Download as pdf or txt
Download as pdf or txt
You are on page 1of 234

AN ANALYTICAL, NUMERICAL AND EXPERIMENTAL INVESTIGATION INTO

THE INTERFACE TEMPERATURE OF SLIDING SURFACES:

NYLON 6,6 AND STEEL

by
ARTHUR LAKES LIBRARY
Paul Panozzo COLORADO SCHOOL OF MINES
GOLDEN, CO 80401
ProQuest Number: 10794294

All rights reserved

INFORMATION TO ALL USERS


The quality of this reproduction is d e p e n d e n t upon the quality of the copy subm itted.

In the unlikely e v e n t that the a u thor did not send a c o m p le te m anuscript


and there are missing pages, these will be noted. Also, if m aterial had to be rem oved,
a n o te will ind ica te the deletion.

uest
ProQuest 10794294

Published by ProQuest LLC (2018). C opyright of the Dissertation is held by the Author.

All rights reserved.


This work is protected against unauthorized copying under Title 17, United States C o d e
M icroform Edition © ProQuest LLC.

ProQuest LLC.
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, Ml 48106 - 1346
A thesis submitted to the Faculty and the Board of Trustees of the Colorado School of

Mines in partial fulfillment of the requirements for the degree of Master of Science

(Engineering Systems).

Golden, Colorado

Date ()'/& '* 1 7

Signed: (
Paul B. Panozzo

Approved: ^
Dr. Andrea Knox-Kelecy
Thesis Advisor

Golden, Colorado
z '. ' ? -
Date f * '

Dr. JoayGosink
Professor and Head,
Department of Engineering
ABSTRACT

A study related to the Pressure x Velocity (PV) Limit of nylon 6,6 was conducted to
improve the prediction of and understand the causes of material failure under varying
loads and operating speeds. The experimental work was conducted using the test setup
and Falex Multispecimen Tester from the ASTM D 3702-90 specification. This
configuration consisted of a rotating polymeric thrust washer mated against a stationary
steel washer. An analytical and numerical study was also conducted to investigate the
accuracy of various analytical and numerical techniques in predicting the experimental
results of the Falex.

Experimental results from the literature and this study suggest that the PV Limit is
dependent upon temperature, geometry and material thermal properties. Tests were run at
the estimated pressure x velocity limits of the material provided by a supplier, but
catastrophic failure was never obtained in our experiments. This was, in part, due to the
fact that the melting temperature of nylon 6,6 was never achieved at the sliding surface.
The reason that the melt temperature was never achieved was that the heat at the interface
was effectively conducted and convected away to the surroundings. One possible reason
for the discrepancy between the results obtained in this study and published data on PV
limit may have been geometry and test method differences. Differences in geometry and
thermal conductivity of not only the steel and polymer thrust washers, but also the
rotating shaft and the support spindle could have been different from that of the test
device of the supplier of the nylon 6,6 data. Differences in geometry result in differences
in conduction, convection and heat transfer, therefore resulting in different surface
temperatures. Convection coefficients also play a determining role in the surface
temperatures: the higher the convective coefficient, the lower the temperatures. Another
factor that could have raised or lowered the measured surface temperatures is the
presence of transfer films.

A finite difference heat conduction code was used to model the geometry of the test
apparatus, while using an average convection coefficient over the whole model to predict
the surface temperatures of the Falex at the sliding surface. The numerical model, which
used a single value of an averaged convection coefficient for all test cases, proved to be a
useful tool for simulating the experimental results and in particular surface temperature.
Simpler analytical techniques, such as one dimensional conduction analysis and one
dimensional fin analysis, proved to be adequate for rough estimated interface
temperatures, but were not accurate enough to use as a predictive tool. It is unlikely that
the one dimensional conduction and fin methods are not accurate throughout a wide range
of heat generation due to the inability to model the test apparatus geometry.

iv
TABLE OF CONTENTS

Page
ABSTRACT ................................................................................ iii

LIST OF FIGURES ................................................................................................... ix

LIST OF TABLES ...................................................................................................... xiv

ACKNOWLEDGEMENTS ........................................................................................... xv

DEDICATION ............................................................................................................ xvi

L INTRODUCTION ................................................................................................. 1

2. POLYMER TRIBOLOGY .................................................................................... 4

2.1 Nylon Tribology ....................................................................................... 4

2.1.1 Nylon ......................................................................................... 4


2.1.2 PV Limit ................................................................................... 5
2.1.3 Temperature Effect on PV Limit ............................................... 7

2.2 Basic Polymer Tribology ............................................ 10

2.2.1 Contact Characteristics of Polymer-on-Metal ............................. 10


2.2.2 Friction, Heat, and Temperature in Sliding Contact..................... 14
2.2.3 Load Effect on Friction ............................................................... 17
2.2.4 Velocity Effect on Friction ......................................................... 18
2.2.5 Wear ........................................................................................... 19
2.2.6 Transfer Film ............................................................................... 22
2.2.7 Thermal Expansion ....................................................................... 24

V
3. HEAT TRANSFER IN TRIBOLOGY 27

3.1 Bulk Modeling ........................................................................................... 27

3.1.1 Surface Temperature Predictions for Moving Heat Source in


Sliding Contact ......................................................................... 27
3.1.2 Flash Temperatures ..................................................................... 33
3.1.3 Transfer/Surface Film ................................................................... 39
3.1.4 Friction and PV Limit ................................................................... 41
3.1.5 Load and Speed ........................................................................... 43
3.1.6 Contact ....................................................................................... 44

3.2 Numerical Modeling .................................................................................. 44

3.2.1 Comparison of the Finite Difference and Finite


Element Methods ........................................................................ 46
3.2.2 Numerical Instability Under HighPeclet Numbers..........................47
3.2.3 Tribology Modeling Assumptionsand Boundary Conditions. . . . 47
3.2.4 Lubricant or Transfer Films in Numerical M odels.........................48
3.2.5 Transient Thermal Analysis ......................................................... 51
3.2.6 Division of Heat Between Contacting Bodies ...............................52
3.2.7 Mesh/Convergence ....................................................................... 53
3.2.8 Contact ......................................................................................... 54

4. EXPERIMENTAL TECHNIQUE .......................................................................... 56

4.1 Falex Setup ................................................................................................ 57


4.2 Material Properties .................................................................................... 61
4.3 Issues of Concern for the Falex Apparatus.... ................................................ 65
4.4 Modifications of Test Apparatus ................................................................ 66

4.4.1 Thermocouple Array ................................................................... 66


4.4.2 Steel Washer Modification ........................................................... 68
4.4.3 Control and Measuring System ..................................................... 70

4.5 Problems Associated with Falex Modifications ..........................................71


4.6 Final Experimental Setup ............................................................................ 72

vi
5. RESULTS AND DISCUSSION OF EXPERIMENTAL WORK 74

5.1 Approach to Experimental Testing and Acquiring D a ta ..............................74

5.1.1 PV Limit Test ............................................................................... 75


5.1.2 Test Matrix ........................................................................... 83
5.1.3 Temperature Measurement Using Nine Thermocouple Array .. 114

6. THERMAL ANALYTICAL AND NUMERICAL SIMULATION OF THE


TEST DEVICE ................................................................................................. 124

6.1 Geometry Effects on Conduction and Convection................................... 125


6.2 One Dimensional Conduction Analytical Calculation............................... 128
6.3 One Dimensional Adapted Fin Calculation ............................................. 131
6.4 Flash Temperatures ............................................................................... 133
6.5 Two Dimensional Conduction using F luent............................................. 133
6.6 Two Dimensional Conduction and Convection Model
Utilizing Fluent ..................................................................................... 134
6.7 Modeling Approach and Model Verification........................................... 136

6.7.1 Convection Holes in Shaft ....................................................... 136


6.7.2 Constant Volumetric Heat Rate at the Sliding Contact Area . . . 138
6.7.3 Average Convection Coefficient over the Surface.................... 141
6.7.4 Ambient Temperature ............................................................. 142
6.7.5 Finite Difference Method ......................................................... 143
6.7.6 Grid Dependence ..................................................................... 146
6.7.7 Convergence of Results and Verification C heck....................... 146

7. RESULTS AND DISCUSSION OF ANALYTICAL AND NUMERICAL


WORK ............................................................................................................... 153

7.1 Modeling Conditions for Analytical and Numerical Results.................... 153

7.1.1 Heat Generation Rate ............................................................... 154


7.1.2 Heat Partitioning ..................................................................... 156
7.1.3 Geometry ................................................................................. 163

7.2 Comparison of Numerical Model to Experimental Data......................... 163


7.3 Comparison of Analytical Surface Temperature Predictions................... 183
7.3.1 Flash Temperature ................................................................. 183
7.3.2 One Dimensional Conduction Analytical Calculation............. 185
7.3.3 One Dimensional Adapted FinCalculation.............................. 188

7.4 Comparison of Analytical and Numerical Surface Temperature


Predictions ........................................................................................... 190
7.5 Numerical Model as a Predictive Tool ................................................... 192

8. CONCLUSIONS AND RECOMMENDATIONS .............................................. 196

REFERENCES ......................................................................................................... 201

APPENDIX A Variations of Tribometer Test Geometry ................................ 205

APPENDIX B Temperature Contours and Axisymmetric Geometry from


Fluent ................................................................................... 210

viii
LIST OF FIGURES

Page
Figure 2.1 Pressure x Velocity Curve of a Self Lubricating Composite Material
(Theberge 1970) 6
Figure 2.2 Actual Maximum Shear Stress Location for a Plastic Contact of a
Cylinder-on-Flat due to Normal Load (Amell et al. 1993).................. 13
Figure 2.3 Comparison of Wear Rate to Sliding Distance (Anderson et al. 1978).. 21

Figure 3.1 Surface Temperature Profiles for a Circular Sliding Source Contact
(Archard 1988) 32
Figure 3.2 Model of Contact Area (Archard 1958) .................................................. 34

Figure 4.1 Falex Multispecimen Test Machine (ASTM 1990)................................. 57


Figure 4.2 Thrust Washer Test Specimen Arrangement (ASTM 1990)................... 58
Figure 4.3 Rotating Test Specimen (Thrust Washer), (ASTM 1990)
(Dimensions are in Inches, Parentheses are in m m ) ...............................59
Figure 4.4 Stationary Steel Washer Specimen (ASTM 1990)
(Dimensions are in Inches, Parentheses are in m m ) ...............................59
Figure 4.5 Thermal Conductivity of Nylon 6,6 at Various Temperatures................. 62
Figure 4.6 Thermomechanical Properties of Nylon 6 ,6 ..............................................64
Figure 4.7 Modified Stationary Steel Specimen (All Dimensions in Inches) 68
Figure 4.8 Radial Placement of Thermocouple in A rra y ............................................70
Figure 4.9 Final Experimental S e tu p .................................... 73

Figure 5.1 Limiting Pressure x Velocity Values ........................................................76


Figure 5.2 PV Limit Test Performed on Tribometer Tester; 133-623 N at
0.0508 m/s (Load = 30-14Ô lb, V = 10 ft/m in)....................................... 77
Figure 5.3 PV Limit Test; 556-1,779 N at 0.0508 m/s
(Load = 125-400 lb, V = 10 ft/min) ....................................................... 79
Figure 5.4 Wear Depth of PV Limit Test; 556-1,779 N at 0.0508 m/s
(Load = 125-400 lb, V = 10 ft/min) ....................................................... 81
Figure 5.5 Original Test Matrix ................................................................................ 84
Figure 5.6 Free Convection for Friction Coefficient and Temperature at Half LNP
PV Limit; 133 N and 0.0508 m/s (Load = 30 lb, V = 10 ft/min) 85

ix
Figure 5.7 Forced Convection for Friction Coefficient and Temperature at Half
LNP PV Limit; 133 N and 0.0508 m/s (Load = 30 lb, V = 10 ft/min).. 85
Figure 5.8 Free Convection for Friction Coefficient and Temperature at Estimated
Half PV Limit; 13 N and 0.508 m/s (Load = 3 lb, V = 100 ft/min). . . . 87
Figure 5.9 Forced Convection for Friction Coefficient and Temperature at Estimated
Half PV Limit; 13 N and 0.508 m/s (Load = 3 lb, V = 100 ft/min). . . . 87
Figure 5.10 Free Convection for Friction Coefficient and Temperature at Estimated
Half PV Limit; 4 N and 1.27 m/s (Load = 1 lb, V = 250 ft/m in) 88
Figure 5.11 Forced Convection for Friction Coefficient and Temperature at Estimated
Half PV Limit; 4 N and 1.27 m/s (Load = 1 lb, V = 250 ft/min) 89
Figure 5.12 Free Convection for Friction Coefficient and Temperature, Preliminary
Testing; 133 N and 0.152 m/s (Load = 30 lb, V = 30 ft/m in).................90
Figure 5.13 Forced Convection for Friction Coefficient and Temperature, Preliminary
Testing; 133 N and 0.152 m/s (Load = 30 lb, V = 30 ft/m in).................91
Figure 5.14 Comparison of Free and Forced Convection at Steady State
Temperatures ....................................................................................... 92
Figure 5.15 Lines of Constant PV for Test Matrix ...................................................... 94
Figure 5.16 Test Matrix: Forced Convection 444 N and 0.0508 m/s
(Load = 1001b, 10 ft/min) ..................................................................... 95
Figure 5.17 Test Matrix: Forced Convection Repeat at 444 N and 0.0508 m/s
(Load = 1001b, 10 ft/min) ..................................................................... 96
Figure 5.18 Test Matrix: Forced Convection at 1,112 N and 0.0508 m/s
(Load = 250 lb, V = 10 ft/min) ............................................................. 97
Figure 5.19 Test Matrix: Forced Convection at 88 N and 0.254 m/s
(Load = 20 lb, V = 50 ft/min) ............................................................... 98
Figure 5.20 Test Matrix: Forced Convection at 222 N and 0.254 m/s
(Load = 50 lb, V = 50 ft/min) ............................................................... 99
Figure 5.21 Test Matrix: Forced Convection at 44 N and 0.508 m/s
(Load = 10 lb, V = 100 ft/min) ......................................................... 101
Figure 5.22 Test Matrix: Forced Convection Repeated at 44 N and 0.508 m/s
(Load = 10 lb, V = 100 ft/min) ......................................................... 101
Figure 5.23 Test Matrix: Forced Convection at 111 N and 0.508 m/s
(Load = 25 lb, V = 100 ft/min) ....................................................... 103
Figure 5.24 Test Matrix: Forced Convection Repeated at 111 N and 0.508 m/s
(Load = 25 lb, V = 100 ft/min) ....................................................... 103
Figure 5.25 Test Matrix: Forced Convection at 44 N and 1.27 m/s
(Load = 10 lb, V = 250 ft/min) ....................................................... 105
Figure 5.26 Test Matrix: Forced Convection Repeated at 44 N and 1.27 m/s
(Load = 10 lb, V = 250 ft/min) ....................................................... 105
High and Low Temperatures, Forced Convection for PV
at 175 (kPa-m/s) ........................................................................... 108
High and Low Temperatures, Forced Convection for PV
at 437 (kPa-m/s) ........................................................................... 108
Measured LVDT Wear at the Min. and Max. Temperature for PV
(175 kPa-m/s) ............................................................................... 110
Measured LVDT Wear at the Min. and Max. Temperature for PV
(437 kPa-m/s) ............................................................................... 110
Comparison of Tribometer Thermocouple and Array Thermocouple
at Same Location (Load = 44 N, V = 1.27 m/s) ............................ 115
Comparison of Tribometer Thermocouple and Array Thermocouple
at Same Location (Load = 444 N, V = 0.0508 m/s) ...................... 116
Comparison of Tribometer and Array Top Row Thermocouples
(Load = 44 N, V = 1.27 m/s) ......................................................... 118
Comparison of Tribometer and Array Top Row Thermocouple
(Load = 444 N, V = 0.0508 m/s) ................................................... 119
Array Temperature Contour at High Temperature
(Load = 222 N, V = 0.254 m/s) ..................................................... 121
Array Temperature Contour at Low Temperature
(Load = 444 N, V = 0.0508 m/s) ................................................... 122

Initial Geometry ............................................................................... 126


Second Geometry Modeled ............................................................. 127
Tribometer Convection Holes and Fluent Axisymmetric
Representation ............................................................................... 138
Pressure Distribution from Finite Element Analysis......................... 140
Two Dimensional Nodal System (Incropera and DeWitt 1990)........ 144
Finite Difference Approximation (Incropera and DeWitt 1990). . . . 145
Dense Mesh Used in Grid Dependence Studies
(shown in Sliding Materials Only) ...................................... .......... 148
Mesh Used in Numerically Simulated Results
(Shown in Sliding Materials Only) ................................................. 149

Volumetric Heat Generation Rate Convergence


(Temperature (K) vs. Length (m)) ................................................. 155
Increased Area of Volumetric Heat Generation Rate
(Temperature (K) vs. Length (m)) ................................................. 156
Approximation of Asperities in Contact Area ................................ 157
Numerical Model of Small and Middle Geometries ........................ 164
Numerical Geometry Comparison with Conduction O nly................ 165

xi
Numerical Model Geometry Comparison for Conduction
and Free Convection 166
Numerical Model Geometry Comparison for Conduction
and Forced Convection ................................................................... 167
Numerical Model Boundary Condition Comparison for Actual
Geometry Representation ............................................................... 168
Comparison of Numerical Model and Experimental Results for
Forced Convection ......................................................................... 169
Rough Best Fit Estimate of Convection Coefficient for Forced
Convection Using the Thermocouple Array
(Load = 1112 N, V = 0.0508 m/s) ................................................... 172
Refined Best Fit of Convection Coefficient for Forced Convection
Using the Thermocouple Array (Load = 1112 N, V = 0.0508 m/s) . . 173
Comparison of hexpenmentai and hmax/min for Forced Convection............ 175
Comparison of the Experimental and Average Convection
Coefficient for Forced Convection ................................................. 176
Comparison of Top Row Array Thermocouple and Numerical
Models Temperature ....................................................................... 177
Comparison of Tribometer Thermocouple and Numerical Model
Temperature ................................................................................... 178
Comparison of Maximum Array Temperature and Numerical Model
Surface Temperature ....................................................................... 179
Numerical Model Surface Temperature for haemal and leverage.............. 180
Comparison of Best Fit Forced Convection Coefficient and Speed. . . 181
Comparison of Best Fit Forced Convection Coefficient and
Normal Load ................................................................................... 182
Comparison of Best Fit Forced Convection Coefficient and
Temperature ................................................................................... 182
Fin Geometry Comparisons to Numerical Surface Temperature........ 189
Comparisons of Analytical and Numerical R esults............................ 191
Comparison of Surface Temperature and Thermal Conductivity
for Polymer-on-Material ................................................................. 193
Numerical Model as a Predictive Tool for a Polymer-on-Material .. . 194

Front View of Tribometer Tester (Not to Scale) .............................. 206


Axisymmetric Representation of Tribometer T e ste r.......................... 207
Axisymmetric Representation of Tribometer Tester with Bearings,
Cooling Holes and Load Plate ................ ........................................ 208
Axisymmetric Representation of Tribometer Tester with Cooling
Holes and Load Plate ....................................................................... 209

xii
Figure B. 1 Temperature Contour at Experimental Maximum Heat Generation for
Geometry with Bearings, Cooling Holes and Load Plate.................. 211
Figure B.2 Temperature Contour at Experimental Maximum Heat Generation for
Geometry with Load Plate ............................................................... 212
Figure B.3 Temperature Contour at Experimental Maximum Heat Generation for
Geometry with Cooling Holes and Extended Load Plate ................ 213
Figure B.4 Temperature Contour at Experimental Maximum Heat Generation for
Actual Geometry ............................................................................. 214
Figure B.5 Close Up Temperature Contour of Contacting Surfaces at
Experimental Maximum Heat Generation for Actual Geometry . . . . 215
Figure B.6 Temperature Contour at Melting Temperature of Nylon 6,6 for
Actual Geometry ............................................................................. 216
Figure B.7 Close Up Temperature Contour of Contacting Surfaces at Melting
Temperature of Nylon 6,6 for Actual G eom etry............................... 217
LIST OF TABLES

Page
Test Matrix of Load, Speed and PV ................................................. 93
Test Matrix: Temperature (°C), Coefficient of Friction and
PV (kPa-m/s) ....................................................................................... 107
Different Methods of Calculating the Wear Rate for the Test Matrix
(A = 175 kPa-m/s and B = 437 kPa-m/s) ............................................. 113
Comparison of Tribometer and Array at Same Location
Thermocouple Temperatures (°C) ....................................................... 117
Comparison of Tribometer and Array Top Row Thermocouple
Temperatures ....................................................................................... 120

Heat Partitioning Values ................................................................. 163


Least Squares Fit of Temperature for Forced Convection
Coefficients to Experimental Data ................................................. 170
Averaged Convection Coefficients (W/m2*K) and PV (kPa-m/s)
for Forced Convection ......................................................................... 174
Calculated Flash Temperature Above the Bulk Surface Temperature
for 5 Contacts Across Contact Area ..................................................... 184
Thermocouple Temperature for Test Matrix of Tribometer
and Array Top Row Center (°C).... ......................................................... 185
Calculated Bulk Surface Temperatures from the Tribometer and
Top Row Array Thermocouple (°C) ..................................................... 187
Infinite Fin Calculations of Bulk Surface Temperatures (°C) for
Three Geometries ................................................................................. 190

xiv
ACKNOWLEDGEMENTS

Special thanks goes to Dr. Andrea Knox-Kelecy for taking me on as her student when
she really had nothing to gain from this work. Honest, trusting people like Dr. Knox-
Kelecy are not looking out for themselves only, and they serve a great duty to society and
educational facilities such as Colorado School of Mines. Andrea always believed in me
and my abilities and she always encouraged my decisions.

Also, special thanks goes out to Dr. Ron Rorrer for making this happen. His drive,
ideas and equipment made this become a reality. Dr. Rorrer’s introduction of the “Zen
Theory” to me says that no matter what you do, if you care about it or not, do it well
because it is a reflection of your character. This statement will carry on with me the rest
of my life, no matter what path life has chosen for me.

It would not be right if I did not mention the people who work in the surface science
lab at Gate Rubber Company. Thanks to Jesse, J.J. and Kate for helping me with all
aspects of my project and for working around my schedule to allow me to use the Falex. I
also appreciate the technical inputs of my third committee member Dr. Dave Munoz.

XV
DEDICATION

This work is dedicated to all whom were supportive in my decision to work towards
my Master of Science. Those people include my parents, brothers, sisters and friends,
their support made this happen.

xvi
1

INTRODUCTION

In today’s modem world the need to reduce weight, maintain strength, improve
efficiency and reduce cost are problems that face companies everyday. One way of
accomplishing these goals is to replace metals with plastics. Plastics are light in weight,
can be high in strength and have low friction, all factors which may help to improve
efficiency. Plastics can also provide a means for inexpensive material costs. Plastics are
being used more in heavy load, high speed and high temperature sliding conditions.
Plastics can be made self lubricating, which will lower cost, wear, and reduce damaging
debris.

Understanding and enhancing plastics under loads, velocities, friction and temperature
conditions in tribological systems will allow for better designs and greater use. Many
analytical (Blok 1963 and Archard 1958), numerical (Kennedy 1981 and Vick et al. 1991)
and experimental (Lancaster 1971 and Ludema 1980) studies have been conducted for
various polymer materials and sliding situations. These studies used a variety of sliding
test conditions, and in general the results have not been compared and contrasted across
methods due to the complexities of their differences. It is not clear which method best
represents the sliding system. Therefore, this study was conducted to show the validity of
analytical, numerical and experimental work performed for one particular test
configuration.

The objective of this study is to develop a cost effective, simple and useful tool for
estimating sliding surface temperatures, which may be either analytical, numerical or
2

experimental. The focus of the work became the development of a simple numerical
model that would accurately simulate thermal conditions within the Falex Multispecimen
Tester and the use of that model as a predictive tool. Before the model could be used as a
predictive tool, it required the use of experimental data to estimate convective heat
transfer coefficients. This allowed for the best approximation surface temperatures and
these surface temperatures were used to compare and contrast with analytical models.

The test matrix used in this study was chosen to cover multiple speeds and loads,
while obtaining one of two different values of Pressure x Velocity (PV). Nylon 6,6 was
the material chosen for this study because it had been studied many times previously
(Vroegop et al. 1980, Ludema 1980 and Lancaster 1971), and therefore it would be easy
to compare and contrast prior work with the work performed here. The test matrix
yielded a good range of temperatures, coefficient of friction (|n) values and wear, all
factors which had the most impact on the softer of the two rubbing surfaces. In the
experimental work performed the polymer is softer than steel. Polymers can be broken
down by a number of factors or combination of factors such as: load, velocity, coefficient
of friction and temperature. Most of the time it can be seen that an increase in load
increases the coefficient of friction, wear and temperature. An increase in velocity
without an increase in load tends to decrease the coefficient of friction, therefore
decreasing wear and temperature. The formation of transfer films will occur under
increasing velocity and low to moderate loads. The coefficient of friction is a function of
load, velocity and temperature. Temperature is a function of load or velocity or some
combination of the two. It is believed in this study that temperature is the driving factor
in the failure of the material.

The following chapters elaborate on the factors surrounding this study. In Chapter 2,
titled POLYMER TRIBOLOGY, a review of the tribology and the characteristics of
3

nylon 6,6 in particular are presented. In Chapter 3, HEAT TRANSFER IN


TRIBOLOGY, certain topics pertaining to heat transfer are addressed, with a focus on
analytical methods to predict and understand bulk temperatures and flash temperature.
Chapters 4 and 5, EXPERIMENTAL TECHNIQUE and the RESULTS AND
DISCUSSION OF EXPERIMENTAL WORK, cover the experimental test apparatus and
the experimental results that followed the testing. In Chapter 6 and 7, THERMAL
ANALYTICAL AND NUMERICAL SIMULATION OF THE TEST DEVICE and the
RESULTS AND DISCUSSION OF THE ANALYTICAL AND NUMERICAL WORK,
the numerical modeling approach and results are presented and discussed along with the
analytical work performed. The final topic, covered in Chapter 8, is the CONCLUSIONS
AND RECOMMENDATIONS, which is composed of a summary and recommendations
made on improving the experimental and numerical work.
4

POLYMER TRIBOLOGY

2.1 Nvlon Tribology

Tribology is the study of friction, wear and lubrication. The polymer nylon is an
important tribological material due to its good wear characteristics as well as its low
weight, low cost and robustness. A review of the tribology of polymeric materials, and in
particular nylon, will be covered in this chapter along with the an investigation into the
factors that eventually lead to Pressure Velocity Limit.

2.1.1 Nvlon

Nylon is the trade name associated with polyamide. Polyamide resins (nylons) are
synthetic polymers that contain an amide group, -CONH-, as a recurring part of the chain.
Nylons are commonly identified by numbers corresponding to the number of carbon
atoms in the monomers. One of the most common commercial products is nylon 6,6.
Nylon 6,6 is the polymer used throughout this study. Some of the key characteristics of
nylon is its resistance to fatigue, repeated impact, and abrasion. Other characteristics are
a low coefficient of friction, high tensile strength and toughness, creep resistance, and
retention of properties over a wide temperature range. The melting point of nylon 6,6 is
269 °C (LNP Engineering Plastics).
5

2.1.2 PV Limit

In sliding systems, both the contact pressure and the relative siding velocity affect
wear. The Pressure Velocity Limit, better known as PV limit, is the value of normal
pressure (across the contact area) x velocity (of sliding contact) at which catastrophic
tribological failure of a plastic occurs due to the resultant temperature, friction and wear.
Catastrophic failure of a polymer shows chemical decomposition to a point where the
material flows away due to melting or it discentigrates. The PV limit is the critical value
above which the sliding system no longer performs adequately. It has been shown
(Theberge 1970) that this critical value of PV will vary with pressure and velocity. At
zero velocity it is believed that there is a critical pressure at which the material flows due
to the normal load alone (cold flow). Similarly, when the load is virtually zero there is a
velocity at which the material melts due to frictional heat generation alone. Thus at
different combinations of non-zero pressure and velocity, the behavior of the material that
leads to failure will be a combination of temperature-related and mechanical failure. The
PV limit is typically represented by a curve that is fitted to these limits of pressure and
velocity (See Figure 2.1).
6

10
Vel oci t y, Z ( f p m )

Figure 2.1 Pressure x Velocity Curve of a Self Lubricating


Composite Material (Theberge 1970)

While the rationalization that variations in pressure and velocity result in changes of
wear, friction and temperature does not completely explain what happens at given
catastrophic values of PV, it begins to give physical understanding of the underlying
phenomena.

PV limit can be defined in several ways. One is that the limiting PV values are those
at which the temperature at the surface of a sliding contact reaches a value near the
melting or softening point of the polymer. Theoretically this is believed to be the flash
temperature at the asperity contacts superposed on the mean surface temperature resulting
from the general dissipation of frictional heat. In this definition, wear is also believed to
contribute to failure as well as temperature. This concept will be explained in chapter 3
(Lancaster 1971). Another definition of PV limit is that point in which wear is at a
minimum until the melting temperature is obtained. This disallows the possibility that
7

catastrophic tribological failure was a result in excessive wear prior to the melting
temperature.

The published PV limits are not always comparable, particularly if there are
differences in the test apparatus used to obtain the PV values. PV limits also do not tend
to be reflective of other material properties such as melt temperature. Two different
materials with the same melt temperature, will not necessarily have the same PV limits.
Ludema (1980) showed that the melting points for a group of polymers does not follow
the same order as their PV limits. For example, Acetal and tetraflouroethylene resin have
the same melt temperatures, but their PV limit values are 3000 (psi-ft/min) and 1800 (psi-
ft/min) respectively. What Ludema did not take into consideration was that PV limit is
not solely dependent on the melt temperature. Acetal and tetraflouroethylene resin have
different thermal conductivités and coefficient of friction values, and these influence the
surface temperature. The higher the thermal conductivity of a material the quicker heat is
pulled away from the surface, thus lowering the temperature. The higher the coefficient
of friction (|i) is, the higher the heat generation (Q) is if the load (W) and speed (V) are
held constant (Q = pWV). Therefore, the PV of the two materials may differ even though
their melt temperatures are the same.

2.1.3 Temperature Effect on PV Limit

In sliding systems, frictional heat generation results in an increase in the temperatures


of the mating materials. Whenever one of the sliding bodies has some capacity to remove
heat away from the interface, but if thermal conductivity of the conducting material is
relatively small, there will be a significant temperature difference between the interface
and the conducting material. The surface will then have a higher macroscopic (bulk)
temperature (Ling and Pu 1964). The increase in temperature affects the material
properties and thus friction and wear. As the temperature of operation approaches the
material transitions, such as a glass transition, the mode of wear can change as well as the
magnitude of friction. In fact, as the melting temperature is approached at the surface,
catastrophic failure or seizure can ensue. The effects of temperature at the sliding surface
interface is the topic of interest, yet they are not easily obtained. There is not an exact
way of measuring the interface temperature for most sliding systems, only an estimate
measurement of the surface temperature. Thus, it is useful instead to predict the
interfacial temperature.

Attempts by Blok (1963) and Archard (1958) have been made to predict flash surface
temperatures both experimentally and analytically using the conduction heat equation, but
the results seldom agree. Furey (1964) believes that through a more rigorous application
of the energy equation, a more realistic and accurate approximation of the interface
temperature can be obtained analytically. The differences between the theoretical and
experimental results of Blok (1963) and Archard (1958) are believed by Furey (1964) to
be due to heat loss by radiation, convection or melting of the materials.

Of the published research, there has been disagreement whether bulk or flash
temperature at the interface causes melting of the polymeric material. Bulk temperatures
are a macroscopic approach to estimating surface temperatures based on ideally smooth
surfaces and apparent area. Flash temperatures are based on real area and localized
temperature rises due to increased loads at asperities in contact. Vroegop, Vermeulen and
Bosma (1980) argue that PV failure (limit) occurs when the temperature at the surface
reaches a value slightly above the glass transition temperature, independent of friction
and speed. Others such as Ettles (1987), believe that a PV failure will depend upon load
and speed, because these variables impact frictional heating, which is the basis of
9

catastrophic failure in a polymeric material. In fact, it is still a matter of dispute whether


interface temperature actually controls the process of friction and wear. One problem in
understanding the PV limit is the long amount of testing time it takes to achieve failure in
a material. After such long durations of run time, it is not clear whether failure is due to
severe wear after extended periods of use or due to the temperature-related factors
mentioned earlier.

Furey’s (1964) experimental work showed that the surface temperature variations with
time were small. Since wear is occurring throughout the test, the location of the contact
areas is also changing with the wear. These results suggest that the temperature rise is not
a function of the gross wear area but rather the contact area. As contact area increases,
surface temperature does as well, but in Furey’s (1964) case the contact area remained
constant over time. Wear occurs no matter what the run time is, but it will be a smaller
factor in determining the PV limit if run times are kept short.

Despite all the inconsistencies, PV limits serve a useful purpose in the preliminary
design stages of wear components using plastics. PV limit data allow a designer to rank
one plastic relative to another for a range of geometries and severe operating conditions
(Ludema 1980). It is clear that size, shape and thermal properties of the counterface are
as important as the mechanical properties of the polymer. It is well known that the
generation of heat at sliding contact of polymers can produce local decomposition, hence
limiting the usefulness of polymers in a sliding application (Ettles 1987).
10

2.2 Basic Polymer Tribology

Basic polymer tribology is a general approach to polymers and not limited to nylon
6,6. The contact characteristics for polymer-on-metal is explained in this section along
with the effects of temperature, friction, velocity, wear, transfer films and thermal
expansion have on polymer sliding systems.

2.2.1 Contact Characteristics of Polvmer-on-Metal

In the sliding system of a polymer on a non-polymeric solid, the surface contact, which
depends upon the surface roughness, appears to be its most important property in
affecting PV limit (Vroegop, Vermeulen and Bosma 1980). The area of contact can be
deformed under load either plastically or elastically. In plastic deformation a material
does not go back to its original form upon release of contact loads. In this case, the
contact area for a sphere-on-sphere can be described as (Archard 1958),

a= [2.1]

where:
a = radius of circular area of contact [mm]
W = normal load [N]
pm= flow pressure [N/mm]

In elastic deformation the material returns to its original form. In this case, the contact
area for a sphere-on-sphere is (Archard 1958),
11

a = 1.1 — [2.2]

where:
a = radius of circular area of contact [mm]
W = normal load [N]

R; — = — + — , (undeformed radius of curvature of protuberance) [mm]


R R j R2

1 1 — D .2 1 — Vj
E; — = —- — + —- — , (Young’s modulus) [MPa]
E \ 2

D = Poisson’s ratio

Green wood-Williamson development of the plasticity index, which includes both


material and topographic properties of the solids in contact, can be used to determine if
the contact is elastic or plastic (Greenwood and Williamson 1966).

The use of the laws of plastic and elastic deformation to calculate true contact areas
may not be valid for microscopically rough, dynamic systems. Macroscopically, the
contact area appears to be smooth between the two counterfaces, and this area is called
the apparent area. On a microscopic scale the surfaces are not smooth and the contact
area is much less than that of the apparent area. This microscopic area is called the real
area. Ettles (1987) believes the main difficulty in applying theories of surface heating to
practical cases is in defining the area of the contact. When considering the ideal case of
the .apparent area, the entire surface area is assumed to be available to conduct heat away
from the surface and the load is assumed to be uniformly distributed so there are no
isolated hot spots. However, since the real area, assumed by Ettles (1987), is smaller than
the apparent area, the smaller path for conduction and the higher local load would cause

ARTHUR LAKES LIBRARY


COLORADO SCHOOL OF MINIS
GOLDEN, CO 80401
12

the surface temperature to be higher than would be expected based on the apparent area.
Once contact is made between two pieces, temperature and wear cause the contact area to
change. This area change is very difficult to model. Although Ettles (1987) assumed real
area would be less than apparent area, it is possible that the real contact area is somewhat
greater than expected because of the heating of the material and corresponding decrease
in the flow pressure, pm(Furey 1964).

Archard (1958) developed both the plastic and elastic cases of Hertzian contact, which
is a parabolic pressure distribution over the contact area, and he also modeled a contact as
a single asperity at a flow stress, usually set as the penetration hardness. Lancaster (1971)
states that Jaeger’s analysis where asperity contacts are modeled, shows temperatures will
tend to be higher if the deformation is plastic. This is because the real contact area is
smaller for plastic than it is for elastic deformation, thus the assumption of plastic contact
will tend to over estimate the real temperatures. The plastic contacts are higher in
temperature due to less contact area and higher concentrated loads at each contact causing
more heat generation at that point.

Kennedy (1984) states that a large majority of the frictional heat energy is transformed
into thermal energy. This statement was based on his analysis of past studies, which have
shown that approximately 95% of the energy expended in plastic deformation process is
dissipated as heat, and that the percentage increases as either the strain increases or the
temperature rises. The remaining approximate 5% of energy is dissipated as radiation
through the removal of wear debris and material chemical decomposition at melting.
Since a metallic material’s hardness and shear strength decrease with increasing
temperature, the material just beneath the surface could become weaker than the material
above it and deform more easily (Kennedy 1984). This can be seen in Figure 2.2 which
13

indicates that the maximum shear stress (Tmax) occurs at a distance of 0.79a below the
surface, where “a” is the radius of the contact area.

Lmes of
consioni
r moM

Figure 2.2 Maximum Shear Stress Location for a Plastic Contact of a


Cylinder-on-Flat due to Normal Load (Amell et al. 1993)

When two different materials with different hardnesses are in contact, determining the
type of contact as plastic or elastic is complex. It was determined in the experimental
work done for this study described in Chapter 5 that wear causes difficulty in determining
the contact type. The harder material’s contact points (asperities) tend to cause wear
scars in the surface of the softer material. In order to get a better understanding of the
surface contact concept, Ettles et al. (1994) used three different methods for estimating
area in trying to predict surface heating. The first method used a measured wear scar as
the contact area. The second method used a Hertzian contact area, and the third method
estimated contact area assuming a plastic contact and a flow hardness pressure of 7500
MPa. Of the three methods used, the best agreement with experimental results was found
14

the contact area. The second method used a Hertzian contact area, and the third method
estimated contact area assuming a plastic contact and a flow hardness pressure of 7500
MPa. Of the three methods used, the best agreement with experimental results was found
using the wear scar. Ettles states that results using the Hertzian equation would have
been satisfactory in the absence of wear. Under wear conditions, the Hertzian equation
showed a maximum error of 25%.

2.2.2 Friction. Heat, and Temperature in Sliding Contact

In general, frictional heat is not generated in the space between the asperities of two
bodies in sliding contact. Most of the heat is generated on the surface and in the surface
layer immediately below the surface of both bodies. As was shown in Figure 2.2, the
maximum shear stresses have been found experimentally to occur below the surface, thus
higher surface temperatures at 0.79a below the surface. Ling and Pu (1964) believes that
this high stress region is the region of frictional heat generation due to either the breaking
of adhered junctions or the thermodynamically irreversible process of plastic deformation
of asperities and the bulk body. Once the surface temperature reaches a point of melting
for one of the materials, it would be expected that the coefficient of friction would
decrease. Ludema (1980) suggests, however, that there is not the sharp decrease in the
coefficient of friction when severe wear occurs that one might expect if molten species
were to suddenly appear in the contact region. This could be because the asperities of the
non-melting material are sinking farther into the molten material until they hit material
that is not molten, thus causing the friction to remain constant.

The low thermal conductivity of polymers and elastomers dictate that the frictional
heat is not easily conducted from the interface. Ettles (1987) states that the limiting PV
15

will occur at a load and velocity, beyond which frictional heating is actually the dominant
factor. This is the regime that Ettles (1987) describes as the thermal control regime. The
thermal control regime is obtained when friction is governed by heat flow, which is at a
maximum allowable surface temperature. Ettles’s (1987) work showed that about 75% of
the dry friction was attributable to adhesion and 25% to deformation friction. He also
implies that in the thermal control regime, the thermal properties of the counterface have
an important influence on friction. In Ettles’s (1987) testing of materials, he has shown
some very different results in sliding a material A against a material B, as opposed to
material B against material A. A possible explanation of this thermal phenomena is that
the dominant sink is reversed in the two cases. If the coefficient of friction and the
thermal resistances remain constant, the interface temperature is directly proportional to
PV.

The reduction of the friction coefficient with increasing load is a well-known feature
of polymers and occurs in conditions where the friction is not thermally controlled.
Archard has shown both experimentally and analytically for slow sliding speeds that the
friction coefficient, |i, is approximately proportional to V 20 and W '25, where V and W
are speed and load (Ettles 1987). Archard states that friction reduction due to increasing
load under constant speed occurs in conditions where the friction is not thermally
controlled (Ettles 1987). He has also shown both experimentally and theoretically for
treaded and smooth tires that at a slow sliding speed a reduction in the coefficient of
friction with increasing load can be seen in the range j l l proportional to W"25 for partial
contact (treaded tires) or p proportional to W 1 for full contact (smooth tires). A
distinguishing feature between friction at low speeds and thermally controlled friction is
that the onset of thermal control should be clearly apparent as the load is increased at
constant speed (Ettles 1987). Tanaka and Yamada (1984) have a similar view to
Archard’s theory that load is proportional to a raised exponent of the coefficient of
16

friction. They state that the coefficient of friction generally decreases with the increase in
load “W” and is proportional to W'n. Although the index “n” is generally very small, its
value is dependent upon the type of polymer and the temperature. The small value of the
index “n” indicates that the contact between the polymer and the steel washer is
approximately plastic under the experimental condition (Tanaka and Yamada 1984). The
index “n” appears to be in the range of approximately -0.25 to -1.0, which is the same as
Archard’s case for partial and full contact, where Tanaka and Yamada’s (1984) work
depends on the polymeric material and the temperature.

According to Lancaster (1971) the coefficient of friction is not directly influenced


by temperature and load but is indirectly influenced through the change in speed. The
work done by Lancaster (1971) was mainly experimental and focused on the tribological
effects of polymer-on-metal. Vroegop et al. (1980) believe that in a nylon-on-steel
configuration friction is controlled by the mean surface temperature of the steel and is
independent of sliding velocity. They also stated that the coefficient of friction
dependence on temperature shows a reproducible behavior. This work done by Vroegop
et al. showed that in a nylon-on-steel configuration, the friction is approximately linearly
dependent upon the mean temperature of the steel surface for temperatures above 77°C
(350 K) and below 47°C (320 K). Between 47 and 77°C, there appears to be a transition
in frictional behavior. The slope of the curve of friction versus temperature shows a
discontinuity at a temperature of 64°C (337 K), approximately 7°C above the glass
transition of dry nylon 6,6. The slope of the curves above the transition temperature
increases with speed and varies with center line average (c.l.a.) roughness of the steel
counterface. Below the transition, no correlation with speed or surface roughness is
observed. Vroegop et al. point out that a lower coefficient of friction will appear above
110°C in air for nylon 6,6. Presumably this is the result of oxidative degradation of nylon
17

6,6. In nitrogen, the friction rises until the temperature reaches the melting point of nylon
6,6 (== 270°C).

Experimental results of Jaeger (1942) and Ludema (1980) have shown a decrease of
friction with increasing speed. However, the friction of some materials tend to rise
slightly with further increase in velocity. Tanaka and Yamada (1984) suggest that as
friction of polymers is accompanied with local melting, the surface softens and the
asperities sink farther into the material thereby causing a larger contact area and higher
friction. Jaeger (1942) states that once thermal control has been initiated, the coefficient
of friction (g) is closely proportional to V‘1/2. He states further that the onset of thermal
control is clear for each material and occurs sequentially with the approximate softening
temperatures. Ettles (1987) states that the coefficient of friction for nylon-on-steel is
strongly affected by a transfer layer, thus the thermal control regime will not occur until a
transfer film has developed.

2.2.3 Load Effect on Friction

Kennedy (1984) observed several studies, which involved a moving spherical heat
source. He concluded that the contact pressure reaches a maximum near the trailing edge
of the contact and that increases in load at constant velocity tend to flatten the contact
pressure and heat generation distributions. Furber et al. (1976) believe that load is a more
useful parameter than pressure, based on their testing done with wear pins. When testing
with a truncated cone style wear pin, the apparent and real areas of contact are continually
changing and therefore the pressure is as well, whereas the load remains constant.
Lancaster (1971) has also remarked that in accelerated wear rates it is important to
18

simulate the load rather than the nominal pressure, as the volumetric rate of wear is load
dependent.

When nylon wears, small separated fragments are produced. The small size of the
nylon fragments and the values of initial roughness of the steel specimens (always lower
than 0.45 pm), suggests that, if the load is lower than or equal to that to which the
maximum coefficient of friction corresponds, wear is of the adhesive type. For higher
loads, other aspects of the worn nylon specimens have been observed: the contacting
surfaces show a characteristic series of cracks perpendicular to the sliding direction and
the wear particles are very large (Kennedy 1984).

2.2.4 Velocity Effect on Friction

Wantabe and Yamaguchi’s (1986) investigated the friction and wear properties of
nylon using three continuous sliding nylon pins on steel tester. Their results of wear and
friction dependence upon temperature showed wear was at a minimum, while coefficient
of friction was at a maximum at nearly the same temperature. They believed that the
resistance to surface flow represented by the shear strength of nylon, becomes larger as
the sliding velocity increases because nylon is a viscoelastic material. Eventually, the
shear strength will peak, thus the friction will peak, due to viscoelastic effects. A large
maximum friction due to the viscoelastic property of the amorphous nylon was observed
for sliding at 10 mm/s. In contrast, for sliding at 0.1 mm/s, the sliding occurred in the
nylon itself, with the local flow of material taking place on the nylon surface. The
maximum friction due to the amorphous state was not marked because the sliding took
place in the molecularly oriented flow layer of the nylon surface (Wantabe et al. 1986). It
is conceivable that when specimens were slid at 10 mm/s and around 100°C the shear
19

strength of the nylon itself was higher than that of the interface and sliding occurred
mainly at the interface of the specimens.

2.2.5 Wear

Wear rate is characterized as the amount of material removed in a specific amount of


time. The wear rate is affected by the material finish, hardness of the rubbing contact
surface, temperature and environment (e.g. relative humidity). Severe wear rates can be
more than 2000 times the mild wear rate based on these factors and is also controlled by
the behavior of the transfer film laid down by a plastic body on a metal surface (Ludema
1980). The ASTM (1990) specification for the wear rate of materials states that 6.1
x 10'4 cm/h (1 x 10*5 in/h) is considered excessive; therefore, mild wear rate would be
somewhat less than ÎO 4 cm/h and severe wear rate would be somewhat greater than ÎO 4
cm/h, such as that seen near the melting temperature.

In adhesive wear the creation of an adhesive junction that is stronger in shear at the
interface than the shear strength of the polymer is formed. As a result, a thin, flake-like
debris is created from the surface (Clerico 1969). For nylon 6,6, adhesive wear
predominates at loads lower than or equal to those at which the maximum coefficient of
friction occurs. At higher loads, cracks perpendicular to the sliding direction have been
observed. The wear of nylon 6,6 increases with increasing load, as expected but the wear
rate is independent of load. This is comparable to results obtained by Lancaster for nylon
6,6 against mild steel with a surface finish of 0.15 pm c.l.a. (Furber et al. 1976). In
adhesive wear, a transfer film will be formed on the surface of the steel washer. For the
transfer film to remain intact, it must have good ductility and a low coefficient of friction,
otherwise it will be abraded off (LNP Engineering Plastics). Once the transfer film has
20

been laid, the contact area acts as nylon-on-nylon instead of nylon-on-steel, thus reducing
wear.

Vroegop et al. (1980) state that two different processes of wear particle formation can
occur, depending on the machining of the steel counterface. These different processes are
characterized by a different behavior of the friction force dependence on temperature. On
surfaces turned on a lathe, wear occurs mainly by roll formation. This is due to the spiral
pattern turned on the lathe from the back and forth motion. On a uniformly turned and
subsequently abraded surfaces wear particles mainly have the form of thin flakes, because
the abraded grooves are more likely to be perpendicular to the axis. An increase in speed
corresponds to an increase in the rate of strain of the asperity contacts, and this will, in
turn, lead to a reduction in the elongation to break (Lancaster 1971).

Lancaster (1971) presented wear-speed curves that showed a characteristic shape in


which a maximum wear rate is followed by a minimum just before the flash temperature
reaches the softening or melting point of the polymer. He also noted a reduction in wear
rate prior to softening have also been observed during wear measurements at elevated
ambient temperatures near the melting point, where frictional heating was negligible.
Wantabe and Yamaguchi (1986) state that, from the temperature dependence curve for
wear, one of the causes of the minimum value obtained may have been that sliding takes
place at the interface in the middle (~ 100°C) temperature range and in the nylon itself in
both the low (== 25°C) and the high (= 150°C) temperature ranges. Such decreases in
wear occur because wear rates of polymers are directly related to the breaking strength (S)
and the elongation of the break (e) by 1/Se. Moderate increases in temperature increase e
to an extent which more than compensate for the loss in strength. In abrasive wear,
unfilled polymers have superior wear when tensile stress and strain to failure (toughness)
are high (LNP Engineering Plastics).
21

linear with sliding distance, except during the initial transient wearing process, as shown
in Figure 2.3. They also showed that friction and wear are controlled by the mean surface
temperature of the steel and are independent of sliding distance.

LOAD 31 0 8 N
200.
SURFACE TEMP 7 0 °C

ISO.

E
E

C
E IOQ

<0

1 so.

sliding distance metre* «io3

Figure 2.3 Comparison of Wear Rate to Sliding Distance (Anderson et al. 1978)

According to Archard (1958), for dry sliding surfaces, wear can be analytically
calculated as,

W=KFVT [2.3]

where:
W = wear volume [m3]
22

F = pressure [N/m2]
T = elapsed time [h]
K = wear factor [cm4min/m N hr]
V = velocity [cm/min]

Clerico (1969) has two subgroups he uses to categorize wear. There is a region close
to the interface where the frictional work is dissipated in a very thin layer. Here the rates
of energy dissipation are high and extensive thermal and chemical degradation may occur.
In addition, large strains may cause rupture of the polymer close to the interface and so
produce a transferred layer of polymer on the counterface, provided that the transfer
process is not hindered by the presence of a boundary film. The second dissipation region
includes a greater volume of the interface zone and is characterized by much lower rates
of energy dissipation. In this subgroup, wear properties that are primarily a function of
the cohesive characteristic of the polymers, thus transfer film characteristics can be
included. The polymer can fail by propagation of subsurface cracks if the transfer film is
removed.

2.2.6 Transfer Films

Tanaka and Yamada’s (1984) work shows that the initial low friction at low
temperature seems to be due to the surface roughness of the counterface, but that friction
increases with the amount of transferred material due to the increase of real area of
contact. This appears to explain the fact that the initial friction is lower than the steady-
state friction at low temperatures, because the softer material is not soft enough at low
temperatures for the counterface’s asperities to sink into it. However, the real area of
contact corresponding to the initial friction increases with temperature become of the
23

thermal softening of the polymer. The metal surface asperities penetrate the polymer
surface more deeply at temperatures near the glass transition temperature. Therefore, the
increase of initial contact area and of penetration of the materials surface asperities cause
very high initial friction at temperatures above the transition temperature. However, the
thickness of the transferred polymer layer increases with sliding distance and the sliding
of the polymer begins to occur on the thick and softened transferred polymer layer at high
temperatures, resulting in lower steady-state friction. Experimental data from the work of
Furber et al. (1976) with filled and unfilled nylons show, by contrast, that the transfer film
increases the surface roughness in all cases. Therefore, the transfer film would have a
detrimental effect on the wearing properties, since the mating couple is changed from
polymer/metal to polymer/polymer. Furber et al. dealt with higher temperatures where
the transferred polymer layer did not increase in thickness with sliding distance; therefore,
the surface asperities penetrated into the softened polymer.

Ludema (1980) found from experimentation with a moving heat source sliding system,
that although the sliding surface temperature was appreciably higher than the crystalline
melting point and the softening point of the plastic (nylon 6,6), and actually reached the
thermally degradation temperature, no measurable wear occurred until destruction of the
transfer film. The destruction occurred when the steel surface reached a temperature in
excess of 50°C above the softening point of the plastic tested. Catastrophic loss of
material from the sliding surface occurred upon destruction of the transfer film. If the
temperature of the transfer film is high such that the film of plastic has low viscosity and
if the molten plastic does not wet the metal surface, the plastic agglomerates into spheres
which are removed by the next slider to pass by. Thus, the sliders are deprived of a
lubricant film and instead yield up some material to establish a new film which again is
quickly detached.
24

2.2.7 Thermal Expansion

The process of thermal expansion can be described as follows. Non-uniformities in


the contact pressure distribution result in more frictional heating and higher surface
temperatures in regions of greatest pressure. Thermal expansion is greatest at the high
temperature locations and this results in even more concentration of contact, eventually to
a few hot contact patches. This contact transition is accelerated by an increased rate of
heat generation (or increased velocity) and in materials with greater thermal expansion.
The process is slowed down or stabilized by wear (Blok 1963).

Thermal expansion tends to increase the amplitude of the surface roughness and
therefore contributes “positive feedback” to increase the localized loads on the sliding
system. However, the wear process introduces a “negative feedback” term, since it
removes material from the highest parts of the surface. Thus the stability of the system
depends on the relation between the rates of wear and thermal expansion; if the expansion
is greater than the wear, any initial irregularity in load distribution will be exaggerated
and the load may eventually be concentrated into a relatively small area (Barber 1969).
He first showed that when the rate of thermal deformation exceeds that of wear the
contact area changes can become unstable. Since that time the process has been called
“thermoelastic instability”. Barber states that the expansion is linear in the initial stages
of the heat input. Thereafter, it eventually tends to reach a constant level when the heat
lost from the solid is balanced by the heat generated (steady state). If the wear remains
linear, it must eventually exceed the expansion, bringing other areas of the surface into
contact. However, the wear is not linear, since it has been found to increase with the
average temperature at the interface of this particular sliding system.
25

Barber (1969), who worked with three metal pins on steel, states if the thermal
capacity, resistance and coefficient of expansion of the pin material are constant, the
average vertical expansion (X) at any time (t) after the commencement of the cycle will be
of the form,

X = cpLVf(t) [2.4]

where:
o = proportion of frictional heat generated (calculated)
that enters one of the materials
p = coefficient of friction (measured)
L = normal load [N]
V = sliding speed [cm/min]
/(t) = function of time period [min]

Barber (1969) states that two conditions must be satisfied if the thermal expansion of
the solid is to have a significant effect on the sliding process: the initial expansion of a
contact must exceed the wear rate and the normal expansion produced must be significant
in comparison with the compliance of the surface. If the wear is always greater than the
thermal expansion, the system will be stable and will tend to distribute the load over the
entire nominal contact area (except on the microscopic scale). For small values of t,
Barber (1969) suggests that the expansion is described by.

2aiiGLV{\ + v)t
A = ------------ -------- [2.5]
PCpA
26

where:
a = coefficient of linear thermal expansion [10"5/K]
JLI= coefficient of friction (measured)

G = proportion of frictional heat generated (calculated)


that enters one of the materials
L = normal load [N]
V = sliding speed [m/s]
v = Poisson’s ratio (known)
t = time [s]
p = material density [kg/m3]
cp = thermal capacity [J/kg]
A = contact area [m2]

The analytical and experimental work described throughout this chapter was
inconsistent in materials, boundary conditions and test apparatus. Although many of the
parameters varied from test to test, many of the tribological factors were consistent. It is
important to note the pertinent information from this chapter. The PV limit supplies
critical information to the failure of a material, but it is dependent on geometry,
coefficient of friction, load, speed, wear and temperature. Load and speed have effects on
friction which generates heat. The combination of the three determine a strong
dependence on the sliding surface temperature until a certain critical temperature is
reached, and then failure is completely related to temperature. Wear can have a long term
effect on the failure of a material, but it usually can be controlled. Catastrophic wear
happens when the critical melting temperature is reached. Transfer films can reduce heat
by reducing friction at the sliding interface, thus reducing wear.
27

HEAT TRANSFER IN TRIBOLOGY

3.1 Bulk Modeling

The concept of bulk modeling uses an approximation approach in determining the


surface temperature of two materials in contact. Analytically, reasonable estimates of
surface temperature can be obtained with bulk modeling from using the heat equation.
This depends on two factors: first a better knowledge of the nature of contact at the
interface, and second finding a solution of the heat equation involving realistic but
complicated boundary conditions (Ling and Pu 1964). Furey (1964) states that the
general procedure involved in these theoretical studies is to calculate the temperatures on
the assumptions that; first, the heat is generated at the area of true contact, considering a
single area of contact as a plane heat source, and second, this heat is conducted away into
the bulk of the rubbing members. In the first part of this chapter, the factors that
influence the bulk modeling such as temperature, flash temperatures, transfer films, load,
speed and contact mechanics will be examined. The second half of this chapter explores
numerical modeling, and it compares and contrasts the different models most commonly
used for this application.

3.1.1 Surface Temperature Predictions for Moving Heat Source in Sliding Contact

The most important factors governing surface temperature magnitudes are the rate of
heat generation which depends upon sliding velocity and the frictional force, the thermal
28

properties of the contacting materials, the true nature of the real area of contact, and the
geometry of the contacting materials. In addition, the nature of the sliding contact,
whether a continuous or moving heat source, will affect temperature. A continuous heat
source contact is obtained in the thrust washer test, whereas a moving heat source contact
is obtained in such single-pass tests as pin on flat surface. Archard (1958) states that, for
single-pass rub phenomena, at slow speeds (based on Peclet number) heat has time to
spread throughout the body of the conducting material and it acts almost as if it were a
stationary heat source. Peclet number is defined as a dimensionless independent heat
transfer parameter. At higher speeds, Jaeger (1942) suggests that the Peclet number (L or
Pe) must be applied,

[3.1]

where:
V = sliding speed of body [m/s]
a = thermal diffusivity [m2/s]
a = radius of circular area of contact A [m]

At high speeds (based on Peclet number) the heat flow is one dimensional along a linear
path, since contact time is too short for the body to conduct all the heat. This statement
constitutes that sliding velocity is the most important factor in a single-pass rub
phenomena, neglecting the other factors.

The term “macroscopic or bulk surface temperature” describes the average


temperature found at the sliding interface. That temperature can be estimated by
extrapolation using Blok’s one dimensional conduction bulk temperature theory. This
29

theory assumes that all surfaces are perfectly smooth. In reality surfaces are not smooth
and surface temperatures are actually nonuniform, with flashes of locally high
temperatures. Even though Blok’s one dimensional estimate of bulk surface temperature
is not realistic in the sense that it ignores factors such as convection and geometry, it is
useful and will be used in comparison to more complicated models (Ling and Pu 1964).

Blok’s (1963) development of a one dimensional conduction bulk temperature theory


equation neglects conduction parallel and transverse to the direction of sliding. This is
because Blok’s one dimensional theory does not take area into consideration, but
considers temperature rise taking place at a point on the contact area instead of over the
whole contact area. Blok’s theory for surface temperature considers a constant flux
source moving over a semi-infinite half-space. The bulk surface temperature rise (AT)
above the ambient surface temperature is calculated by having a heat flux q” traveling
over a half space for some time t.

f Y /2
[3.2]

where:
q” = heat flux per unit area [W/m2]
t = time [sec]
k = thermal conductivity [W/m*K]
p = density [kg/m3]
cp = specific heat [J/kg*K]

Blok assumed that the frictional heat was partitioned in a manner that made the peak
temperatures on the contacting surfaces equal. This enabled the determination of a heat
30

partitioning factor for the heat fluxes entering the two bodies. However, because the
local spatial peak temperatures that were equated occur at different locations, the
assumption violates the requirement that, at all points of intimate contact, the surface
temperature of the two bodies must be equal.

A partition factor K must be taken into account between the two different materials.
The partitioning factor (K) is calculated from the thermal diffusivities of the two
materials. Thermal diffusivity (a) is defined as,

a =— [3.
PCr

the partitioning factor K, is estimated as,

a,
K= [3.4]
O f , + o ;2

In this case, K is the percent of the heat flowing into material 1.

It may be found, by considering that a proportion “1-K” of the total flux enters an
opposing semi-infinite body through a static area of diameter D, that the maximum bulk
surface temperature rise (ATmax) at the trailing edge is approximated by,

[3.5]
31

where:
K = partitioning factor (calculated)
(X= coefficient of friction (measured)
P = interface pressure [N/m2]
B = length of contact [m]
V = relative sliding velocity [cm/min]

Equation [3.5] is obtained from equation [3.2] by substituting Q = KpPV and estimating
time as t = B/V.

Stationary contacts give symmetrical temperature profiles with the maximum


temperature at the center of the source as might be expected (Berry and Barber 1984).
This is only true when Hertzian contact, a parabolic pressure distribution over the contact
area, is assumed. Another way the maximum temperature is at the center of the source is
if the heat flux is uniform and there is a uniform convection coefficient present on the
outer boundaries of the sliding system or if the outer boundaries are insulated.

Archard and Rowntree (1988) developed a similar solution for a moving heat source
circular contact of diameter D, but it was for determining average temperature rise (AT)
over the ambient surface temperature,

D V
AT = \2\KfiP —----- [3.6]
Kkpcp

where:
D = diameter of circular contact [m]
32

and
K = l/(l+0.869(ks/k)L"1/2) [m]
k = thermal conductivity of stationary body [W/m*K]
ks = thermal conductivity of the slider [W/m*K]
L = Peclet number = Va/2ct (calculated)

Archard showed that for a circular contact that the maximum temperature, 0max, is
found at the center of the trailing edge of contact, and in terms of the mean temperature
(0m= 0 ~ ©ambient)» tS ©max = 1•640m (See Figure 3.1).

D I R E C T IO N OF V E L O C I T Y
OF H EA T S OURCE

Figure 3.1 Surface Temperature Profiles for a Circular


Sliding Source Contact (Archard 1988)

The surface temperature varies from a reference of zero at the leading edge of the heat
source to a maximum at the trailing edge if the heat flux is uniformly distributed.
33

3.1.2 Flash Temperatures

Sometimes chemical changes occur at the interface of mating materials that cannot be
explained by bulk temperature calculations. When this happens, it is assumed that the
frictional energy is initially dissipated in a very small volume of material as it flows away
from the interfacial contact. High temperatures are produced at the contacts and this
produces flash temperatures at the asperities. The philosophy behind flash temperatures
has not been experimentally proven nor has it been accurately calculated analytically,
only estimated. The work done by Blok (1963), Archard (1958) and Jaeger (1942)
yielded bulk temperatures at the interface of contact and then used them to predict the
maximum temperature at the contact interface.

Archard (1958) describes with heat flow equations the theory behind “flash
temperatures”. A simple model was used to describe a moving asperity called body B,
traveling over a stationary flat object called body C (See Figure 3.2). The contact area
between the two forms a circular contact area (A). Body B receives heat from a
stationary heat source and body C from a moving heat source.
34

Figure 3.2 Model of Contact Area (Archard 1958)

The contact shown above may not be a continuous contact area (A), but instead it may
be subdivided into a number of smaller contact areas (ÔA). The temperature calculations
can be applied to deduce the temperature distribution for the general contact region A or
for the individual contact areas ÔA. Archard (1958) states, “It has been shown
experimentally and theoretically that the aggregate resistance of a larger number of small
closely grouped constrictions, each ÔA, is very little greater than the resistance of one
large constriction A.” However, in general, the largest temperatures are those deduced
for the whole contact region rather than those deduced for the smaller individual contact
areas.

Archard estimated a mean temperature over the interfacial contact area of body B and
C as shown in equation [3.7]. He assumed that in a stationary heat source situation, heat
is supplied to a fixed area of body B and steady-state conditions will exist. He also
assumed that all the heat from the heat source is supplied to body B and not divided
between bodies B and C,
35

e„=T T - [3.7]
4ak„

where:
0m= 0 - 0buik, mean rise in temperature [K]
a = radius of circular area of contact A [m]
Qb = rate of heat supplied from area B [W]
kb = thermal conductivity of body B [W/m*K]

According to Archard’s estimate the mean temperature rise is directly proportional to the
rate of heat supplied and inversely proportional to the radius of the contact area and
thermal conductivity.

Jaeger suggest that equation [3.8] for a slow moving heat source applies for L < 0.1,
while equation [3.9] for the high speed heat source is applicable to L > 5. For large
values of the Peclet number (L), the time taken for the heat to penetrate to a depth a is
large compared with the time during which the heat source is applied.

In a fast moving heat source, the depth to which heat penetrates into body C is of
interest. The time of contact is small compared with the dimensions of the contact. This
allows the problem to be treated as linear heat flow because the heat flow does not have
sufficient time to dissipate radially and the mean temperature of the body never reaches
that of the heat source. Jaegar’s (1942) theory gave a simple explanation of the
temperature distribution over the area of a fast moving contact. He showed that the
surface temperature is proportional to t1/2 (where t = duration of heat supply); therefore,
for a square contact area, it is proportional to the square root of the distance from the
leading edge of the contact. One approximation of this theory is the assumption that the
36

maximum value of the temperature at a depth “z” occurs immediately below the value of
6max at the surface. This is not strictly true but the errors involved are comparable with,
or less than, other errors involved in this simplified approach.

When the velocity is small, the slow moving heat source body B has ample time to
dissipate enough heat to establish an approximate steady-state temperature distribution on
stationary body C. In this case the temperature can be estimated as,

y - 81

where:
6m= 6 - Gbuik, mean rise in temperature [K]
a = radius of circular area of contact A [m]
Qc = rate of heat supplied from area C [W]
kc = thermal conductivity of body C [W/m*K]

which is equation [3.6] applied to body C instead of body B.

At higher speeds however, this equation no longer holds because there is not enough
time for the temperature distribution of a stationary contact to be established in body C.
The speed at which this occurs depends on the Peclet number, as previously defined. In
this case, Archard suggests the temperature can be estimated as.

f a. v '2
e = 0 3 ia [3.9]
k u \Vaj
37

where:
6m= mean rise of temperature [K]
kc = thermal conductivity of body C [W/m*K]
Qc = rate of heat supplied from area C [W]
xc = distance across area C [m]
a = radius of circular area of contact A [m]
V = speed of sliding [m/s]

The flash temperature theory can be developed further, because the flash temperature
for each body is calculated based on the assumption that all the heat is supplied to each
body. A better approximation of the true temperature takes into account that the heat is
divided between the two bodies B and C. Archard suggests the following relation,

1 1 1
[3.10]

where:
6b = mean temperature rise in body B [K]
6c = mean temperature rise in body C [K]

The maximum flash temperature is directly proportional to the heat flux, but
temperature is not determined solely from the heat flux and the properties of the body B.
In fact, it was determined that both the width of the heat source and the velocity at which
the heat source travels contributes to the maximum temperature. Blok’s (1963) work
states that at high speeds of travel, the flash temperatures tend to concentrate themselves
in a surface layer that is very thin as compared with the width of the band shaped heat
source. As a result of this, thermal stresses are set up by the flash temperatures in the
38

surface layers. Archard and Rowntree (1988) also believed that flash temperatures are
concentrated in a space of 10"4 meters in dimension and that they occur for about 10"3
seconds or less.

The experimentally obtained flash temperatures can reach extremely high values such
that the physical properties of the material are altered. Archard and Rowntree (1988) in
experiments of sliding steel-on-steel have shown that transformations of the chemical
compositions of the metal have occurred where the bulk temperature is much less than the
transformation temperature. This suggests that the interface temperatures are much
higher than the bulk temperature and that the interface temperature reaches the phase
transformation of steel. The phase transformation is critically dependent upon the
maximum temperature rather than the mean temperature achieved during the duration of
the rubbing contact. This study shows supporting evidence that flash temperatures do
exist, but they cannot be directly measured nor predicted accurately.

Anderson and Robbins (1978) and Vroegop et al. (1980) used Archard’s method to
estimate flash temperatures. Since the calculation assumes that the load is carried by one
plastic contact over the contact area, the calculated flash temperature never reached the
experimentally obtained melting temperature.

A limiting assumption in determining flash temperatures for a plastically deformed


contact is that the equation assumes that the flow hardness (pm) is independent of
temperature, and in many cases hardness decreases almost exponentially with temperature
(Lancaster 1971),

Pm = poe'ae [3.11]
39

where:
Pm = flow hardness [MN/m2]
p0 = initial flow hardness [MN/m2]
a = material constant [K"1]
0 = temperature rise [K]

Lancaster states that exponential dependence is the major source of error in the
flash temperature calculation outlined above.

3.1.3 Transfer/Surface Film

There are two different types of interface films found in sliding contact situations.
The first type, known as a ’’transfer film”, is a solid film deposited on the typically metal
counterface that a polymer is in sliding contact with. The second type of film, known as a
“surface film”, is a lubricant intentionally placed between the two contacting surfaces.
Both films tend to reduce friction and wear, and temperatures reach a maximum in the
film between the contacting materials.

Archard (1958) states that under an elasto-hydrodynamically (lubricated) condition,


wide variations in viscosity are to be expected because the viscosity increases with the
pressure. Since the pressure distribution is approximately Hertzian and the sliding
contact is a moving heat source, the viscosity rises from a low value on the front of the
moving contact, reaches a maximum at the center, and falls towards the exit side of the
contact. Heat generation should be greatest at highest viscosity, which occurs in the
center, according to Archard’s theory.
40

For most practical speeds and film thicknesses (most of which are very thin), nearly all
the heat penetrates to the substrate (stationary contact) and its temperature is almost the
same as that obtained without a film. Thus the effect of the film will be to raise the local
surface temperature and to lower or leave unaffected the temperature of the substrate.
The substrate temperature will not be increased by the presence of the film (Archard
1958).

Archard’s (1958) original assumption was that the heat, generated by viscous losses,
was distributed uniformly throughout the parallel film. This is not strictly valid because
wide variations in viscosity will exist. The maximum temperatures are achieved in
localized regions at the surface and not in the substrate. The maximum film temperature
in the contact region may be more than five times greater than the surface flash
temperature (Archard 1958).

Archard (1958) determined theoretically through electrical analogy that at low speeds
of sliding the surface temperature is increased, but the temperature drop across the
constriction resistance (the film layer) does not affect the temperature at the surface of the
steel washer (substrate temperature). But, at high speeds the influence of the film is
determined by its thickness compared with the depth of penetration of the heat. It can be
seen that flash temperature theory, which assumes that the effect of a transfer or surface
film can be neglected, clearly requires reconsideration to include elasto-hydrodynamic
conditions.
41

3.1.4 Friction and PV Limit

Kennedy (1984) defines frictional heating as a dry sliding situation in which all the
energy loss resulting from friction is transferred to the contacting bodies as heat. This
definition is appropriate if wear can be neglected, which is not always realistic.

Yamaguchi and Kashiwagi (1982), based on their theoretical works, state that the
friction coefficient (p) increases with velocity. The increased p causes PV to decrease for
the same surface temperature (%) as shown in equation [3.12],

yb- T ti) [kg/cm2 * cm/s] [3.12]

where:
H - overall heat transfer rate [cal/cm2 * °C*s]
k - heat equivalent of work [cal/kg * cm]
A - heat radiating area [cm2]
a - sliding area [cm2]
Tb - surface temperature [°C]
Ta - environmental temperature [°C]

Ludema (1980) states that the PV limit reflects the fact that rubbing of two surfaces
together produces heat. The quantities controlling the temperature rise of sliding surfaces
are related as follows,

fiWV
A 7- [3.13]
42

where:
AT —Tsurface - Tambient [K]
JI = coefficient of friction (measured)
W = applied load [N]
V = sliding speed [m/s]
a = radius of contact [m]
J = mechanical equivalent of heat [J]
ki,2= conductivity coefficient of body 1 and 2

This equation fails to correctly incorporate the surface profile, which is represented by the
quantity “a”.

Ludema points out that the coefficient of friction varies over a wide range. Therefore,
variations in pressure and velocity do not have equivalent effects, that is, load and speed
may have different effects on the coefficient of friction. Ludema found that friction levels
remain constant over a wide speed range, until at some threshold speed “surface melting
occurs...in solids of low thermal conductivity under conditions of high contact stress.”
Beyond this threshold speed the friction decreases.

Lancaster (1971) stresses the importance of the flash temperature effect and assumes
there is no direct influence of speed and load on the coefficient of friction, only an
indirect effect through the change in temperature. Ettles’s (1987) work showed that
friction decreases with velocity for low velocities but increases with velocity at higher
velocities in what he calls the thermal control regime. He suggests that this regime
occurs as the bulk temperature reaches a softening value such that asperities cannot be
supported by the bulk. This will lead to an increase in the contact area, and eventually to
the apparent area (visual contact area). Once the softening value is reached in thermally
43

controlled friction, Ettles suggest that the constant flux source should be replaced by a
constant temperature source at the surface, moving over a semi-infinite solid half-space.

3.1.5 Load and Sneed

Archard and Rowntree’s (1988) comparison between theory and experimental work
suggests that theory forecasts that the flash temperature depends upon the parameter
W 1/2V, where W is the normal load and V is the sliding speed. A tenfold reduction in the
load might therefore be expected to cause the transition speed (speed where a material’s
chemical composition changes) to increase by a factor of three. The experiments,
however, show that the transition speed shows only a twofold increase.

Ling and Pu’s (1964) work using an analytical integration method showed that the
interface temperature is rather constant for the case of uniform heat generation at average
normal pressure, on the order of approximately 100 psi. This is not necessarily true in
general. For very high normal pressures (above 1,000 psi) with a rigid rider on the slider,
the heat generation is far from uniform. The non-uniform heat generation will have
become partially-steady in the sense that, the interface temperature is not constant and the
difference in temperature from when the heat source is over a certain area and when it is
not varies over time. Otherwise the interface temperature remains steady as long as the
heat supplied by the source per unit time remains constant and the speed remains uniform
(Blok 1963).
44

3.1.6 Contact

Ling and Pu’s (1964) work with flash temperature theory focused on modeling the
asperity contact area as a large number of very small localized areas as opposed to one
continuous area of equal size. The results of their model was strongly influenced by the
estimation of asperity sizes in contact and the variation in temperature among the
asperities. It has also been shown (Archard 1958; Ling and Pu 1964) that for a given load
the asperities in contact do not differ in number, but they do differ in location. Lancaster
(1971) provided a means of accounting for the softening of the plastic surface by taking
into account that the hardness is dependent on temperature. Ling and Pu state that the
heat flux distribution depends on the history of actual contacts and the mechanical
behavior of the materials in question. In their model, the simplifying assumption is made
that the total actual area of contact for a given load does not change with time but the
number of contacts and their locations do change with time. The distribution of heat in
space and time is assumed to be a stochastic process, and Ling and Pu used a stochastic
calculation to account for the spatial shifting of contact points between surfaces. Yet, the
best results obtained analytically for flash temperatures were approximately 100°C too
high based upon experimental observation.

3.2 Numerical Modeling

The analytical work done by Blok (1963), Archard (1958) and Jaeger (1942) are
limited to unidirectional sliding with a single contact area between semi-infinite regions.
Also, the estimation of the division of frictional heat in these works is approximate and
does not capture the spatial and time variations (Vick et al. 1991). In an effort to improve
45

on the weaknesses present in analytical theory, several researchers have used numerical
modeling.

The analytical methods, for example, the integral method used by Ling and Pu (1964),
have the advantage of producing closed form solutions and can display explicitly the role
of various system parameters. However, purely analytical methods tend to be rather
mathematically complex and difficult if not impossible to apply to complicated
geometries (Vick et al. 1994). Kennedy (1981) states that transform techniques are
applicable to bodies with finite size, but their mathematical complexity for non-simple
geometries has limited their usefulness.

Numerical methods, including the finite element, boundary element and finite
difference methods, offer the flexibility to model complex problems including surface
films, divided contact, and variable velocity. However, they are not without limitations.
Numerical methods are computationally intensive and do not give closed-form solutions
(Vick et al. 1991). Vick et al. state that both the finite element and finite difference
methods require discretization of the entire domain and suffer from numerical oscillations
at high Peclet numbers. Kennedy et al. (1983) state that most finite element programs for
surface temperature analysis have been limited to quasi-steady problems in which the
material properties were temperature independent. This technique has also been found
susceptible to numerical oscillations at high sliding velocities owing to the dominance of
convective diffusion terms. Vick et al. (1991) comment that the finite difference method
is quite flexible but requires an excessively fine grid and considerable computer time to
account for steep temperature gradients near the contact area. They state that finite
element methods appear more efficient than finite difference. The boundary element
method has the advantage that it requires discretization on the boundaries only and is
particularly useful for problems involving simple geometries.

GOLDEN,CO 80401
46

3.2.1 Comparison of the Finite Difference and Finite Element Methods

The finite difference method was used by Rashid and Seireg (1987b) and the finite
element method was used by Kennedy et al. (1983) to solve for heat partitioning and
transient cases in linear sliding systems. Both the finite element and finite difference
methods do not require artificial partitioning of the frictional heat. Using finite element
method for heat transfer analysis is convenient when node point temperatures are needed
for thermal stress analysis. Also, finite element analysis can be adapted somewhat more
easily to irregular boundaries than finite difference analysis. It is seen that both finite
difference and finite element methods offer strengths and weaknesses, but it is not
possible to directly apply the results of one method to a model of the other method, since
the nodes of the finite difference method are not coincident with those of the finite
element method.

Sonn, Kim and Hong (1995) state that finite element formulations can be adopted for
both elastic and heat problems. Advantage of using the finite element method for
thermoelastic analysis coupled with heat and elastic phenomena is that the efficiency of
the calculation is increased because the same node point variables of the same mesh
layout are used in both heat and elastic analysis. They conducted a thermoelastic analysis
of brake disks represented as the coupled heat and elastic problems including non-
isothermal and multi-point contact.

In comparison to finite element method, Rashid and Seireg (1987a) state that the finite
difference scheme can be useful only by subdividing the temperature field into a number
of small regions and performing an energy balance between each node and its adjoining
nodes. This allows for adequate heat transfer analysis, but not for any kind of thermal
47

stress analysis. Therefore, the finite element method appears to be more versatile than the
finite difference method.

3.2.2 Numerical Instability Under High Peclet Numbers

Vick et al. (1994) investigated the advantages and disadvantages of various modeling
methods, with the objective of determining the method that reduced the instability at high
Peclet numbers the most effectively. They suggest that, “The numerical instability is
simply a result of the integration rule used to handle the singular integral equation and is
not an intrinsic characteristic of the solution method.” This differs from the finite
element method, which inherently displays numerical oscillations at high Peclet numbers
and requires special manipulation at the expense of numerical accuracy to eliminate the
instability (Kennedy 1981). At high Peclet numbers, convection dominates conduction in
the direction of sliding. The oscillatory motion causes heat to be convected away from
the contact area, thus decreasing the effective thermal resistance in the moving region and
allowing the moving region to receive a higher percentage of the frictional heat. At very
low Peclet numbers, the partitioning of heat is almost even and the temperature rise is
symmetric and small (Vick et al. 1994).

3.2.3 Tribology Modeling Assumptions and Boundary Conditions

Kennedy et al. (1983) numerically modeled a metal-on-metal slider with thermal


boundary conditions and assumed the following: the top surface of the slider and the
bottom surface of the slab were both assumed to be held at an arbitrary constant
temperature of 0°C.
48

The boundary element method used by Vick et al. (1994) to model a slider over a
stationary flat object, with a film in between resulted in an energy equation that represents
a balance between energy storage, energy convected in the sliding direction, and three-
dimensional conduction. Their model allowed both the heat transfer coefficient and the
external fluid temperature T«, to vary along the surface. They state that for more general
cases of finite boundaries, heat can be added, removed, or can build up against insulated
surfaces. This has a direct effect on the distribution of frictional heat and resulting
temperatures.

In Gecim and Winer’s (1986) integral transform model, they applied a number of
assumptions and boundary conditions to their journal bearing model. First, they
neglected any axial variations. Second, conduction in the circumferential direction in the
solid was neglected in comparison with convective heat transport in the same direction
with a high rotational speed. Third, the convection coefficient was assumed to be
constant. Finally, the heat flux distribution was assumed to be uniform.

Vick et al. (1994), and Gecim and Winer (1986) applied a uniform heat flux, while
Rashid and Seireg (1987b) used a Hertzian pressure contact which resulted in a parabolic
heat flux distribution. Kennedy (1981) also used a parabolic distribution of contact
pressure and heat generation.

3.2.4 Lubricant or Transfer Films in Numerical Models

The steady state bulk temperature is determined by the energy balance between the
total heat input and the cumulative losses. If cooling of the sliding surface, which is a
localized heat source on a rotating cylinder, is the only mechanism of dissipating the
49

energy, then the bulk temperature is determined by convection alone, independent of the
film effects (Gecim and Winer 1986). Rashid and Seireg (1987b) agree that, for a semi­
infinite solid moving under a stationary heat source, the interface temperature of the
sliding solid decreases with increasing conductivity of the film layers and an increasing
thickness, due to the convection influence under such conditions. Studies with lubricant
films also showed that at the maximum obtainable film thickness, the difference in
surface temperature is mainly controlled by the lubricant film thickness and its
conductivity, not convection.

For relatively thick insulative surface layers, the maximum lubricant film temperature
for the same frictional heat input rapidly increases with the reduction in the fluid film
thickness, whereas layered contacts have a tendency to achieve steady state as the
lubricant film thickness increases. Thin coating layers (< 3 pin.) show an increase in the
maximum temperature of the lubricating film with increasing film thickness. When the
surface layer is less conductive than the lubricating film, the same friction level will
produce a very high surface film temperature as the fluid film thickness decreases (Rashid
and Seireg 1987a).

Under linear sliding conditions, a slower solid has a higher maximum temperature
compared to the faster moving solid in the presence of a film with low thermal
conductivity. This is due to the slower solid having a longer residence time under the
heat source as compared to the faster solid, therefore a higher maximum surface
temperature is expected to occur on the slower solid. A thick lubricating film has a dual
effect on the maximum surface temperature; first a thick lubricant film reduces the
friction coefficient, second it increases the difference between the magnitude of the heat
flux into the solids (Rashid and Seireg 1987a).
50

Gecim and Winer’s (1986) work with different surface films of various thermal
conductivities, including high conductivity silver on steel, at both low and high speeds,
yielded interesting results. At slow speeds (100 rad/s), for a silver film to have an effect
on the penetration depth, the thickness must be > 10 pm, otherwise the resultant
temperature is approximately the same as without a film. At the same speed, if the silver
film is =150 pm, thermal penetration is into the film alone. At higher speeds (1000 rad/s)
and film thickness of 10pm, there is less heat penetration, but the silver film has achieved
thermal penetration at a thickness of 1 pm as opposed to 10 pm at the slower speed. At
this higher speed and =100 pm thick silver film, thermal penetration should still be seen.

When dealing with silver, the thermal conductivity is relatively high, but in the case of
polytetraflouroethylene (PTFE) the thermal conductivity is a low number. Gecim and
Winer’s (1986) work also involved PTFE at both low and high speeds. At slow speeds a
surface film >lpm was needed to have an effect on the penetration depth and a film of =
20 pm was needed to achieve thermal penetration in the film alone. At higher speeds the
film must be 0.1 pm thick for an effect on the penetration depth to occur, and 2 pm thick
for thermal penetration in the film alone so that only the films’ properties are needed
when determining the surface temperature.

Vick et al. (1994) have also modeled films, using a boundary element model. They
tested both moving and stationary substrates coated with films of the same thickness and
assumed that they were flat conforming pieces that have constant area. They determined
that the temperature rise is linear with respect to the friction force (pW) and contact area
(Ac) is held constant, the heat lost over the non-contacting surfaces of the films is
negligible; therefore, the interface temperature can be more accurately calculated
compared to a non-film layer model.
51

Vick et al. showed that a constant film thickness with a low conductivity substrate
increases the contact area temperature, because conduction from the contact area is
inhibited. Similarly, a highly conductive substrate aids in the removal of energy and
lowers the interface temperature. These trends become more pronounced as the film
becomes thinner since the substrate gets closer to the contact area.

Vick et al. found that for all values of the Peclet number (Pe), a moving substrate with
a film could yield a minimum temperature rise (temperature rise of substrate relative to
ambient) at a film thickness (lj) of infinity, but a maximum temperature rise is then
attained for lj = 0. The temperature rise also increases for a given film thickness as the Pe
increases, since Pe could increase due to a larger velocity or larger contact area.

3.2.5 Transient Thermal Analysis

Sonn et al. (1995) state that two useful solution methods for transient problems are
mode superposition and direct temporal integration method. Transient finite element heat
conduction analysis of a journal bearing by Kennedy (1981), showed that within a very
short time after establishment of the contact zone the temperature distribution in both
stationary and moving bodies approached a steady-state relative to a stationary observer.
The length of time required for this quasi-steady state (steady state with respect to a
stationary observer) temperature distribution to be reached was found to be approximately
the length of time for the moving body to move a distance of 2.5 to 3 times the width of
the contact patch. Because this time is so short, it may be concluded that in the vicinity
of actual patches of contact a steady state temperature distribution occurs in the stationary
body, while a quasi-stationary distribution occurs in the moving body. Much of the effect
52

of velocity occurs at low rotational speeds, or Pe < 10. Thus, it is not necessary to do a
transient temperature analysis.

Vick et al. (1991) showed situations involving time varying velocity, when oscillating
contacts never reach steady state and a transient analysis is required. This problem is
corrected by implementing a fully implicit time integration. Steady state results can be
obtained by using a fully implicit time integration directly after one time increment by
using a sufficiently large value of the time step.

3.2.6 Division of Heat Between Contacting Bodies

Berry and Barber (1984) state that even for the lowest Peclet numbers in the high
speed range (Peclet number > 5), 74% of the frictional heat will flow into the softer of
two solids of equal conductivity, since it would accommodate the moving contacts.
However, it is unclear what would happen if the softer material has a lower thermal
conductivity.

Vick, Golan and Furey’s (1994) work found, similarly, that the nature of the materials
forming the contact area governs the frictional heating temperature rise. The distribution
of energy between two similar stationary semi-infinite bodies is 50% to each body

It was shown by Vick et al. that with similar materials for a moving and a stationary
body that the moving body will receive 75% of the (fictionally generated heat, while the
stationary piece receives 25%. A film on the moving substrate appears to have a greater
effect on interface temperature than a film on the stationary substrate, due to convection
of heat to the air. A film on the stationary surface can alter almost 25% of the frictional
53

energy, resulting in a small temperature change while the same film on the moving
substrate can alter a maximum of 75% of the frictional energy, resulting in the much
greater temperature change. All films on the moving substrate, unless highly conductive,
do not allow heat to conduct into the moving substrate. The presence of a low
conductivity moving film decreases the distribution of energy to the moving region. Thus
the properties of the moving substrate are seldom important (Vick et al. 1994).

3.2.7 Mesh/Convergence

Vick et al. (1994) have shown using a boundary element model that for the case of no
surface films, accurate results for the average surface temperature over the contact zone
can be obtained using only a single boundary element to represent the contact. Each
contact or element is a square represented by 2A x 2B, where AX = AY = 2A = 2B. In
their model, the value given to A and B was 10 units, for a total length of 20. If
temperature distributions and maximum temperature are required, the mesh needs to be
refined significantly, particularly in the sliding direction and in the regions directly below
the frictional heat source. The numerical results generated with more than one element
converged and showed no noticeable difference between 9 and 16 elements for the
division of heat and maximum temperature rise, which correspond to element sizes of 6.9
units per element for 9 elements and 5 units per element for 16 elements, respectively.

Convergence of thermal results for models of sliding contact is quite rapid, usually
requiring three or fewer iterations for a finite element model. Kennedy (1981) states that
by making the finite element mesh smaller, one can reduce the error where numerical
oscillations are occurring. In all mesh convergence test cases the maximum surface
54

temperature was predicted to be 20-30% higher than the measured thermocouple


temperature, which is located 1.25 mm below the surface.

Rashid and Seireg (1987a) state that the choice of mesh sizes for any finite difference
problem is dictated by the truncation error of the difference equation and in some cases by
the numerical stability criteria. In the direction of primary heat flow path, they found that
fewer nodes were needed in the temperature field preceding the contact zone entrance
compared to the ones following the contact zone trailing edge, due to convective heat
transfer.

3.2.8 Contact

The journal bearing model of Floquet et al. (1977) ignored changing temperatures of
the inner radius shaft, and assumed that conduction takes place radially both directions
from the contact arc, into both the shaft and the half bearing. The problem with
variations in the surface temperature due to random contacts is known as “temperature
jumps”. At the non-contact areas, the temperature drops off as the air pockets are cooled.
Temperature jumps are difficult to assess. They are less important in steel-plastic
contacts than in steel-steel contacts for two reasons. First of all, since plastics are softer
than metal, their surfaces form to that of the steel and the real and apparent contact areas
are approximately the same. Second, the air trapped between asperity contacts acts as an
insulator or thermal barrier when highly conductive steel surfaces touch but are less
effective insulators when low conductivity plastic liners are placed along the conduction
path. The experimental work of Floquet et al. (1977) with journal bearings showed that
third body effects such as wear debris, are negligible and that they could be neglected in
the numerical model.
55

The works described in this chapter will be compared and contrasted to the work done
for this project. Many of the boundary conditions and results are similar, but there are
also many dissimilarities. The test apparatus and numerical model that will be presented
are different from the ones in these previous studies. The model still accounts for steady
state sliding on a stationary flat object, but the heat flux is constant with each point of
contact, not varying with time, because the contact area is constant with the geometry
studied.
56

EXPERIMENTAL TECHNIQUE

In order to better understand temperature effects on friction and wear related to the PV
limit, experimental data was obtained for nylon 6,6 sliding against a stationary steel
counterface. The size, shape and thermal properties of both the counterface and the
plastic material affect the temperature characteristics and wear between two different
sliding materials. Therefore, the choice of test apparatus, which was that from ASTM
specification D 3702-90, and materials was an important decision.

Many different kinds of testing apparatus have been used when studying the effects of
friction, load, speed, and lubricants. Most test devices use a slider over a stationary flat;
for example, one type consists of a one or three pin hemispherical slider on a stationary
flat. Another type of device is ajournai bearing, which is a rotating shaft on a stationary
half bearing. These test methods apply in theory to Archard’s (1958) constant heat source
moving over a stationary body. The theory states that a steady state condition cannot
exist in the stationary body at high speeds, because the contact time is not long enough to
reach thermal steady-state. Much research has been performed using these techniques
and these tests have resulted in improved fundamental material understanding; however,
we preferred to use an apparatus which provided a steady-state thermal condition, which
was more readily modeled.

A common test apparatus found in industry for the study of polymer tribology consists
of a polymer thrust washer rotating on a stationary metal washer. This situation can be
considered a stationary heat source rotating over a stationary flat object. This
57

configuration differs from those described in the previous paragraph in that the stationary
heat source will allow the interface between the sliding bodies to achieve a steady-state
temperature if the test is run long enough. Since the asperities of the softer material form
to the surface of the counterface, contact is nearly constant with the heat source, not
allowing convective cooling to occur throughout the sliding surface. Therefore, this
method results in higher wear, greater temperatures and continuous changes in the contact
area.

4.1 Falex Setup

The Falex Multispecimen Test Machine available from the Falex Corporation, shown
in Figure 4.1, was the test apparatus used to conduct the experimental work described in
this chapter.

Figure 4.1 Falex Multispecimen Test Machine (ASTM 1990)


58

We chose the Falex system because one of our goals on this project was to provide
information useful to industiy, and the Falex is among the most widely used devices in
industry. The Falex Multispecimen Test Machine can be used for the ASTM
specification D 3702-90, “Standard Test Method for Wear Rate of Materials in Self-
Lubricated Rubbing Contact Using a Thrust Washer Testing Machine”. The key features
of the Falex are the test specimen, which can spin at a range of speeds, and the steel
washer which is held stationary and in contact with the test specimen under a normal load
that can be varied (See Figure 4.2).

- « OT A f t T S H N O L C ( W I T H
O ft'VC U N S 4 B A I L S E A T )

Figure 4.2 Thrust Washer Test Specimen Arrangement (ASTM 1990)

From the standpoint of heat transfer, the geometry of the rotary spindle, rotary specimen
holder and stationary specimen holder are also important. These parts are made out of a
stainless steel, which is a good heat conducting material, thus allowing heat to be pulled
away from the sliding surface.

The size, shape and surface roughness of the test specimen (thrust washer) and steel
washer used in the experiment were based on the ASTM specification. The thrust washer
59

was machined from an extruded bar of Polypenco nylon 6/6 to the following ASTM
dimensions (See Figure 4.3).

TO B E S P E C I F I E D

'.oes U iis*

Figure 4.3 Rotating Test Specimen (Thrust Washer), (ASTM 1990)


(Dimensions are in Inches, Parentheses are in mm)

The geometry of the stationary steel washer, which was made from AISI C-1018 steel,
is shown in Figure 4.4. The surface finish was 41 ± 5 pm ( 16 ± 2 pin).

.031 M A X . X « 1

Figure 4.4 Stationary Steel Washer Specimen (ASTM 1990)


(Dimensions are in Inches, Parentheses are in mm)
60

In the normal Falex operation, a thermocouple measures the stationary steel washer
subsurface temperature. This measurement allows for an estimate of the actual surface
temperature at the center of sliding contact and is believed to yield a bulk surface
temperature, defined by Ling and Pu (1964). This thermocouple is a ANSI Type K,
nickel-chromium and nickel-aluminum, and is located 0.0024 m (0.093 in) below the
sliding interface and 0.0032 m (0.125 in) into the steel washer from the outer radius. The
diameter of the Falex thermocouple is approximately 0.0016 nominal meters (0.0625 in).

Torque is measured by a load cell mounted on the stationary specimen holder. The
load cell measures force at a radius (or torque) caused by the rotating specimen in contact
with the stationary specimen, so that torque is equal to the friction force multiplied by the
radius. From this torque measurement the coefficient of friction (j l i ), can be calculated by
dividing the friction force (F), by the normal load applied (N). The load on the rotating
specimen is applied by a pneumatic cylinder located at the base of the Falex. The load is
held constant even as wear is occurring. Wear is measured from a linear variable
displacement transducer (LVDT) or by weight measurements taken before and after
testing. The rotating specimen is driven by a variable speed motor that applies a constant
speed.

The Falex has a cooling fan located in the rear of the test chamber. Tests can be run
with or without the cooling fan on. However, steady state is achieved much faster with
the fan on. The data acquisition and control system of the Falex machine allows the
specimen rotating speed, specimen load, total test time, step time and sampling time
parameters to be set. Step time is used if there is a change in the speed or load of the test
at some desired time, and sampling time is the amount of time in between each
measurement taken by the Falex system. The Falex data acquisition system returns
values of speed, load, torque, coefficient of friction, wear, temperature from the
61

thermocouple inserted in the stationary steel washer and minimum/maximum values for
these factors, averaged over a set time period, which was the sampling time at 90 seconds
for each recorded data measurement.

4.2 Material Properties

The materials chosen for this testing procedure were materials commonly used in
previous experiments such as; Vroegop et al. (1980), and from ASTM specification D
3702-90. Nylon 6,6 was chosen because its material properties are robust, it is capable of
many uses and its commonly used in industry. In addition, PV limit, wear, and
coefficient of friction data for nylon 6,6 are readily available for comparison with our
results.

In polymer-on-metal sliding systems, it is common that a transfer film (material laid


down by the polymer onto the counterface material) develops. Nylon 6,6 is not known to
have heavy wear, thus nylon 6,6 would not be expected to produce a significant transfer
film compared to other polymers like polytetraflouroethylene (PTFE). Avoiding transfer
films was a concern, because in the presence of a transfer film, the heat generation area
and conduction path are altered in a way difficult to predict. These experiments were
conducted to determine interface temperatures by use of thermocouples and one
dimensional continuous conduction theory (bulk surface temperatures). Transfer films do
not allow for an accurate measurement of bulk surface temperatures.

Another important material property of nylon 6,6 is thermal conductivity. Tests were
run at varying temperature conditions to estimate how thermal conductivity varies with
temperature (See Figure 4.5). The thermal conductivity appears to be constant between
62

50 and 225°C. The sudden increase above 225°C is due to the approaching melt
temperature (phase change) for nylon 6,6 (See Figure 4.5).

0.5 --

0.4 --

0.2 --

0 50 100 150 200 250


T e m p e r a t u r e (C )

Figure 4.5 Thermal Conductivity of Nylon 6,6 at Various Temperatures

The thermomechanical properties (loss modulus and storage modulus) of nylon 6,6 are
important when trying to understand behavior as temperature increases. All materials
contain damping, which is important in maintaining contact between the two mating
surfaces, thus allowing for consistent conduction. Damping or hysteretic losses in
polymers come as a result of oscillating elastic motion of the polymeric chains.
Polymeric materials will damp vibration much more effectively than a metallic material.
Polymers have loss moduli in the range of 1 to 200% of the storage moduli. This can be
explained by comparing two structures that have the same stiffness (storage modulus, E')
but different damping (loss modulus, E"). For the same energy input at one end of a
63

structure, less energy will be transmitted at the other end for the one with greater
damping. This combination of E' and E" is known as the complex modulus.

E* = Ez + iE" [4.1]

For the purely elastic case, Young’s modulus (E) can be related to the shear modulus (G)
by:

E
G =— [4.2]

Three important thermomechanical properties have been graphed (See Figure 4.6)
from the data obtained from shear tests conducted on the nylon 6,6. The first parameter is
the shear storage modulus (G'), which is the “in phase” modulus. The second parameter
is the shear loss modulus (G"), which is the “out of phase” modulus. The shear loss
modulus is internally generated heat. The third factor is called the shear tangent loss
factor (tan Ô). It is a function of the shear loss and shear storage modulus as shown in
equation [4.3].

tan 8 = —- [4.3]
G

The tangent loss factor also is related to friction, which is not experimentally shown in
Figure 4.6.

In Figure 4.6, it can be seen from the shear storage modulus (G’) that nylon 6,6 is rigid
up to approximately 60°C, then the glass transition starts around 75°C. The next
64

transition, into the rubbery state, is at about 100°C, and the melt temperature appears to
occur at roughly 260°C. This apparent melt is in reasonable agreement with published
melt temperatures for nylon 6,6 which are approximately 270°C, depending on how the
nylon was formed (LNP Engineering Plastics).

Nylon 6,6

1.0 0 E -0 1

I.00E -02
- G ' [d y n /cm .sq .J
- G " [d y n /cm .sq .J -- 8 .0 0 E -0 2

Rigid - -T a n D elta [d e g re e s ]
-- 7 .0 0 E -0 2
Glass
1.00E + 10 -- Transition -- 6 .0 0 E -0 2
Rubbery
5 .0 0 E -0 2

M elt A- - 4 .0 0 E -0 2

-- 3 .0 0 E -0 2

-- 2 .0 0 E -0 2

1 .0 0 E -0 2

1 .0 0 E + 0 7 0 .0 0 E + 0 0
-50 -10 30 70 110 150 190 230
T e m p e r a t u r e (C )

Figure 4.6 Thermomechanical Properties of Nylon 6,6

The stationary steel washer material was chosen based on the ASTM specification D
3702-90. This specification calls for a steel washer of AISI C-1018 steel. The thermal
conductivity for this material is 60.1 W/m*K with a Rockwell hardness of Rc 20±5. The
thermomechanical properties of steel are not considered to be of great importance in
relation to nylon 6,6. Metallic materials often have very low levels of damping, with loss
modulus on the order of 0.1 to 1.0% of the storage modulus.
65

4.3 Issues of Concern for the Falex Apparatus

The subsurface temperature and wear measurements provided by the Falex appeared to
be suspect as the heat generation increased. Therefore, we developed alternate ways of
measuring both the subsurface temperature and wear. The LVDT on the Falex appeared
to be inaccurate for measuring the wear. This device often measured negative wear after
thermal expansion had taken place. Therefore, an alternate method of weighing the test
specimens before and after each test was used in our experiments. This method yielded a
closer estimate of wear and provided data reliable up to four significant figures in grams.

The Falex thermocouple measurements are subject to errors. The surface temperature
is very important in determining the factors that surround failure, and hence having a
reliable measurement is critical. However, localized melting can not be detected at the
depth of the Falex standard thermocouple. Also, the thermocouple is relatively large, and
accurate positioning is difficult to achieve, particularly since it must be repositioned for
each test. The thermocouple becomes bent over time and does not always stay seated in
position as vibration increases. Also there are air gaps around the thermocouple that
could alter the conduction of heat. An additional concern is that with only one
thermocouple, it is impossible to draw conclusions regarding the two dimensional
temperature distribution and heat dissipation through the steel washer. Having more
temperature data in both the axial and radial directions would help elucidate the
importance of geometry and conductivity in the stationary specimen.
66

4.4 Modifications of Test Apparatus

The tribometer testing apparatus did not meet all the requirements of this study;
therefore, modifications were made to the current apparatus. The modifications consisted
of adding more thermocouples to aid in the prediction of bulk surface temperatures and to
predict conduction axially and radially through the steel washer. The steel washer was
modified to accommodate the additional thermocouples. A control and measuring system
was developed to handle the added thermocouples.

4.4.1 Thermocouple Array

In order to address the concerns that surround the surface temperature measurement of
the Falex, a new temperature measuring system was developed. In developing the new
temperature measuring system, it was desired to obtain sliding surface temperatures as
well as temperatures away from the surface. However, as described, further, surface
temperature measurements are difficult to obtain. Embedded thermocouples can be used
to measure bulk temperatures within the sliding bodies, but would not work well for
measuring contact temperatures. Measured bulk temperatures can be used effectively in
determining boundary conditions for an analytical study or for calculating the distribution
of frictional heat between the two contacting bodies (Kennedy 1984).

There are different methods for estimating temperatures at the contacting surface. One
method is the use of “dynamic thermocouples”, which are formed when two different
metals are in sliding contact. As they come into contact, they generate an electro­
magnetic force (e.m.f.) nearly proportional to a temperature difference. Questions persist
about the accuracy and meaning of the thermal e.m.f. produced by a dynamic
67

thermocouple, and also this thermocouple is limited to metals and linear one time (single
pass) testing (Kennedy 1982). These requirements do not hold for our test procedure or
the materials being used in our study. Another method is to use lasers and heat sensitive
devices to attempt to measure the surface temperature; however, this limits one of the
contacting materials to being transparent, which again is an unacceptable limitation for
our study.

The imbedded thermocouples proved to be the most practical temperature measuring


device for our experiments. We chose one of the smallest thermocouples commercially
available to reduce the amount of disturbance in the conduction path. Also, a conductive
paste was added to fill any air gaps that existed between the thermocouple and the hole.
The paste provided an excellent means of conducting heat, thus increasing the speed of
response and improving accuracy. A 40 gauge thermocouple wire was chosen. This is
approximately the size of a human hair, or 0.003 inches in diameter. The nominal cross-
section of the two wires each individually wrapped in a Teflon sheath, then wrapped
together, was 0.015 x 0.024 inches. The thermocouple was an ANSI 40 gauge Type T,
copper-constantan purchased from Omega. The temperature range for this type of
thermocouple is -270 to 150°C. This maximum temperature is far below the melt
temperature of nylon 6,6. However, this did not appear to be a problem during the testing
performed, since the thermocouple still appeared to function adequately at the highest
temperatures reached in the experimental work, which were approximately 200 °C. The
response time on the Teflon sheathed type T thermocouple is 0.23 seconds. The paste
from Omega, a filled silicone paste, withstands up to 200 °C and has a thermal
conductivity of 27.7 W/m°C. Although this conductivity is below the value of 60.1
W/m°C for the stationary steel washer, it is considerably higher than the value of 0.0262
W/m°C for air.
68

Next, to achieve the goal of measuring a more accurate subsurface temperature and to
map the two dimensional heat conduction through the stationary steel specimen, a nine-
location array of three rows of three thermocouples each was inserted (see section 4.3.2)
into the stationary specimen. The locations of the nine thermocouples are shown in
Figure 4.7.

4.4.2 Steel Washer Modification

To accommodate the addition of the nine thermocouples, the steel washer obtained
from Falex was modified (See Figure 4.7).

□P
C io c
H o le
Dio..

(p

FRONT RIGHT SIDE


FQlex Thermocouple
D ia. 0 . 0 6 7 x 0 .1 2 5 DEEP
o o o (Ref.)
o o o

\ '— 3 HOLES x D ia. 0 . 0 2 5 x 0 .0 5 2 DEEP


E - 3 HOLES X D ia. 0 . 0 2 5 x 0 .0 9 1 D E E P
3 HOLES x D ia. 0 . 0 2 5 x 0 .1 9 7 DEEP

Figure 4.7 Modified Stationary Steel Specimen (All Dimensions in Inches)


69

The added thermocouples are located 90° counterclockwise from the Falex
thermocouple (See top view of Figure 4.7). The first row is at an axial depth halfway
between the top surface of the stationary steel specimen and the Falex drilled
thermocouple hole, or 0.046 inches. This row was intended to provide a better
approximation of the surface temperature without getting too close to the surface. The
second row is at the same axial depth from the top surface as the Falex drilled
thermocouple hole, or 0.093 inches. This row was intended as a check on the accuracy of
the Falex thermocouple, and it also could detect if the first row disrupted the flow of heat
into the second row by measuring considerably lower temperatures. The third row has the
same displacement from the second row as does the first row has from the second row.
The role of the third row (depth of 0.186 in.) is to provide more information about the
axial temperature gradients. Each column of thermocouples was set at three different
depths, so that radial variations could be measured. The leftmost column from the top
view is the deepest drilled hole. This radial depth, shown as “A” in Figure 4.9, is at a
radial position of 0.197 in. from the outer radius, inside the polymer-metal contact area.
The second column, which is in the middle, is at a radial position of 0.091 in. from the
outer radius, that is in the exact radial center of the contact area and is at the same radial
position as the Falex thermocouple. The third column is the shortest of the three radial
depths. The radial depth places these thermocouples at 0.052 in. from the outer radius,
just outside of the polymer-metal contact area, as shown by location “B” in Figure 4.8.
The thermocouple columns were each rotated far enough circumferentially from each
other that the neighboring columns did not interfere with conduction in the horizontal
direction. In addition to the nine thermocouple array, three additional thermocouples
were used to take ambient temperature measurements at high, middle and low positions
along the shaft of the Falex.
70

F alex T h erm o co u p le
L ocation

Figure 4.8 Radial Placement of Thermocouples in Array

4.4.3 Control and Measuring System

With the addition of the thermocouples, another data acquisition system needed to be
put in place. A Fluke Helios Plus 2287A data acquisition system was used to record the
12 additional thermocouples signals. The Fluke Helios is a data acquisition and control
subsystem that can be used with a personal or mainframe computer. The Helios
combines a full range of measurement capabilities with easy programmability. This
system is a medium speed, highly accurate measurement and control system that has a
maximum of 1000 channels. In order for the Helios to receive thermocouple signals
certain option cards needed to be added. First, the analog to digital (A/D) converter had
to be added and the 161 High Performance A/D Converter was chosen over the 165 Fast
A/D Converter. Both A/D converters were ample, but the need for high accuracy of
analog to digital conversion of scanner input voltages was considered more important.
Since the High Performance A/D Converter card was chosen, the 162 Thermocouple/DC
Volts Scanner was the appropriate card to couple with the A/D converter for
thermocouple signals. This scanner operates as a self-calibrating analog multiplexer,
linking the High Performance A/D Converter to external measurement points. Finally,
71

the system needed a way to connect the thermocouples to the data acquisition system.
This is where the 175 Isothermal Input Connector was installed. This card supplies a
maximum of 20 thermocouple channels that each have a high, low and shield terminal,
which are routed to the scanner.

Once the system was assembled, a program to run and record the data of interest was
needed. A modification to an existing program was used. The original program, written
in Qbasic 4.5, took temperature measurements only when instructed to do so by the
operator from the menu. The program was adapted to repeatedly read signals from 12
thermocouples at intervals setup initially by the operator. The program continued taking
these measurements at the assigned intervals until the stop time was reached or until
manually stopped by the operator. The data being collected were stored to a file on the
hard drive. The program was run from a remote laptop computer and connected to the
Helios through an RS-232-C cable.

4.5 Problems Associated with Falex Modifications

Some of the problems that occurred with the modified test setup were related to the
size and fragility of the equipment being used under sometimes harsh test conditions.
The stationary steel washer with its nine thermocouple holes was costly to make because
of the machining hours involved. The drilled holes, only 0.025 inches in diameter, were
extremely small and at the lowest limit possible when machining on a drill press.

The very small Omega thermocouples used were determined, in retrospect, to be too
fragile for this type of testing. Under certain test conditions the thermocouples tended to
vibrate out of their holes. The thermocouple wires were approximately rectangular, with
72

a widest dimension of 0.024 inches, while the diameter of the drilled hole was 0.025
inches. There was enough clearance between the wires and holes that even the
conductive paste could not hold the wires in. The wires were not rigid, but hung down
under their weight, which also tended to pull them out. Also, since the thermocouples
bent easily and had only a very thin Teflon protective coat on them, they broke quite
easily, and required re-beading several times. In the process of re-beading the
thermocouples, the wire can become twisted which can produce an unwanted junction
point. This results in a temperature measurement at the twist rather than at the bead at the
thermocouple end. These problems made acquiring the temperature data tedious because
of the very careful and continuous inspection the thermocouple required.

4.6 Final Experimental Setup

The final setup (See Figure 4.9) was very similar to the setup previously described, but
the problem of the thermocouples falling out was addressed. To fix the vibration
problem, the thermocouples needed to be isolated. Therefore, two bolts were screwed
into holes drilled in the front of the stationary specimen holder and locked down with
nuts to secure them. A spring was then attached between the two bolts and locked into
place by nuts on each end. The thermocouple wires were run through a slot in the spring
and put in their corresponding holes. The spring, which helped isolate the thermocouples
from the vibration in the stationary steel washer and also supported most of the weight of
the wires, kept the thermocouples from backing out. It also helped ease the stresses on
the beads and resulted in fewer broken thermocouples.
v- Data Acquisition

, Falex Multispecimen Tester

y heeos
/ Tempereiure
Recorder

Figure 4.9 Final Experimental Setup


74

RESULTS AND DISCUSSION OF EXPERIMENTAL WORK

In Chapter 2, the experimental work by Vroegop et al. (1980), Lancaster (1971), etc.
and the different setups used in previous studies were thoroughly investigated. The
experimental setup being used in our study differs fundamentally in sliding contact from
most of those described in the literature. It was determined from the experimental work
performed for this study, that in contrast with the literature, the heat generation at the
sliding surface and partitioning of heat between each contacting material differed. This
was true even though the materials used were the same (ex. Vroegop et al. 1980), but the
geometries of the sliding pieces differed.

5.1 Approach to Experimental Testing and Acquiring Data

The originally proposed test matrix for the experimental work was intended to allow
investigation of heat transfer and tribological behavior for PV levels near the PV limit,
according to Pressure x Velocity (PV) limits of nylon 6,6. However, experiments run at
the published PV limits did not result in failure for these test conditions. Therefore, more
tests were run at elevated PV levels in a free convection environment in an attempt to
obtain catastrophic failure of the nylon 6,6 thrust washer. However, catastrophic failure
of the nylon 6,6 was never obtained, even at the highest values of PV we could achieve
with the tester. The experimental matrix for the PV tests covered a wide range of speed,
load and temperatures. The tests were conducted in both free (5-15 W/m2*K) and forced
(15-200 W/m2*K) convection environments. Temperatures in the stationary steel washer
75

were obtained with a single thermocouple supplied with the tester and the array of
additional thermocouples imbedded in the steel washer for this study. These
thermocouples were used to estimate surface temperatures and both radial and axial heat
dissipation. The test matrix was repeated to determine repeatability of the test method.
Wear was also monitored in an effort to assure that material failure was due to
temperature, speed and load, and not simply excessive wear.

5.1.1 PV Limit Test

The test apparatus used to conduct our experiments was a widely used industrial tester
made by the Falex Corporation. The Falex tester will be referred to as a tribometer tester
for the remainder of the thesis. Published data are available for the PV limit of nylon 6,6.
However, failure was not seen in the experiments at these PV levels, as discussed
previously. Published data from the LNP Corporation (LNP Engineering Plastics) state
that nylon 6,6 has the limiting PV values shown in Figure 5.1.
76

10000

1000 --

s
I

10--
LNP PV D a t a

0.01 0.1 1 10
V e lo c ity (m /s )

Figure 5.1 Limiting Pressure x Velocity Values

The method for conducting the PV limit test on the tribometer tester used a fixed
velocity throughout the test and an increasing load, incremented by 44 N (10 lb) every
two hours. The test was run with the cooling fan off, which was considered a free
convection environment. Based on the LNP data and a contact area of 1.31 cm2 (0.02029
in2), it was believed that the PV limit would be obtained between 267 - 311 N (60-70 lb)
for a speed of 0.0508 m/s (10 ft/min). From Figure 5.2 it can be seen that the values for
load were considerably higher than the ones obtained by LNP and yet failure was not
obtained (LNP Engineering Plastics).
77

1.2 200

■- 180

-- 160

- 140
0.8 --

- 0.6 - : -io o ||

0.4 --
- 60

Coefficient of Friction -- 40
0.2 -- Load
Subsurface Temperature
-- 20

0 4 8 12 16 20 24
T im e (h)

Figure 5.2 PV Limit Test Performed on Tribometer Tester; 133-623 N at 0.0508 m/s
(Load = 30-140 lb, V = 10 ft/min)

There are several potential reasons why the material did not fail as expected. First, the
two nylon materials were manufactured differently, allowing for the nylon molecules to
align differently thus affecting strength and wear. Also, PV tests conducted by others are
likely to have variations in materials, geometries of the contacting pieces and test
apparatus design, thus providing different conduction and convection away from the
contacting bodies. It is believed that the test apparatus used at LNP was of the thrust
washer on a steel washer configuration, but it is believed to be a modified drill press
rather than a Falex tester.
78

The LNP PV limit at 0.0508 m/s (10 ft/min) was reported as a PV of 105 kPa-m/s
(3,000 psi-ft/min), or a pressure of 2,067 kPa (300 psi). In the PV limit test conducted for
the current study, the load range was set to run between 133-623 N (30-140 lb), which
should have exceeded the PV limit of the LNP material by over twice the reported value,
since the resulting pressures were 1,016 to 4,755 N/m2 (148-690 psi). At the LNP
published PV limit of 267 N, the tribometer tester showed a subsurface temperature of
65°C, which is approximately 200°C less than the LNP obtained temperature. It can be
seen in Figure 5.2 that at 22 hours the temperature dropped off rapidly. This is due to the
fact that the required torque had exceeded the torque the motor could produce. The gears
for the motor were changed out so that the motor could produce over 135.6 N*m (100 ft-
1b) of torque, which would be expected to accommodate all values of torque required in
the next PV test.

A final test was conducted at a range of 556-1,779 N (125-400 lb) and 0.0508 m/s (10
ft/min) in an attempt to obtain the melt temperature. The force of 1,779 N or a PV of 700
kPa-m/s, is more than six times the PV limit of the LNP material (See Figure 5.3).
79

1.2 450

- 400

-- 350

I
0.8 -- -- 300

-- 250
"5 0.6
c
-- 2 0 0 g-
8
o
0.4 -- -- 150

-- 100
- Coefficient of Friction
0.2 —Load
- - Subsurface Temperature -- 50

0 4 8 12 16 20 24
T im e (h)

Figure 5.3 PV Limit Test; 556-1,779 N at 0.0508 m/s


(Load = 125-400 lb, V = 10 ft/min)

Under these heavy load conditions, there still was not any catastrophic failure of the
material. In addition the approximate maximum surface temperature reached was only
170°C, which is approximately 100°C lower than the estimated melt temperature of nylon
6,6.

The difference in PV limits between LNP and the test performed on our tribometer
tester is believed to be due to first, differences in the conducting materials of the spinning
holder and shaft and the stationary holder and shaft, and second, differences in the
convection. It is believed that the thermal conductivity of the non-contacting pieces for
80

the LNP test apparatus is much less than that of the stainless steel pieces used on the
tester used in this study. Geometry of the shafts and holders could also have differed.
The difference in convection suspected is due to the possibility that LNP may have
pumped heat into the surrounding air, thus increasing the interface temperature. It is also
possible that the sliding surfaces in the LNP device were better secluded from ambient
conditions, resulting in a lower free convection coefficient, thus reducing convective
cooling and raising the interface temperature.

The importance of wear during these tests must also be considered. Wear is
considered excessive for PV tests, according to ASTM D-3702, when the wear rate
exceeds 6.1 x 10'4 cm/h. Figure 5.4 shows that, for a test with the fan off (free
convection), the linear variable displacement transducer (LVDT) readings taken by the
tribometer tester are not reliable for predicting wear, since this is in part due to the test
never reaching steady state for the fan off during some of the two hour intervals,
primarily between 17 to 24 hours. If steady state is not reached then thermal expansion is
still taking place, thus the negative wear depths indicated from 17 hours until the end of
the test. The wear depth is important because as wear occurs the distance between the
contact surface and the imbedded thermocouple decreases. The smaller the distance
between the contact surface and thermocouple, the higher the recorded temperatures since
there is less material through which a temperature drop can occur.
81

0.01

0.008

0.006
Wear Depth

0.004

0.002 -

20
- 0.002 - -

-0.004

-0.006

-0.008

- 0.01

T im e (h)

Figure 5.4 Wear Depth of PV Limit Test; 556-1,779 N at 0.0508 m/s


(Load = 125-400 lb, V = 10 ft/min)

To get a more reliable, although integrated measure of the wear, the weight for the
nylon thrust washer was taken before and after the PV test. The loss of weight for the
thrust washer over the 24 hour period was 0.0337 grams, resulting in a volumetric loss of
0.0043 cm3:

Volume = = - ^ Z l L = 0.0043 ^ [5.1]


P 7.872— f
cm
82

This results in an average wear rate over the 24 hours of 1.79 x 10‘4cm3/h (or 5.38 x 10"5
in/h), which is considered excessive in the ASTM (1990) specification.

Another method of estimating wear is to measure the thickness of the thrust washer
before and after the test. This method yielded an estimate for the wear rate of 9.31xl0*4
cm/h:

At (cm) 0.022352 cm . cm
w=i ^ r ^ j ^ = 93M 0 t [5-2]

(or 3.67 x 10"4 in/h)

These two methods for measuring the average wear rate are not the most appropriate for
PV tests, due to the varying loads. The wear rate can also be calculated for the nylon 6,6
thrust washer using equation [2.3], but this would require individual calculations at each
load and time step (2 hours). Then the results at each load and time interval would be
totaled to find the final wear volume, which can be converted into the wear rate:

wear (cm3) 1.05 x 10"1cm3 , cm


W = -------t— ^ — ----- — —- = 3.34x 10-3— [5.3
Area (cm ) •time (h)1.31 cm 24 /r h

(or 1.31 x 10"3 in/h)

These values for wear are much higher than the ASTM D-3702 recommended maximum
value of 6.1 x 10 4 cm/h (1.0 x 10'5 in/h). The wear seen here is much higher than that
seen in the test matrix, due to higher loads subjected during the PV test.
83

5.1.2 Test Matrix

A target for the test cases to run was based on the LNP data. It was determined that
due to the inability to reach material failure through melting, the test matrix should at
least be comprised of a wide range of temperatures. All tests were run for both fan off
and fan on conditions to study the effects of free and forced convection on the materials.

The first data in the testing matrix were taken at three different values of P and V
below the LNP PV limit line, and one value above the LNP PV limit line as shown in
Figure 5.5. The same PV was attempted for all tests. At the speeds of 0.0508, 0.508 and
1.27 m/s this was nearly accomplished, while at 0.254 m/s the PV was higher, due to a
load that was above the LNP PV limit line. Originally, the goal was to take data at
0.0508, 0.508 and 5.08 m/s. Unfortunately, the velocity of 5.08 m/s was considered
extreme due to severe vibration for the test device. Therefore, the maximum speed of
1.27 m/s was used. It was desired to test at a range of loads that would generate a wide
range of temperatures, and to do this, half of the LNP PV limit load was considered a
good starting point. At the maximum speed (1.27 m/s), it was not possible to obtain the
desired load due to the resolution of the pneumatic cylinder on our tester. The fourth test
point, above the LNP PV limit line, was the result of a speed calibration error when trying
to run the first test (133 N, 0.0508 m/s). The velocity was actually three times higher than
that of the first test (See Figure 5.5).
84

10000

1000 --

S O
. 100 --

10--

LNP PV D a t a
■ E x p e r im e n ta l

0.01 0.1 1 10
V elocity (m /s)

Figure 5.5 Original Test Matrix

The testing conducted at the four combinations of P and V shown in Figure 5.5 was
preliminary and used the tribometer tester thermocouple as the only means of measuring
the approximate surface temperature. The effects of free and forced convection also were
investigated for each of these tests, since this information was expected to have a
significant influence on determining our final test matrix.

The first test (See Figure 5.6 and 5.7) was run at the slowest speed of 0.0508 m/s, with
a fixed load of 133 N, for forced convection (fan on) and free convection (fan off):
85

200

-180

-- 160

-- 140
0 .8 - -

Co e ffi c ie n t of Friction -- 120

Temperature (C)
LL Subsurface Tem perature
0 .6 -- -- 100
%
A"üT»iHfTi*r"n yn-yrw >r> y nnv n »h»t so
0. 4 --
-- 60

-- 40
0.2 -

-- 20

0 2 4 6 8 10 12
T im e (h )

Figure 5.6 Free Convection for Friction Coefficient and Temperature at Half LNP PV
Limit; 133 N and 0.0508 m/s (Load = 30 lb, V = 10 ft/min)

1 .2 200

— — C o e ff ic ie n t of F ric tio n
0 .8
S u b s u rfa c e T e m p e ra tu re

Temperature (C)
0.6

0.4

0 .2

0
0 2 4 6 8 10 12
T im e (h)

Figure 5.7 Forced Convection for Friction Coefficient and Temperature at Half LNP PV
Limit; 133 N and 0.0508 m/s (Load = 30 lb, V = 10 ft/min)
86

By comparing Figures 5.6 and 5.7, it can be seen that the steady state temperature was
highest in the free convection case, thus resulting in a higher coefficient of friction. For
the forced convection case, the time to reach steady state was approximately two hours,
while the free convection case was at steady state after approximately four hours. The
transient period is due to both thermal and initial wear effects. It appears, at the higher
temperature for free convection, that the coefficient of friction is more uniform than for
the forced convection case. This increased stability in the coefficient of friction could be
attributed to the deposit of a transfer film or a faster wear rate.

The next test point considered was that with the same PV as the previous test but at a
speed of 0.508 m/s for both free and forced convection (See Figures 5.8 and 5.9). It can
be seen from the graphs that for both the free and forced convection cases, the coefficient
of friction was non-uniform. This is attributed to friction-induced vibration and the low
force (13 N) being applied, which appeared to be bouncing instead of maintaining a
constant load. Under these test conditions, the free convection case reached steady state
at around two hours. This is once again shorter than for the forced convection case,
which reached steady state at approximately three hours. The time constant differs from
the previous case at 0.0508 m/s, and could be related to the heavy oscillations of the
coefficient of friction. A higher temperature and coefficient of friction were obtained in
the free convection case, than the previous case for 0.0508 m/s.
87

0 .8
Coefficient ot Friction

120 _

100 Ë

■Coefficient of Friction
S u b s u rfa c e T e m p e ra tu re

2 4 6 8 to 12
T im e (h)

Figure 5.8 Free Convection for Friction Coefficient and Temperature at Estimated
Half PV Limit; 13 N and 0.508 m/s (Load = 3 lb, V = 100 ft/min)

200

-- 180

-- 160
Coe ffi cien t of Friction
Subsurface T em perature -- 140
0 .8 --

120
I£ --

1
I -- 80

- 60

-- 40
0.2

-- 20

0 2 4 6 8 10 12

Figure 5.9 Forced Convection for Friction Coefficient and Temperature at Estimated
Half PV Limit; 13 N and 0.508 m/s (Load = 3 lb, V = 100 ft/min)
88

The third speed in the original matrix (5.08 m/s) was determined to be too fast for the
bearings on our tribometer tester, resulting in high vibration and noise problems.
Therefore, the speed chosen for the high speed was 1.27 m/s. This speed was considered
to be adequate for noise and vibration purposes, yet still more than a factor of two above
the next highest speed. In order to obtain a PV consistent with the first two tests a load
below one pound was needed; however, this was not obtainable because less than one
pound was below the resolution of the tester. Therefore, a velocity of 1.27 m/s and a load
of one pound load were chosen in order to stay near the half PV used in the first two cases
(See Figures 5.10 and 5.11).

1.2 200
-- 180

-- 160

^ “ Coefficient of Friction -- 140


0.8 -- Subsurface Temperature
I -- 120
E
0
0.6 -- -- 100 2

•- 80
o1
0.4 --
-■80

-- 40
0.2
-- 20

0 2 4 6 8 10 12
T im e (h)

Figure 5.10 Free Convection for Friction Coefficient and Temperature at Estimated
Half PV Limit; 4 N and 1.27 m/s (Load = 1 lb, V = 250 ft/min)
89

1.2 200

- 180

-- 160
^ — Coefficient of Friction
Subsurface Temperature -- 140
0.8 --
I -- 120 ~
I
° 0.6-- -- 1 0 0 s

I
o8
- - 80
0.4 -•
-- 60

-- 40
0.2 -
-- 20

0 2 4 6 8 10 12
T im e (h)

Figure 5.11 Forced Convection for Friction Coefficient and Temperature at Estimated
Half PV Limit; 4 N and 1.27 m/s (Load = 1 lb, V = 250 ft/min)

The free convection case again obtained a higher temperature; therefore, a higher
coefficient of friction was a resultant of the higher temperatures. Both test cases reached
steady state within the first hour. The heavier than normal oscillations in coefficient of
friction were due to the light load being applied, and the resolution of the data acquisition
system.

It can be seen from the previous three tests that forced convection on the tester has a
noticeable decreasing effect on the temperature. It can also be seen that the loads applied
were too light, thus oscillations in the measured values of the coefficient of friction

ARTHUR LAKES LIBRARY


COLORADO SCHOOL OF MINES
GOLDEN, CO 80401
90

resulted. Oscillations in friction coefficient were also a result of the resolution of the data
acquisition system. With the combination of pressures and velocities chosen, it can be
seen for the forced convection case that the temperatures were > 10°C over the ambient
condition 26°C. These temperatures obtained with forced convection were too low to
allow thorough study of the conduction and convection effects and they prohibited the
possibility of seeing material effects at higher temperatures.

The following two tests, run at 0.152 m/s and 133 N, were a result of a calibration
error for speed on the first test at 0.0508 m/s. Although the PV was not the same as the
previously tested speeds, it does give a good indication of moderate speed (0.152 m/s)
and higher force (133 N). This yielded a better view of what combination of pressure and
velocity would be appropriate for the test matrix (See Figures 5.12 and 5.13).

200
-- 180

- 160

-- 140
0.8 --
.2 -- 120
0.6 100
I
-- 80
I
0.4 --
-- 60

Coefficient of Friction -- 40
Subsurface Temperature
-- 20

0 2 4 6 8 10 12
T im e (h)

Figure 5.12 Free Convection for Friction Coefficient and Temperature, Preliminary
Testing; 133 N and 0.152 m/s (Load = 30 lb, V = 30 ft/min)
91

1.2 200

-- 180
1 --
-- 160

-- 140
0.8

+ 120ü
- 0.6 -- -- 1 0 0 s

I
-- 80
0.4 --
60

-- 40
0.2
■Coefficient of Friction
- Subsurface Temperature -- 20

6 10 12
T im e (h)

Figure 5.13 Forced Convection for Friction Coefficient and Temperature, Preliminary
Testing; 133 N and 0.152 m/s (Load = 30 lb, V = 30 ft/min)

Even though this PV test condition was run by error, the results obtained were more
uniform and yielded more suitable temperatures for what was trying to be achieved for
the test matrix. The coefficient of friction and temperature were higher for the free
convection case once again. At these higher loads (> 44 N) the coefficient of friction was
more uniform for both cases than for the lower loads of the previous tests. Also at the
higher loads, steady state is reached within the first hour in the forced convection case
just as was seen in the test at 0.0508 m/s (10 ft/min). The free convection case took over
four hours to reach steady state, due to the dependence of the time constant on the
convection coefficient.
92

The forced convection case, under stable test conditions, generally reached steady state
before the free convection case. This comparison is shown clearly in Figure 5.14.

120

100 --

80 -

60 -
I

40 - i

20-- forced convection


free convection

0 1 2 3 4 5 6 7 8
T im e (h)

Figure 5.14 Comparison of Free and Forced Convection at Steady State Temperatures

By reviewing the initial tests results, a new test matrix was created. The new matrix
was tested under a forced convection condition only. This was done to in an effort to
ensure that most of the data was obtained at thermal steady state before wear was an
attributing factor to the results. The new matrix had a wide range of loads over the four
speeds, and the matrix represented only two different PV values. By maintaining two PV
values at varying load and four speeds, the results would show the effects of speed and
coefficient of friction on a material. Also for these test cases, more detailed
measurements were taken in an attempt to isolate the effects of convection into the air
93

from conduction of heat into the steel washer, and to better estimate interface
temperatures. These tasks were accomplished through the implementation of a nine
thermocouple array in the steel washer (see subsection 4.4.1 for an explanation of the
setup).

The test matrix in Table 5.1 and Figure 5.15 consisted of a wide range of values of
load and speed at two PV levels. The temperature results shown on the graphs in sections
5.1.1 and 5.1.2 were recorded by the tribometer tester thermocouple.

Table 5.1 Test Matrix of Load, Speed and PV

44 88
NORM
ALLOAD(N)
111 222 444 1112
0.0508 A B
pv(175) pv(437)
0 (kPa-m/s) (kPa-m/s)
0.254 C D
pv(175) pv(437)
VELOCITY
(m/s) 0.508 E
(kPa-m/s)
F
(kPa-m/s)

pv(175) pv(437)
(kPa-m/s) (kPa-m/s)
L2Z G
pv(437)
(kPa-m/s)
94

100000

10000 --

1000 --
Pressure (kPa)

100 --

10--
PV (175 kPa-m/s)
PV (437 kPa-m/s)

0.01 0.1 1 10
V elocity (m /s)

Figure 5.15 Lines of Constant PV for Test Matrix

The PV (175 kPa-m/s) test shown as “A” in Table 5.1, was at 0.0508 m/s (10 ft/min)
and a force of 444 N (100 lb) (See Figure 5.16 and 5.17). The test was performed twice
to show repeatability.
95

1.2

-- 140
0.8 --
- - 120 ~

“ 0.6 -f f 100 ™

r-----------
I
80

40
0.2 -/
■Coefficient of Friction
-Subsurface Temperature + 20

6
T im e (h)

Figure 5.16 Test Matrix: Forced Convection at 444 N and 0.0508 m/s
(Load = 100 lb, V = 10 ft/min)

The time range used to obtain the averaged temperature and coefficient of friction for
each test is marked on the figures with a rectangle. It can be see on Figure 5.16 that the
averages were taken between 10-12 hours. Most averages were taken at this time because
of the uniform generated results. Occasionally, one of the measuring devices (usually the
thermocouple) would stop yielding valid data; therefore, the averages would be taken at
an earlier steady state time.
96

1.2 200

- 180

-- 160

-- 140
0.8 --

-- 120

Tem perature (C)


° 0.6 -- -- 100

- 80
0.4
- 60

- - 40
0.2
^ — Coefficient of Friction -- 20
Subsurface Temperature

0 2 4 6 8 10 12
T im e (h)

Figure 5.17 Test Matrix: Forced Convection Repeat at 444 N and 0.0508 m/s
(Load = 100 lb, V = 10 ft/min)

As stated earlier, these tests, represented in Figures 5.16 and 5.17, are repetitions at the
same conditions. These two tests showed agreement in coefficient of friction and
temperature within 6.5%. Both tests reached steady state within one hour and remained
uniform throughout the test. The averaged steady state temperatures of the first and
second test were 47°C and 41°C, and the coefficient of friction values were 0.46 and
0.45, respectively. It is interesting to note an increase of 6°C in temperature, while the
coefficient of friction only increased by 0.01.
97

The PV (437 kPa-m/s) test shown as “B” in Table 5.1 was performed at a force of
1,112 N (250 lb) and a velocity of 0.0508 m/s (10 ft/min) (See Figure 5.18). This test
was run only once based on the uniform results for the PV (175 kPa-m/s) case “A” at a
speed of 0.0508 m/s (10 ft/min). It was assumed another test at these conditions would
show repeatability.

200

-- 180

-- 160

-- 140
0.8 --

I -- 120
|
o
0.6 -- -- 100 «
g
y
Io v>wv». _ go

-- 60

-- 40
0.2 -■
— Coefficient of Friction
Subsurface Temperature -- 20

0 2 4 6 8 10 12
T im e (h)

Figure 5.18 Test Matrix: Forced Convection at 1,112 N and 0.0508 m/s
(Load = 250 lb, V = 10 ft/min)

The test run at 1,112 N (250 lb), case “C”, was even more uniform than the tests at 444 N
(100 lb), and both reached steady state around the same time. At 1,112 N (250 lb) the
coefficient of friction was 0.48 and the temperature was 78°C at the subsurface
thermocouple location.
98

The results of tests performed at a velocity of 0.254 m/s (50 ft/min), cases “D” and
“E”, can be seen in Figures 5.19 and 5.20. These tests were useful in combining a more
equal balance of both speed and load, while still matching the two PV levels from the
matrix.

1.2 200

180
1 --
-- 160

-- 140
0.8
Coefficient of Friction

-- 120

Temperature (C)
0.6 -- - 100

0.4 -
jlu 60

-- 40
0.2
•Coefficient of Friction
-- 20
- Subsurface Temperature

6 8 10 12
Time (h)

Figure 5.19 Test Matrix: Forced Convection at 88 N and 0.254 m/s


(Load = 20 lb, V = 50 ft/min)
99

200

-- 180

-- 160

-- 140
0.8 -■

g 120 —
£
0.6 -- 100 g

1
8
o
-- 80
0.4
-- 60

-- 40
0.2 --

^ — Coefficient of Friction
-- 20
Subsurface Temperature

0 2 4 6 8 10 12
T im e (h)

Figure 5.20 Test Matrix: Forced Convection at 222 N and 0.254 m/s
(Load = 50 lb, V = 50 ft/min)

At both 88 N (20 lb) and 222 N (50 lb) for the tests conducted at 0.254 m/s (50 ft/min), it
can be seen that the coefficient of friction results are not as uniform as those conducted at
0.0508 m/s (10 ft/min). Both tests at 0.254 m/s (50 ft/min) reach steady state at
approximately one hour. As seen in the previous case at 0.0508 m/s (10 ft/min) the
higher loads yield more uniform data and a higher temperature and coefficient of friction.
The temperature difference between the 88 N (20 lb) and 222 N (50 lb) was 58°C, where
44 N (20 lb) was 51°C and 222 N (50 lb) tests was 108°C. The coefficients of friction
were 0.50 for 88 N (20 lb) and 0.69 for 222 N (50 lb).
100

The first comparison of PV at 175 (kPa-m/s) (case “A”: P = 3,389 N/m2, V = 0.0508
m/s and case “C”: P = 672 N/m2, V = 0.254 m/s) showed a small increase from 0.46 to
0.50 in coefficient of friction. The increase in temperature was smaller than what had
been seen relative to changes in temperature due to an increase in the coefficient of
friction. The temperature increased from 47°C at 0.0508 m/s (10 ft/min) to 51°C at 0.254
m/s (50 ft/min) for the same PV. The PV at 437 (kPa-m/s) also showed the same
characteristics as the PV at 175 (kPa-m/s). The coefficient of friction increased from 0.48
at 0.0508 m/s (10 ft/min) to 0.69 at 0.254m/s (50 ft/min) and the temperature followed at
78°C and 108°C, respectively. It would appear at this time that an increase in speed has a
increasing effect on both the temperature and the coefficient of friction.

The next round of tests were conducted at 0.508 m/s (100 ft/min) for two different
forces; at 44 N (10 lb) and 111 N (25 lb) (cases “F” and “G”). These tests showed some
non-uniformity in coefficient of friction from vibrations the first time they were run, so
both were repeated to ensure accuracy. Results for the two repetitions of the test run at a
PV of 175 (kPa-m/s) (44 N) can be seen in Figures 5.21 and 5.22.
101

200

-- 180

-- 160

-- 140
0 .8 --

-- 120

| 0.6 -- -- 1 0 0 Ï

I 0.4
“ C o e ffic ie n t of Friction
— S u b s u r fa c e T e m p e r a tu r e
-

-- 60
80

-- 40
0.2

-- 20

0 2 4 6 8 10 12
T im e (h )

Figure 5.21 Test Matrix: Forced Convection at 44 N and 0.508 m/s


(Load = 10 lb, V = 100 ft/min)

■Coefficient of Friction
S u b s u r fa c e T e m p e r a tu r e

Figure 5.22 Test Matrix: Forced Convection Repeated at 44 N and 0.508 m/s
(Load = 10 lb, V = 100 ft/min)
103

200

-- 180

-- 160

-- 140
Coefficient of Friction

-- 120

Temperature (C)
0 .6 " I -- 100

-- 80

0 .4
-- 60

-- 40
0.2 -■ 1 'C o e ffic ie n t of Friction
S u b s u r fa c e T e m p e r a tu r e
-- 20

0 2 4 6 8 10 12
T im e (h )

Figure 5.23 Test Matrix: Forced Convection at 111 N and 0.508 m/s
(Load = 25 lb, V = 100 ft/min)

1.2 200

-- 180

1
-- 160

- 140
0.8

120
Temperature (C)
0.6 100
80

0.4
60

-- 40
0.2
— — C o e ffic ie n t of Friction
S u b s u r fa c e T e m p e r a tu r e -- 20

0
0 2 4 6 8 10 12

Figure 5.24 Test Matrix: Forced Convection Repeated at 111 N and 0.508 m/s
(Load = 25 lb, V = 100 ft/min)
104

Once again at the higher force,! U N (25 lb), the tests experienced more uniformity in
friction than at 44 N (10 lb), although the coefficient of friction is still not very uniform
for both tests run at 111 N (25 lb). Both tests reached steady state at around two hours.
The coefficients of friction showed poor agreement, at 0.69 and 0.75 for the first and
repeated test case. The friction value for the repeated case is much higher in comparison
to other tests of equal PV. This could be attributed to the sporadic nature of the results
and the choice of average time span in which the data were averaged. The temperatures
for the two tests were 95°C and 107°C, respectively, which again showed poor
agreement, since they differed by approximately 10%.

It can be seen for all the tests that the PV data at 437 (kPa-m/s) up to this point showed
a minimum and maximum coefficient of friction to be 0.48 and 0.75. The minimum and
maximum temperatures were 78°C and 108°C, respectively. As it was in the case of 175
(kPa-m/s) PV, the PV at 437 (kPa-m/s) shows that the highest temperature does not
correspond to the highest coefficient of friction. This is attributed to erroneous
temperature measurements recorded by the tribometer thermocouple.

The final tests conditions were run at a PV of 437 (kPa-m/s) (case “G”: P = 336 N/m2)
with a speed of 1.27 m/s (250 ft/min). It was decided that the force needed to run a test
at a PV of 175 (kPa-m/s) was too light since the results would most likely display
unacceptable non-uniformity. Therefore, a force of 44 N (101b) was used in this test case
(See Figures 5.25 and 5.26) for a resulting PV of 437 (kPa-m/s).
105

200

180

120

T e m p e ratu re (C )
100

60

40
■C o efficien t of Friction
- S u b s u r f a c e T e m p e r a tu r e

6
T im e (h )

Figure 5.25 Test Matrix: Forced Convection at 44 N and 1.27 m/s


(Load = 10 lb, V = 250 ft/min)

200

-- 180

160

140
0 .8 --

120
0.6 T e m p e ratu re (C)
-- 100

-- 80

0.4
-- 60

-- 40
C o e ffic ie n t of Friction
S u b s u r fa c e T e m p e r a tu r e -- 20

0 2 4 6 8 10 12
T im e (h )

Figure 5.26 Test Matrix: Forced Convection Repeated at 44 N and 1.27 m/s
(Load = 10 lb, V = 250 ft/min)
106

The repeated test case was the only one of the two tests run at this speed of 1.27 m/s (250
ft/min) that yielded accurate results. The data obtained in the first of the two tests were
invalid due to severe vibration. Other loads were attempted at this speed to try to reduce
vibrations in the system. Lighter loads resulted in non-uniform friction data and heavier
loads resulted in severe vibrations of the test equipment. The test at 44 N (10 lb) was
attempted several times before useful data could be obtained, due to the fact that
thermocouples would break or vibrate out (this was fixed by using a spring holder system,
see subsection 4.4.2). Also, at times the tribometer tester would shut down from
excessive vibrations. The vibration during the running of these two tests was the highest
seen during all the test conducted. The oscillations seen in the coefficients of friction for
the tests at 1.27 m/s (250 ft/min) were smaller in amplitude than the oscillations seen in
the tests run at 0.508 m/s (100 ft/min). As seen in Figure 5.25, the slope of the first test
data is exponentially decreasing until approximately nine hours and then it appears to
reach steady state. The repeated test, shown in Figure 5.26, was much more stable,
reaching steady state in approximately two hours. The coefficient of friction for the first
test was 0.78 and the repeated test showed a value of 0.89. The noticeable decline in
coefficient of friction until approximately nine hours in the first test, resulted in a much
lower value than that recorded in the repeated test. The temperatures for each test were
149°C and 139°C, respectively. This did not coincide with the higher coefficient of
friction resulting in a higher temperature. The first test was discarded because of all the
irregularities that are involved with it.

This test case was taken at PV of 437 (kPa-m/s), and when considering all tests at PV
of 437 (kPa-m/s), the minimum and maximum coefficient of friction were 0.48 and 0.89,
respectively. The minimum and maximum temperatures were 78°C and 139°C. These
results were obtained at the 0.0508 m/s (10 ft/min) and 0.254 m/s (50 ft/min) cases where
107

the maximum temperature coincides with the maximum coefficient of friction, and it
would appear that the temperature and the coefficient of friction are speed dependent.

The results of the full test matrix showing tribometer thermocouple temperatures,
coefficient of friction and PV values are summarized in Table 5.2.

Table 5.2 Test Matrix: Temperature (°C), Coefficient of Friction and PV (kPa-m/s)

44 88
NORM
ALLOAD(N)
111 222 4M 1112
0.0508 47.1°C / 0.46 77.5°C / 0.48
41.3°C / 0.45
pv(175) pv(437)
(kPa-m/s) (kPa-m/s)

0.254 50.9°C / 0.5 108.4°C / 0.69


pv(175) pv(437)
VELOCITY
(m/s)
(kPa-m/s) (kPa-m/s)

0.508 48.0°C / 0.63 95.2°C / 0.69


49.1°C/0.65 106.5°C / 0.75
pv(175) pv(437)
(kPa-m/s) (kPa-m/s)

L2Z 138.7°C / 0.89


pv(437)
(kPa-m/s)

The maximum and minimum temperatures for the PV of 175 and 437 (kPa-m/s) with
forced convection are plotted in Figures 5.27 and 5.28.
108

60

5 0 --

40 - ~fi

I
20 --

Min
10 -- M ax

0 2 4 6 8 10 12
T im e (h )

Figure 5.27 High and Low Temperatures, Forced Convection for PV at 175 (kPa-m/s)

160

140 --
.'x^rv-xV!_ ^ ' '"N
120 --

100 --

I
80
IE
60

40 -
Min
M ax
20 -

0 2 4 6 8 10 12
T im e (h )

Figure 5.28 High and Low Temperatures, Forced Convection for PV at 437 (kPa-m/s)
109

The difference between maximum and minimum temperature for PV of 175 and 437
(kPa-m/s) varied significantly. It can be seen that the difference in temperature for the
PV at 175 (kPa-m/s) is 4°C, while the difference in temperature for the PV at 437 (kPa-
m/s) is 61°C. One explanation for the temperature difference between the two PV levels
could be that the wear taking place at the higher PV is excessive.

The wear taking place over the course of the testing will alter the temperature
measurement. Once the wear has exceeded the allowable limit, then the distance between
the contact surface and the thermocouple has changed significantly enough to yield higher
temperatures from the change in distance. An approach to resolve this problem is to
compare the LVDT data from our tester to the change in weight measurement of the
nylon thrust washer or by using the change in thickness of the thrust washer. The LVDT
wear depth for both the maximum and minimum temperature cases at both PV levels can
be seen in Figures 5.29 and 5.30. These counteract the expectation that as wear
progresses, temperature increases for constant coefficient of friction, pressure, and
velocity. However, as described in the following paragraphs, the LVDT measurement is
not a reliable measure of wear.
110

Wear Depth (In)

T im e (h)

Figure 5.29 Measured LVDT Wear at the Min. and Max. Temperature
for PV(175 kPa-m/s)

0.02

0
Wear Depth (In)

- 0.02

-0.04

-0.06

-0.08

- 0.1

Tim e (h )

Figure 5.30 Measured LVDT Wear at the Min. and Max. Temperature
for PV(437 kPa-m/s)
Ill

It is quite noticeable in the maximum temperature cases for both PV levels that thermal
expansion dominated the LVDT measurement, which is why the wear depth is negative.
A better estimate of the wear rate can be calculated using either the weight loss
measurement, the thickness measurement or the distance traveled with a wear factor.

A calculation for the volume of the maximum wear rate at a PV of 437 (kPa-m/s) from
equation [2.3], for the time traveled yields:

^ in5 ■min ^
^=200x10 -10
x50 x250 J L x 20 Zzr = 0.005 in'
f t —lb —hr \in 2 j Vmin J
[2.3]
or 0.082 cm:

This yields the wear volume which can be converted to a wear rate of 3.13 xlO"3 cm/h
( 1.23x10 ^ in/h). The wear rate can be found using the change in thickness:

w= = a m 6 5 M = 4 8 x i 0 . 4^ or 1 9 x l0 . 4 m
time (hr) 20 (hr) h h

The final method of solving for the wear rate at maximum temperature for the PV of 437
(kPa-m/s) is using the weight loss of the nylon 6,6 thrust washer.

y = Aweight {g) = 0007 { g) , = 8.89 x 1 0 - W


8 8
7.872
\cm~ cm
112

This number is the wear volume, and it can be converted into a wear rate of 3.39xl0"5
cm/h ( 1.34x10"5in/h).

All three methods appear to yield different wear rates, with only the weight loss
method falling within the acceptable wear limits set in ASTM D-3702. The calculations
of wear rate at a PV of 437 (kPa-m/s) do not give enough insight to determine what
method is most accurate, although ASTM D-3702 specifies thickness measurements. A
list of the wear rates for the other tests in the matrix can be seen in Table 5.3.
113

Table 5.3 Different Methods of Calculating the Wear Rate for the Test Matrix
(A = 175 kPa-m/s and B = 437 kPa-m/s)

444 N. 0.0508 m/s: PV(A) 444 N. 0.0508 m/s: PV(A)


Test 1: Test 2:
wear rate: 1.25x1 O'3 cm/h wear rate: 1.25x1 O'3 cm/h
thickness: 1.03x1 O'3 cm/h thickness: 1.12x103 cm/h
weight: 1.32x1 O'4 cm/h weight: 1.13x1 O'4 cm/h

1112 N. 0.0508 m/s: PV(B) 88 N. 0.254 m/s: PV(A)


Test 1: Test 1:
wear rate: 3.12x10'3 cm/h wear rate: 1.25x1 O'3 cm/h
thickness: 4.88x1 O’4 cm/h thickness: 1.14x10 3 cm/h
weight: 4.17x10 3 cm/h weight: 1.21x1 O'4 cm/h

222 N. 0.254 m/s: PVfB) 44 N. 0.508 m/s: PV(A)


Test 1: Test 1:
wear rate: 3.12x1 O'3 cm/h wear rate: 1.25x1 O'3 cm/h
thickness: 3.38x10 4 cm/h thickness: 3.84x10 4 cm/h
weight: 4.22x1 O'5 cm/h weight: 3.63x1 O'5 cm/h

44 N. 0.508 m/s: PV(A) 111 N. 0.508 m/s: PVfB)


Test 2: Test 1:
wear rate: 1.25x10 3 cm/h wear rate: 3.12x1 O'3 cm/h
thickness: 7.82x10 4 cm/h thickness: 3.89x10 4 cm/h
weight: 5.99x1 O'5 cm/h weight: 6.07x1 O'5 cm/h

111 N. 0.508 m/s: PV(B) 44 N. 1.27 m/s: PVfB)


Test 2: Test 1:
wear rate: 3.12x1 O'3 cm/h wear rate: 3.12x1 O'3 cm/h
thickness: 4.45x10*4 cm/h thickness: 4.83x10"4 cm/h
weight: 4.7x1 O'5 cm/h weight: 3.4x1 O'5 cm/h

By comparing the results for all of the tests, it can be seen for the three different methods
that the weight loss measurement results in the smallest predicted wear. The thickness
method provided the second smallest wear prediction and the wear rate method predicted
the greatest amount of wear. The thickness measurement would appear to be the best
114

method because it does not require any assumptions in material properties, and the only
place for error to occur is in measuring the thickness of the thrust washer. By contrast,
the weight loss method requires the density of nylon 6,6 and the wear rate method uses a
wear factor, both dependent upon material properties. The density of nylon 6,6 is
expected to vary among samples so that the exact density of the sample used in the test
case is unknown; therefore, an average value for density was used instead. It appears that
the density of nylon 6,6 does not vary much between manufacturers or have much to do
with the processing method. In the case of the wear rate method, a wear factor is needed
to predict the wear rate. This wear factor is an averaged value for nylon 6,6 and does not
accurately represent the specimen used in this test case. This method is also dependent
on the PV, and it can be seen for every test case with the same PV that the wear rate is
identical no matter what conditions surrounded it. The results found in this exercise show
that all but one (weight loss method) of the wear rates methods yield excessive wear rates
according to ASTM D-3702. However, it is hard to say if the wear rate in the ASTM
standard is appropriate for the large PV values conducted in these tests.

5.1.3 Temperature Measurement using Nine Thermocouple Array

The Helios Temperature Recorder was used as the primary temperature measurement
device, to estimate the bulk surface temperature and to monitor the radial and axial
temperature variations in the steel washer using a nine thermocouple array. The locations
of the thermocouples in the steel washer were important in obtaining the best
approximation of the bulk surface temperature at the center of the contact area. They also
allowed for radial measurements away from the contact area. The thermocouples were
placed at three different depths to monitor diffusion of the heat away from the surface.
The second row of thermocouples was located at the same depth as the tribometer tester
115

thermocouple. The middle thermocouple in the second row is at the same radial and axial
proximity as the tester thermocouple, but shifted circumferentially by 90°. This allowed
for a direct comparison of the tribometer tester thermocouple to the thermocouple array
for this location. The middle thermocouple located on the top row can be compared to
the tribometer tester thermocouple in relation to providing a better experimental estimate
of the bulk surface temperature.

A more direct comparison of the results from the tribometer thermocouple and the
array thermocouple at the same location can also be observed (See Figures 5.31 and
5.32).

180

160

140

120

o
100
Array thermocouple broke
due to vibration

Tribometer
Array

40

0 2 4 6 8 10 12
Time (h)

Figure 5.31 Comparison of Tribometer Thermocouple and Array Thermocouple


at Same Location (Load = 44 N, V = 1.27 m/s)
116

50

45

40 - -I

35

o
30 -

20 --

15 --

10 --
— Tribometer
Array
5 --

0 2 4 6 8 10 12
Time (h)

Figure 5.32 Comparison of Tribometer Thermocouple and Array Thermocouple


at Same Location (Load = 444 N, V = 0.0508 m/s)

Since the one array thermocouple was located at the same location as the tribometer
thermocouple, it can be used to see if the smaller thermocouple with smaller hole in the
array provides different data compared to the larger tribometer thermocouple. It can be
seen that the array temperature measurements are higher than the tribometers’s by 7°C for
the high temperature case and by 2°C for the low temperature case. This proves that the
size of the thermocouple makes a difference. Due to the smaller size, the array
thermocouple did not affect the conduction path as much as the one thermocouple of the
tribometer; therefore, resulting in a more accurate bulk surface temperature. In the high
temperature test, the array temperature reading shows some instability early on and
117

eventually drops off just prior to the eight hour mark. This drop off point is where the
thermocouple broke due to vibration. The high and low temperature case shows a
symmetric resemblance between the two measuring devices in measuring data.

The temperature results of the two measuring devices taken at a steady state condition
can be seen for the entire test matrix in Table 5.4.

Table 5.4 Comparison of Tribometer and Array at Same Location


Thermocouple Temperatures (°C)

88 N. 0.254 m/s
TestTribometer:
1: 1112 N. 0.0508 m/s

77.5
T
est1:
Tribometer: 50.9
1: 222 N. 0.254 108.4
TestTribometer: m/s

Array: 87.6 Array: 50.1 Array: 106.1


% difference = 6.12 % difference = -0.79 % difference = -1.07

1: 444 N.47.10.0508 m/s TestTribometer:


TestTribometer: 1: 44 N. 0.508
48
m/s
TestTribometer:
1: 111 N. 0.508 m/s

95.2
Array: 49.1 Array: 56.4 Array: 99.8
% difference = 2.08 % difference = 8.05 % difference = 2.36
TestTribometer:
2: 41.3
TestTribometer:
2: 49.1
TestTribometer:
2: 106.5
Array: 45.8 Array: 56.4 Array: 111.1
% difference = 5.17 % difference = 6.92 % difference = 2.11

: 44 N.138.7
TestT1ribometer: 1.27 m/s

A ll Temperatures in °C
Array: 145.4
% difference = 2.36

It can be seen that the array bulk surface temperatures are higher than the tribometer
temperatures at most test conditions. The few test cases that do not follow this trend are
believed to be due to a faulty thermocouple bead or false junction in the wire of the array
thermocouple. The difference in temperature between the tribometer and array
118

thermocouples could be related to how well the tribometer thermocouple was inserted, or
if the array thermocouples were clear of any disturbances (metal-to-metal junctions)
below the bead. It was also observed that the temperature difference between the two
devices increased as the bulk surface temperature increased.

Another function of the array thermocouples was to try to obtain a more accurate
measurement of the bulk surface temperature by measuring temperature closer to the
surface. The distance of the closest row of the array from the surface was half that from
the tribometer thermocouple location to the contact surface. The array thermocouple was
expected to yield more accurate results than the tribometer thermocouple as can be seen
in Figures 5.33 and 5.34.

180

160

140

120

T 100
80

60

40
A rray
20

0
0 2 4 6 8 10 12
Time (h)

Figure 5.33 Comparison of Tribometer and Array Top Row Thermocouples


(Load = 44 N, V = 1.27 m/s)
119

50 --

40 --

S« 30

20 --

^ — Tribometer
10 -- Array

0 2 4 6 8 10 12
T im e (h)

Figure 5.34 Comparison of Tribometer and Array Top Row Thermocouples


(Load = 444 N, V = 0.0508 m/s)

The difference in temperature between the array and the tribometer measurements were
believed to be due to two effects: the array thermocouple was substantially closer to the
interface and that the array thermocouples were smaller in diameter, resulting in a smaller
hole and less interruption in the heat path. The difference in temperature for the high
temperature case is 10.2°C and 2.6°C for the low temperature case. These differences in
temperature are critical when using nylon 6,6 in an application under such conditions. If
the tribometer temperature is used as a guideline, the inaccurate lower temperature
prediction could lead to use of the material in a condition too extreme and have an
adverse affect on the performance of the material, or even result in premature failure.
120

A comparison of all the tests were performed to show the differences between the top
row array measurement and the tribometer measurement (See Table 5.5).

Table 5.5 Comparison of Tribometer and Array Top Row Thermocouple Temperatures

TestTribometer:
1: 1112 N. 0.0508 m/s
1: 88 N. 0.254
TestTribometer: 50.9
m/s
1: 222 N. 0.254 108.4
TestTribometer: m/s

77.5
Array: 89.8 Array: 50.8 Array: 107.9
% difference = 7.35 % difference = -0.10 % difference = -0.23

TestTribometer: 1: 44 N. 0.508
1: 444 N.47.10.0508 m/s TestTribometer: 48
m/s
TestTribometer:
1: 111 N. 0.508 m/s

95.2
Array: 49.7 Array: 58.7 Array: 106.3
% difference = 2.69 % difference = 10.03 % difference = 5.51
TestTribometer:
2: 41.3
TestTribometer:
2: 49.1
TestTribometer:
2: 106.5
Array: 46.7 Array: 56.2 Array: 117.1
% difference = 6.14 % difference = 6.74 % difference = 4.74

1: 44 N.138.7
TestTribometer: 1.27 m/s

A ll Temperatures in °C
Array: 148.9
% difference = 3.55

In the results shown above, a couple of cases can be seen where the array temperature is
not higher than the tribometer measurement as expected. These results are believed to be
due to the top row center thermocouple error. The largest percent difference in
temperatures is 10.03%, seen in the 44 N (10 lb), 0.508 m/s (100 ft/min) case. The
smallest positive percent difference is 2.69%, seen in the 444 N (100 lb), 0.0508 m/s (10
ft/min) case. It is interesting that both these high and low percent difference cases have
the same PV of 175 (kPa-m/s). It also can be seen for PV of 437 (kPa-m/s) that the
121

slowest speeds yield the largest percent difference in temperature and the fastest speeds
yield the smallest percent differences.

The final goal of the array thermocouple design was to understand the axial and radial
dissipation of heat through the steel washer. The array measurements would allow for a
better understanding of the effects of geometry, as well as the effect of convection near
the edges. The array of thermocouples used in the steel washer show the axial and radial
temperature distribution for high and low temperatures (See Figures 5.35 and 5.36). Note
that the data point for the middle row inner radius thermocouple would be expected to be
at a higher temperature, and therefore this data point is suspect.

N ylon sample contact area

110

109 --

108 --

107 -

rr 106 --

2 105 --
Top Row
104 - Middle Row
Bottom Row

103 --

102
101 -

100
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Radius (mm)

Figure 5.35 Array Temperature Contour at High Temperature


(Load = 222 N, V = 0.254 m/s)
122

N ylon sample contact area

49.8

49.6 --

49.4 --

49.2 -- Top Row


Middle Row
Bottom Row

E 48.8 --

48.6 --

48.4 --

48.2 --

48
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Radius (mm)

Figure 5.36 Array Temperature Contour at Low Temperature


(Load = 444 N, V = 0.0508 m/s)

The results shown by the temperature array contours indicate, for the most part, a
maximum temperature for each row at the middle column. This trend is plausible
because the middle column of thermocouples are located directly below the middle of the
contact area, whereas the first and third column lie on the inner and outer edges,
respectively. The first and third columns are not located under the contact area, thus are
subject to two dimensional conduction effects. It can seen in both test cases that the
bottom two rows have a sharper drop off at the inner and outer edges. This can be
attributed to convection and two dimensional conduction. The steel washer is pulling the
heat from the contact surface, and the heat is dissipating more readily in the cooler
123

material away from the surface, thus yielding a larger temperature drop compared to the
top row at the edges. The lack of symmetry between rows could be attributed to
inaccuracies in the thermocouples, although it is very possible that convection and
geometry issues are the main reason.

It can be seen from the experimental results that wear plays an important part in
measuring the bulk surface temperature. The results also imply that PV limit is heavily
dependent upon the conduction and convection of the test apparatus. From the
experimental bulk surface temperatures obtained, it was shown how the careful placement
and sizing of a thermocouple can yield a more accurate bulk surface temperatures that the
standard thermocouple on our tribometer tester. The location of thermocouples in an
array also indicated how the heat dissipates axially and radially through the steel washer.
This data was very helpful in building the numerical model described in Chapters 6 and 7.
124

THERMAL ANALYTICAL AND NUMERICAL

SIMULATION OF THE TEST DEVICE

Many analytical models have been developed to predict the surface temperature of a
sliding element on a stationary flat. There are two temperature types of interest when
modeling such a system. The first is bulk temperature, which is obtained from classical
thermal calculations, and the second is flash temperature. Analytically, three methods
have been used in the current study to predict the surface temperature. The first method
is a one dimensional conduction theory developed by Blok (1963) to estimate bulk
surface temperatures. The second method is a fin calculation, which considers one
dimensional conduction and convection. The third method is flash temperature theory,
also developed by Blok (1963).

The bulk surface temperatures can also be modeled numerically, and in this case the
finite volume code Fluent was used for modeling. A major challenge of numerical
modeling is the selection of appropriate boundary conditions, in this case chosen to match
experimental tests on the tribometer tester. The numerical models should be more
accurate than the analytical models, since they account for two dimensional geometric
and thermal variations in the experimental test apparatus. Three geometries of varying
complexity and accuracy have been used to represent the tribometer tests. The first
geometry investigated was the simplest case of only the two axisymmetric contacting
pieces. The second geometry investigated included the first geometry, and in addition it
modeled the axisymmetric pieces that hold the contacting pieces. The final geometry was
an axisymmetric representation of the entire experimental test apparatus. In all cases, a
125

uniform heat generation region was supplied at the sliding interface and the outer axial
boundaries were assumed to be at ambient temperature. These models will be compared
and contrasted to each other and to the experimental data obtained from the tribometer
system.

6.1 Geometry Effects on Conduction and Convection

The initial geometry chosen to represent the tribometer was only composed of the
lower half of the thrust washer and the whole stationary specimen (See Figure 6.1). This
model is axisymmetric, and can be rotated around the axis to represent the geometry as
three dimensions, but the governing equations are in two dimensions, r and y. By
building the model axisymmetric, the axis of the model defines the centerline of a
cylindrical mesh. To determine the appropriate physical value for a particular variable at
a point on the axis, Fluent (1995) averages the values at the centers of the cells
surrounding the node under consideration. This averaged value is then assigned to each
of the boundary cells bordering the axis in the circumferential direction. The
axisymmetric representation thus eliminates the singularity associated with the zero
control volume element area at the centerline. One drawback to this feature is that any
aspect of the geometry that is not axisymmetric, cannot be represented exactly in the
model.
126

th ru st w asher

"steel
w asher

Figure 6.1 Initial Geometry

The initial geometry was determined to be inadequate, since the upper and lower thermal
boundary conditions were unknown. The intent of the modeling work was to model a
large enough portion of the geometry that the upper and lower boundaries would be at
ambient temperature. However, it was clear, based on comparison with experimental
results, that the upper and lower boundaries of the geometry shown in Figure 5.1 were
actually at a temperature significantly above ambient. By setting the boundaries at
ambient temperatures the solver converged the solution to this temperature at just these
boundaries, independent of the generated heat input. To match the interface surface
temperatures found in the experimental work, the required convection coefficients were
substantially higher than those normally seen in a natural convection air environment (5-
15 W/m2*K). Also, as the heat generation was varied, the convection coefficients varied
considerably so that an average “typical” value could not accurately be determined.

The next level of modeling included the rotating specimen holder and the complete
thrust washer mated with the stationary specimen and its holder (See Figure 6.2). This
additional geometry allowed for better conduction and an increase in surface area allowed
for enhanced convection to take place, thus more cooling.
127

ro tatin g specinen
holder

Figure 6.2 Second Geometry Modeled

The results obtained for this model were considered to be more accurate than those for the
initial geometry. This was determined by the accuracy of the surface temperatures and
the average convection coefficients needed to obtain these surface temperatures.
However, this geometry still was not adequate, since the boundary conditions at the very
top and bottom (y-direction) of the model were still not actually at ambient temperature,
as determined from the experimental work. Therefore, there was a clear need to model
more geometry above and below the model shown in Figure 6.2 in order to achieve axial
boundaries that were actually at ambient temperature. The overall result of the second
geometry was that the interface temperature predictions are higher than the experimental
measurements, unless the convection coefficient is increased to reduce the interface
temperature. The problem resulting from increasing the convection coefficient to
compensate for the difference in surface temperatures between the model and the
experimental data is that it must be changed every time the heat rate changes and that the
values used exceed realistic values for natural convection in air.
128

It was realized that the whole Falex test apparatus needed to be modeled to determine
whether temperatures would converge in the +y and -y directions to the ambient
conditions, so that a non-ambient wall temperature at those locations would not need to
be specified (See Appendix B). This brought up several concerns when trying to model
the geometry using axisymmetry. The remaining geometry to be modeled had features
that were not axisymmetric, but important to the conduction and convection of the overall
model.

The results from the final model yielded results that were considered acceptable. The
criterion that made the final model acceptable was that the temperature for both the top
and bottom boundaries was at ambient, as confirmed by experimental measurements.
Another result from the final model of the full geometry was that the convection
coefficients used to match the predicted and experimental temperatures in the vicinity of
the sliding surfaces were in an acceptable range for natural convection in air and did not
change significantly with heat input at the sliding interface.

6.2 One Dimensional Conduction Analytical Calculation

In sliding systems there is an increase in the surface temperature due to frictional heat
generation. The increase in temperature can be predicted by calculating the bulk or
average surface temperature for a total generated heat rate. The surface area over which
the frictional heat is generated is ignored in the simplified one dimensional case. The
bulk surface temperature can result in melting of the counterface materials. In the current
analysis, the total heat (Q) is calculated from.

Q = (iWV [Watts] [ 6 . 1]
129

where:
g = coefficient of friction (measured)
W = controlled normal load [N]
V = controlled velocity [m/s]

Next, the total heat must be partitioned between the two surfaces. Ideally, the heat
partitioning is determined from the geometry and thermal properties of the counterfaces
and test structure. For the one dimensional case, the heat is assumed to be partitioned
solely based on the thermal diffusivities of the two counterfaces (Blok 1963),

Ql — OCQtotal [6 .2]
and
Q 2 — (l~O0Qtotal [6.3]

where:

a = --------- - [6.4]
(a, +0f2)

and
ai= ki/piCp,i, thermal diffusivity of lower material [m2/s]
« 2= k2/p 2Cp,2 , thermal diffusivity of upper material [m2/s]
k = thermal conductivity [W/m°C]
p = density [kg/m3]
cp = heat capacitance [J/kg°C]

Once the total heat generated is known from equation [6.1] and partitioned between two
contacting bodies has been estimated, one dimensional temperature variations can be
130

estimated. For example, for the tribometer tests, by using the subsurface temperature
measurement near the interface in the steel washer from the thermocouple located directly
under the contact area in experimentation, an estimate of the surface temperature can be
made given a geometry for both pieces, (See Figure 6.1). Using Fourier’s Law,

q " ~ — /v ~ ~ =
A* A* r , ) = -A,
j - [ W / m 2] [6.5]

where:
Ax = distance from interface into material 1 [m]
Tx = measured temperature at distance Ax from the interface [°C]
Ts = surface temperature of material at interface [°C]
ki = thermal conductivity of material 1 [W/m*K]
A] = cross-sectional area of material 1 [m2]

Thus, the estimate for bulk surface temperature is:

/ Ô1AxX
7=7 + [6.6]
Ai^i

Once an approximate geometry is determined in a design and an estimate of the


coefficient of friction is obtained, the interfacial temperature can be estimated.
131

6.3 One Dimensional Adapted Fin Calculation

Another method for estimating the surface temperature and axial temperature
distribution for a thrust washer in contact with a steel washer is to assume the test
structure behaves as two fins. The fin calculation allows for convection in addition to one
dimensional conduction, thereby adding realism to the model. The biggest drawback to
using a fin model is the geometry, since neither counterface is an ideal fin. The thrust
washer has a non-uniform axial cross-section and the stationary washer is not an infinite
fin. However, several assumptions were made to adapt the geometry to a fin model. For
example, only the lower annulus of the thrust washer was modeled as a fin (See Figure
6.1). The initial intent was to set boundary conditions based on measured temperatures.
However, accurate temperature measurements could not be taken at the base of the thrust
washer or the steel washer, due to the rotating specimen holder and the stationary
specimen holder (See Figure 6.2). Since it was difficult to set appropriate temperature
boundary conditions, the assumption was then made to treat both fins as infinite fins.
This means both pieces were of constant cross-sectional area sufficiently long that
ambient temperature was obtained at the end of each fin. The heat was partitioned
between the thrust washer and steel as follows,

Qgen = Qthrust + Qsteel [6.6]

where:
q gen = heat generation [Watts]
qthrust, qsteel = M [Watts]
M = (hPkAc)1/20b [Watts]
and
h = convection coefficient [W/m2oC]
132

P = perimeter of geometry [m]


k = thermal conductivity [W/m°C]
Ac = cross-sectional area [m2]
0b = Tb-Too [°C]
Tb = interface temperature (at base of each fin) [°C]

One can solve for the temperature at the base (Ty, See Figure 6.1), which is considered
the interfacial temperature. To solve for the temperature at x = L, the tip of the fin.

0
-T = e~mx [6.7]
Ob

where:
0 = T-Tm [°C]
m = (hP/kAc)1/2 (calculated)

The boundary conditions are that as length (x = L) goes to infinity then Tl-T=o= 0.

Practically speaking, infinite fin conditions hold approximately for L ) ------, since 0 is
m
x% of 0y at that length. The assumption that the geometry is constant is obviously
approximate, because it can be seen from the actual device that this is not true. We used
this infinite fin estimation using approximately 40 Watts (max. heat input for
experimental test range) of heat input and a convection coefficient of 12 W/m2*K
(averaged free convection coefficient) to determine how long the polymer and metal
would have to be in the +y and -y directions to be infinite. It was found that the polymer
would need to be 21 times longer and the metal needs to be 126 times longer, in fact this
length was longer than the entire measuring device in the -y direction. Therefore, the
133

infinite fin is not a valid approximation of the actual device. However, as presented in
Chapter 7, the infinite fin assumption yielded reasonable agreement. Conduction shape
factors were investigated to determine whether their use would allow the tribometer
geometry to be represented more accurately. However, it was seen that the parameters
that are implemented when using shape factors would contradict the parameters for the
infinite fin (Incropera and DeWitt 1990).

6.4 Flash Temperatures

Flash temperatures occur due to the fact that sliding contact usually occurs at small
asperities on the surface, so that the actual contact area is less than the apparent area.
When this occurs the frictional energy is dissipated in small localized volumes of
material. The instantaneous temperatures, dependent on asperity size, can flash up to 50-
100°C or more above the bulk interfacial temperature. These flashes result in localized
melting at the contacts which in turn can result in discoloration or reaction byproducts
that would not be predicted for the mating conditions. In subsection 3.1.2, Archard’s
(1958) theory on predicting flash temperatures (see eq. [3.6] and [3.8]) is presented. No
attempt was made in the current study to measure flash temperatures experimentally.

6.5 Two Dimensional Conduction Using Fluent

A numerical computer model was developed using a two dimensional axisymmetric


model of the tribometer test device, as discussed in section 6.1. As discussed in section
3.2, there are many options when trying to decide on a solver. Although finite element
methods have been widely used for thermal modeling, the available finite element
134

package was not sophisticated enough to cover all of the possibilities that were being
considered. A lack of knowledge of the boundary element method was a drawback to its
use, in addition to the very complex boundary conditions needed to execute this method.
Therefore, we chose to use a finite volume (finite difference) method, since a reliable
program was readily available. This program, written by Fluent Inc., is a general purpose
package solving integral conservation equations for energy and has the ability to handle
complicated geometries. Fluent can be used to model conduction, convection, radiation
and three dimensional sliding surfaces. For these reasons, Fluent version 4.3 was chosen
as the numerical solver to be used to model the tribometer system.

The first and most basic case modeled with Fluent was a two dimensional conduction
model representing the one dimensional case studied by Blok (1963). The purpose of this
model was to try and predict the interface temperature of the sliding materials, and to
compare and contrast results to Blok’s one dimensional theory. When only conduction is
taking place in the model, all the outer boundary walls are set at a constant temperature.
This temperature is usually an ambient temperature, or it could be a measured
temperature from experimental data. Conduction for this problem is solved by a form of
Fourier’s Law (see eq. [6.5]). This is a very simple model, which neglects convection
and radiation.

6.6 Two Dimensional Conduction and Convection Model Utilizing Fluent

The model was taken to another level of complexity by the addition of convection.
Convection brings more realism to the model, since there is always natural convection
unless a vacuum is involved. Under the conditions of the tribometer tester, there is
natural convection or forced convection from the rotating thrust washer and from the
135

cooling fan on the tribometer (if turned on). There were two options for modeling the
convection using Fluent. One was to allow Fluent to model the air flow surrounding the
tribometer apparatus and thereby the convection heat transfer. In this case flow regions
would be included in the model geometry and flow boundary conditions would be
needed. The second simpler option was to use Fluent to model the conduction in the
solid region only and to provide a convective boundary condition to the solid walls. This
was the method used in the current modeling. For convective boundary conditions,
Fluent (1995) uses the convection heat transfer coefficient (h) which is defined by
Newton’s Law of Cooling;

q” = h(T s-T a) [6.8]

where:
q = heat flux from the surface [W/m2]
h = convective coefficient [W/m2oC]
Ts = temperature at the surface [°C]
Ta = ambient temperature [°C]

Fluent requires that a user-determined convection coefficient and ambient temperature be


input as the convective wall boundary condition. Therefore, Fluent solves for the wall
temperatures. Whereas, in the conduction only case, instead of solving for the wall
temperatures, the temperatures were entered. The wall surface temperatures could not be
accurately measured during experimental work.
136

6.7 Modeling Approach and Model Verification

The approach to modeling the tribometer system using Fluent was based on three main
assumptions:

1. Axisymmetric geometry
2. Constant volumetric heat rate over the sliding contact area
3. Average convection coefficient over surface

The model was verified by showing grid independence (as discussed in more detail in
section 6.7.7) and by using experimental data to compare against Fluent’s results.

6.7.1 Convection Holes in Shaft

Geometry deviation from the idealized assumptions of two dimensional axisymmetry


was considered to be one of the biggest contributors to inaccuracy of the numerical
results. In sections of the experimental apparatus above and below the two sliding
surfaces, the geometry was very non-axisymmetric. Two areas of particular concern were
the bearings located high on the rotating shaft and the load plate located near the bottom.
The bearings located on the main upper shaft were very difficult to represent accurately in
the axisymmetric model, and they not only conducted heat away but also generated heat.
The bearings were modeled as a solid mass in the shape of a solid cylinder, estimated to
be the same size in area as the actual bearings. The large plate, used to apply load to the
samples and located at the bottom of the tribometer stationary shaft was also difficult to
represent accurately in the model. Again it was not axisymmetric, so an approximation of
its area was used to model it symmetrically as a ring (See Appendix A). Fortunately, it
137

was determined for all the test cases run that the bearings and load plate did not impact
the conduction heat transfer, since temperatures of the bearings and load plate were at
approximately the ambient temperature.

An additional non-axisymmetric aspect of the tribometer tester geometry was located


directly below the stationary specimen holder. Two sets of intersecting holes drilled in
the shaft below the holder, which were assumed to be present to aid in heat convection,
could not be modeled directly in our axisymmetric model. Therefore, to account for the
loss of material and change in conduction path, the diameter of the model was decreased
by assuming the volume of the holes were subtracted from the original shaft. We felt this
was the best approximation that could be made with an axisymmetric model. Two
models, one that accounted for the convection holes as described above and one without
the holes in the shaft, were compared (See Figure 6.3). The geometry shown in Figure
6.3b yielded surface temperatures up to 5 °C higher than the surface temperatures
predicted based on the geometry shown in Figure 6.3c. In addition, with the geometry of
Figure 6.3c, the lower shaft temperatures were considerably above those measured. It
was determined that the incorporation of the holes into the shaft was needed to obtain a
more reasonable approximation and a more accurate solution to the model.

ARTHUR LAKES LIBRARY - '


eOLORADO SCHOOL OF MINES
G8UBI.C0 80401 —
138

q) A c t u a l Falex b) Axisymmetric c) Axisymmetric


Hole G eometry Fa le x Hole Representation
Representation N e g l e c t in g Holes

Figure 6.3 Tribometer Convection Holes and Fluent Axisymmetric Representation

6.7.2 Constant Volumetric Heat Rate at the Sliding Contact Area

Modeling the heat generation at the sliding surface was not a straightforward process.
The first attempt was to input the heat source as a heat flux (W/m2) since the model was
in two dimensions. However this did not work, because the two dimensional model did
not correctly partition the heat into each piece, but instead it evenly divided the heat into
both pieces neglecting the differences in thermal conductivity and geometry. The second
attempt was to model the heat generation as a volumetric heat source, which required a
thickness for the heat source to be established. An actual thickness for the heat
generation region is not known nor can it be accurately measured. However, it seemed
reasonable to use a thickness similar to that of a surface film (10'5m). This thickness
could be any value within reason, but the larger the thickness the lower the maximum
temperature for a designated heat rate. The maximum temperature can be adjusted
139

through the convection coefficient, but for larger thicknesses the convection coefficient
becomes unrealistically high above what is generally acceptable for free and forced
convection in air (5-200 W/m2*K).

The volumetric heat rate (W/m3) is the total heat rate divided by the volume. The heat
rate is the product of the coefficient of friction, load and speed (see eq. [6.1]). The load
and speed are chosen to agree with experimental values, and the coefficient of friction is
taken as an average from experimental data once it reached steady state for the given load
and speed.

Another issue that affects the estimate of volumetric heat generation rate is the
distribution of the normal stress at the sliding interface. Normal stress is sometimes
assumed to be uniform, although it is typically Hertzian in reality, which means the
contact stress distribution is parabolic. For Hertzian contact, the maximum pressure, and
therefore the maximum heat generation is at the center of the contact, and has a
magnitude of,

W 'E '
Pmax — [6.9]
\ kR j

where:
pm ax = maximum surface pressure [MPa]
W ’ = W/unit length for square contact [N/mm]
W = normal load [N]
E’ = composite modulus [MPa]
1 l-u ,2 1 - 1>22
+
E' Ex E
140

R = composite radius [mm]


1 1 1
R ~ Rx R2

For our numerical simulation, the contact surface is treated as perfectly smooth. To
investigate the load distribution over the surface of contact a finite element analysis was
conducted on the contacting geometries (See Figure 6.4).

P r e s s u r e D istrib u tio n

S te e l W a s h e r

Figure 6.4 Pressure Distribution from Finite Element Analysis


141

The results shown in Figure 6.4 show that the pressure distribution predicted by the finite
element analysis program (Marc) is approximately uniform for a uniform load on the
upper surface of the thrust washer. Therefore, it was assumed that a uniform heat
distribution over the contacting surfaces for all of the simulations.

6.7.3 Average Convection Coefficient over the Surface

The local convection coefficient over a surface that is subjected to free, forced, or
mixed convection will vary. The tribometer test device was comprised of three different
regions dominated by different levels of convection. The first region is the rotating shaft
bearings above the thrust washer. These bearings are encased, the friction between the
bearings and the shaft tend to produce heat, and the average convection coefficient would
be expected to be low. The second region is the area below the rotating shaft bearings
and above the load bearings located below the stationary specimen holder. This region
would be expected to have a higher average convection coefficient with the cooling fan
on. If the cooling fan is off, then the convection coefficient would be expected to be
similar to that of the first region. The third region is below the second region and is more
directly subjected to the ambient room temperature. This area would be expected to have
a convection coefficient similar to that of the second region when the cooling fan is off.
In spite of expected differences in typical convection coefficients in three regions
described above, since there was no way to accurately determine local convective
coefficients, an average value was determined for the entire test device. Two different
average convection coefficient values were determined, one for the cooling fan on and
one for the fan off conditions on the tribometer tester. These average convective
coefficients, determined over a range of speeds and loads, were chosen so that good
agreement between predicted and measured temperatures near the sliding interface was
142

obtained. The process used to select the convective coefficients will be explained further
in section 7.4.

6.7.4 Ambient Temperature

Ambient temperature was initially assumed to be room temperature. Over the course
of testing at both high and low temperatures, it was noticed that the testing chamber
temperature was higher than room temperature. Since the ambient temperature used in
the model had a direct influence on the surface temperature, it was important to use an
accurate temperature. Therefore, a measuring system was devised to record the ambient
temperatures in the three regions described in section 6.7.3. One thermocouple was
placed on the housing for the rotating shaft bearings to measure its local ambient
temperature. This area recorded the highest temperatures of the three regions with the
cooling fan on or off. A second thermocouple was located in the chamber surrounding
the second region. This area recorded near ambient temperatures with the cooling fan on,
but the temperatures were an average of 4°C higher with the cooling fan off. A third
thermocouple was located in the vicinity of the third region. The ambient temperature for
this area showed little difference from room temperature when the cooling fan was on
because the surface temperatures were cool, relative to when the fan is off. However,
when the cooling fan was off the ambient temperature for the third region was an average
temperature of 30C higher than when the cooling fan was on. This region had the lowest
ambient temperatures of the three regions.

The ambient temperatures of the three regions did not vary as much as expected when
the cooling fan was on, with maximum variations of ±1°C over the range of heat
generation rates tested. However, in the cooling fan off case the ambient temperatures of
143

the three regions varied ±4°C over the range of heat generation rates investigated. An
averaged ambient temperature for all three regions was used to model situations for the
cooling fan on case. For the cooling fan off case an averaged ambient temperature for
each region, based on the measured values, was used.

6.7.5 Finite Difference Method

In the numerical model of the tribometer. Fluent (1995) solves the governing partial
differential equations for the conservation of energy. For two dimensional steady state
conduction with constant properties, no heat generation, and Cartesian coordinates, the
differential form of the heat equation is:

where the temperature distribution T(x,y) is determined in the medium. In solving the
heat equation, the dependent variable T is determined as a continuous function of the
independent variables x and y. The finite difference solution for a nodal system such as
shown in Figure 6.5 will yield temperatures at each node, so that resolution of the
solution will depend on grid refinement.
Figure 6.5 Two Dimensional Nodal System (Incropera and DeWitt 1990)

To solve the heat equation using a finite difference method, the geometry must be
represented by a nodal network. The finite difference equations to be solved at each node
are obtained by integrating the differential heat equation, given in equation [6.10]. For
d 2T
example, the term - r - r becomes:
ox'

[6. 11]
dx1 Ax

where the temperature gradients may in turn be expressed as a function of the nodal
temperatures. That is.

T _ T
cfT_ m + l./i m .n
[6 . 12]
dx m + l/ 2 .n Ax

and
145

'r rr
dT_ m .n m - l.n
[6.13]
dx A jc

This can be seen graphically in Figure 6.6.

m+ 1

Figure 6.6 Finite Difference Approximation (Incropera and DeWitt 1990)

Substituting equations [6.12] and [6.13] into [6.10], the following difference
approximation is obtained;

d 2T
[6.14]
dx2 (Ax)‘

By following the same procedure it can be shown that;

d 2T
dy2 Ay

T 1 ’T _ v ' I '7 '


m ./i+ l m .n - 1 m .n
[6.15]
146

By using a system for which Ax = Ay and substituting equation [6.14] and [6.15] into
equation [6.10], the finite difference equation for an interior node is obtained;

Tm, n+1 Trn, n-1 "f" Tm+i?n + Tm.i>n - 4Tm) n = 0 [6.16]

This approximate finite difference form of the heat equation may be applied to any
interior node that is equal distance from its four neighboring nodes. The form of the
finite difference equation for r-y coordinates will be similar.

6 .7 .6 Grid Dependence

Many different grid configurations were investigated in the model until one was finally
chosen. The chosen grid configuration used a concentrated mesh in the axial (vertical)
direction near the interface of the mating materials. This grid allowed for accurate
temperature calculations in the area where the steepest temperature gradients were taking
place. Radially (horizontally), the mesh was uniform in the contact area; away from the
contact area, it was denser closer to the contact area than in the rest of the surrounding
area. Typical grid configurations used in the model refinement are shown in Figures 6.7
and 6.8.

6 .7 .7 Convergence of Results and Verification Check

Temperatures were calculated at each of the many nodes throughout the model.
Several grid refinement tests were run to eliminate unnecessary nodes, thus reducing
model runtime. These tests were run with both high and low volumetric heat generation
147

rates to determine mesh refinement required at both the maximum and minimum heat rate
interface temperatures so that results differed by no more than 1°C with further
refinement. It was observed that the interface temperature did not change over the
temperature range up to the melting temperature for 8,300 to 21,266 elements over the
entire model. The element count was heavily weighted at the interface zone, since these
elements saw the steepest temperature gradients and needed to be tightly spaced to obtain
the maximum interface temperature. The maximum number of nodes that could be
placed radially (horizontally) in the contact area without significantly increasing the
solution time was 16. With 16 nodes, the nodal spacing in the contact area was
1.029 x 10'4m (See Figure 6.7).
148

Steel Washer

Thrust Washer

Figure 6.7 Dense Mesh Used in Grid Dependence Studies


(shown in Sliding Materials Only)

After many reductions in mesh density, the final model contained 6 nodes radially
(horizontally) across the contact area, so that each node was radially spaced 2.572xl0"4m.
This reduction in overall nodes still allowed agreement, to within 1°C, with the results
from the model with 16 nodes, but decreased the overall runtime by minutes (See Figure
6 .8 ).
149

Steel Washer
/

Thrust Washer

J ___________ L

Figure 6.8 Mesh Used in Numerically Simulated Results


(Shown in Sliding Materials Only)

A Fluent (1995) calculation of the energy equation is considered converged when the
difference equations are balanced at each node in the solution domain. Residuals are a
measure of the magnitude of the error in the solution at each iteration for a variable. In
general, the finite difference energy equation solution is considered to be converged for
enthalpy residuals on the order of IxlO'6, but the tribometer model required the enthalpy
residuals below IxlO'7 before convergence was satisfactory. Enthalpy is used in
convergence because the energy equation is being solved and enthalpy is the appropriate
form of energy represented in the energy equation. Temperature (T) is related to the
150

enthalpy through the material specific heat (Cp) and mass (m) according to: AH = mCpAT,
where H = enthalpy, which we are solving for.

Convergence is often deterred by a number of factors. Large numbers of


computational cells and overly conservative underrelaxation factors are often the main
causes. To speed convergence of the model, several techniques were used. First, the
number of nodes were reduced as much as possible, while still maintaining accuracy.
This allowed for fewer computational cells. The second factor to speed convergence was
underrelaxation.

Underrelaxation reduces the change in each variable allowed during each iteration.
Underrelaxation is often required due to the nonlinearity of the equation set being solved
by Fluent. It is not generally possible to obtain a solution by fully substituting the
improved values for each variable, which have been generated by the approximate
solution of the finite difference equation. Convergence can be achieved by iteratively
refining the value of the variable <}>p at node P, which is calculated from the old value,
4>p,oid, the computed change in <()p, A<|)p, and the underrelaxation factor, a, as follows
(Fluent 1995):

<1>p= 4>P,oid + <xA<])p [ 6 .1 7 ]

In Fluent (1995), the default underrelaxation parameters for the energy equation are set
at low values in order to ensure convergence in the largest possible number of cases.
However, the low default values of a do not give rise to the fastest rate of convergence,
and an improvement can often be obtained by an increase in a. The smaller the factor, a,
the heavier is the degree of underrelaxation and the greater is the degree of control
exercised over the change permitted from one iteration to the next. While small
151

underrelaxation factors damp out nonlinearities, they also serve to inhibit the rate of
convergence for more straightforward problems. The largest allowable factor for
underrelaxation, a = 1, was used in all cases presented in Chapter 7.

Fluent (1995) uses iterative techniques to obtain an approximate solution to the set of
discretized equations. Iterative methods of solution require much less effort than an exact
solution, but attention must be paid to solution accuracy at each iterative stage.
Controlling solution accuracy requires some basic understanding of the equation solution
procedure for enthalpy.

The default technique used by Fluent (1995) for enthalpy is the Multigrid (MG) Block
Correction. Block correction is generally recommended, and particularly for the enthalpy
equation, for the solution of coupled convection/conduction problems and in any problem
involving regions with highly different thermal conductivities. The block correction
technique speeds up convergence by applying a quasi-one-dimensional correction to the
current solution field in order to satisfy global conservation, with the correction vanishing
as the local balance is achieved throughout the domain. Since the MG block correction is
mainly intended for pressure equations, a line by line solution technique known as Line-
Gauss-Seidel (LGS) was used instead. In LGS the equations are solved simultaneously
for small groups of cells, one at a time. LGS is good at reducing local errors, but the
speed of the solution deteriorates with increasing grid size. This was not a problem with
the tribometer model, and the LGS was faster than the Multigrid block correction method,
with models converging in approximately 40 iterations.

This chapter has been intended to set the background needed to understand the thermal
model of the tribometer system developed in Fluent. The following chapter compares the
152

results found experimentally to those generated by Fluent and shows the capability of the
Fluent model as a predictive tool.
153

RESULTS AND DISCUSSION OF ANALYTICAL

AND NUMERICAL WORK

In Chapters 3 and 6 the analytical and numerical techniques used to estimate heat
transfer and surface temperature were described. The analytical techniques investigated
were one dimensional conduction, one dimensional fin, and flash temperature
approximations. The numerical technique used a two dimensional axisymmetric finite
difference model. In this chapter, the results generated by the analytical and numerical
studies will be compared to the experimental data from the tribometer tester to determine
what prediction method if any yields acceptable results. The comparisons first start with
the heat generation method, the heat partitioning scheme and how each model considers
the geometry of the model. Secondly, the numerical model results will be compared to
the experimental data. The third comparison is of the analytical surface temperature
predictions in comparison to the numerical surface temperatures. The final topic
presented is the use of the numerical model as a predictive tool for different sliding
materials.

7.1 Modeling Conditions for Analytical and Numerical Results

The modeling conditions needed to run the numerical simulations consist of several
factors such as: heat generation, heat partitioning and geometry. These factors help
distinguish between different conditions. For example, two different tests could have the
same pressure x velocity (PV), but have completely different coefficients of friction and
154

temperatures. A test matrix needed to be set up that covered a broad range for load and
speed, yet it was desired to hold PV levels constant over a range of speeds. The test
matrix was initially setup by running a PV test, and from that, determine other tests at the
same constant pressure x velocity.

7.1.1 Heat Generation Rate

The heat generation rate for all numerical cases used in this study uses the
relationship, Q = pWV. The flash temperature, bulk temperature and fin calculation all
use this same heat generation rate equation as well. The measured input needed from
experimental work was the coefficient of friction. The imposed normal load and velocity
are known and comparison of the experimental tests that resulted in the highest measured
bulk surface temperature will be used as example calculations. The heat generated at the
highest temperature is,

Q = 0.89(44.48)( 1.27) = 50.28 Watts [5.1]

where:
W = 44.48 N (10 lb)
V = 1.27 m/s (250 ft/min)

The heat generation rate for the numerical model had to be converted to a volumetric
heat generation rate, as described in chapter 6. The area of contact between the polymer
and steel washer was taken as the apparent area, but a thickness was needed to estimate a
volumetric heat generation. A thickness on the order of 10"5m (transfer film thickness)
was used, because it was the best approximation for the distance between contacting and
155

non-contacting asperities. It can be shown in Figure 7.1 that the boundary conditions
used to investigate the effect of the heat generation volume were insulated boundaries at x
= 0 and temperature was zero at x = L.

2.000E-01 8.00bE-01

! F L U E N T Grid File /• CO N FIG U R A TIO N = ddd • / i Sep 17 1996 j


« | Node Values Along J-Position = 1 j Fluent 4.32 i
I T em perature (K) Vs. 1-Direction Length (M) Fluent Inc. |

Figure 7.1 Volumetric Heat Generation Rate Convergence


(Temperature (K) vs. Length (m))

Other choices in thicknesses would result in a lower or higher surface temperature


depending upon the thickness. It can be shown from Figure 7.2 that if the thickness is
increased for the same boundary conditions, a higher surface temperature will be reached
at the x = 0 boundary independent of materials and interface conditions.
156

FLU EN T G rid File /• C O N FIG U R A T IO N = d d d l */ M ay 22 1997


L » N ode Values A long J-P osition = 1 R u e n t 4.32
T em perature (K ) Vs. I-D irection L ength (M) flu e n t Inc.

Figure 7.2 Increased Area of Volumetric Heat Generation Rate


(Temperature (K) vs. Length (m))

When using the volumetric heat generation rate in a thickness of 10"5m for the numerical
model, the convective coefficient values required to obtain agreement within
experimental data were considered to be reasonable for free and forced convection.

7.1.2 Heat Partitioning

Different heat partitioning techniques were used in the analytical and numerical
models. Some of the techniques involved geometry in determining the partitioning while
others did not. All the partitioning techniques involve the thermal properties of the
materials in sliding contact. The partitioning schemes for flash temperatures, bulk
surface temperature, fin calculation and the numerical model will be covered.
157

The flash temperature calculation was based on the assumption that all heat is supplied
to each body. For all cases, the contacts were considered elastic and the asperity area was
that of the steel washer. The choice of number of asperities is somewhat arbitrary to
determine from the varying heights obtained with a profilometer. The number appears to
range across the contact area, from approximately 5-25 asperities per 0.5 mm of unit
length. The five most prominent asperities are marked with a square while the remaining
possible asperities that could have contact are circled (See Figure 7.3).

2.5

1.25 -■

II 0.(E o.: Us

-1.25 --

-2.5
Length (mm)

Figure 7.3 Approximation of Asperities in Contact Area

In this study the contact area is assumed to be in the shape of a square, with multiple
circular contacts over the entire area. Therefore, using the equation for elastic contact
(assumed spherical) in a square contact area with five contacts. The contact radius at the
maximum experimentally obtained heat generation for the steel washer is.

fsW -JtV '3 3(0.00013 jv)(7.9xl0'6m)


= 6.2 x 10 7m [7.1]
[ 4 E ' ) { 4 2 6 X 1 0 % ) ,

where:
W = W/number of asperities over contact area = 44 N/339,250 = 0.00013 N
R = 7.9 x Iff6 m
E’= 3.26 x 109 N/m2

Therefore, for the maximum heat generation experimentally obtained for five contacts
across the contact area on the steel washer, the flash temperature was solved for a
stationary heat source. The mean temperature rise for the steel washer is,

0.0038 W
6 [3.6]

where:
Qb = Qgen/number of asperities over contact area = 50.28/13,332 = 0.0038 W
kb = 60.1 W/m*K
159

In the case of five contacts across the contact area on the thrust washer at the maximum
heat generation experimentally obtained, a slow moving heat source was used. The mean
temperature rise for the thrust washer is,

Q 0.0038 W
= = 4.(6.2 x 1 0 - m) ( 0 .2 4 3 % ^ ) = 6’3° 6 ^ ™

where:
Qc = Qgei/number of asperities over contact area = 50.28/13,332 = 0.0038 W
kc = 0.243 W/m*K
a = 6.2 x 10'7 m

For the maximum heat generation experimentally obtained with 25 contacts across the
contact area, the mean temperature rise for the steel washer is.

^ = 4 i k b = 4 (6 .2 x K) = ^ ^

where:
Qb = Qgen/number of asperities over contact area = 50.28/339,250 = 0.00015 W
kb = 60.1 W/m*K
a = 6.2 x 107 m

Likewise, for 25 contacts, the mean temperature rise for the thrust washer is.
a 0.00015 W
= 248.9 K [3.8]
4 a K 4 . ( 6 2 x 10"7m )(o .2 4 3 % .

where:
Qc = QgeVnumber of asperities over contact area = 50.28/339,250 = 0.00015 W
kc = 0.243 W/m*K
a = 6.2 x 10 7 m

These calculated mean temperatures are deceiving, since all the heat generated is assumed
to be into each of the bodies, not being partitioned between the two.

When partitioning the bulk surface temperature, the method requires the diffusivity of
both materials. For example, the diffusivity of the steel washer is.

k 60.1
a = = L59 x 10'5 [3.2]
pcp 7,872-481

where:
ksteei = 60.1 W/m*K
p = 7,872 kg/m3
cp = 481 J/kg*K

Once the diffusivity is determined for the polymer and the steel, heat between the two
materials can be partitioned,
161

1-5
a = 7— ^ --------- r = 7 ----------------- 1 -5 ! X l ° --------------— r = 0 . 9 2 [ 3 .3 ]
( 1^ x 1 0 = + 1 .4 2 x 1 0 " )

where:
O C ste e F 1 . 5 9 X 1 0 " 5 m 2/ s

O tn y lo n = 1 4 2 X 1 0 '6 H lV s

Thus, the partitioning into each body at the maximum experimentally obtained heat
generation is:

Qsteei = ccQtotai = 0.92 x 50.28 W = 46.26 W [6.2]


and
Qnyion = (l-cOQtotai = 0.08 x 50.28 W = 4.02 W [6.3]

The bulk surface temperature method does not take into account the geometry involved in
sliding contact, but it does partition between different materials.

When determining the partitioning factor for fin calculations, assuming an infinite fin
as described in chapter 6, the cross-sectional area and perimeter of both washers is:

Ac{steel] = n r l - n r ^ = n (0.0158 m f - n (0.0079 m f = 0.000588 m

Psteei - 2 n r 0 + 2k r( = 2k (o.0158 m) + 2k (0.0079 m) = 0.1493 m

and

A c(thru.u) ~ n r o = 7T(0.0143 m) —K (0.0127 m) = 0.000135 m2

Pthrust = 2 # ^ + 2k rt —2k (0.0143 m) + 2^: (o.0127 m) = 0.1698 m


162

A constant convection coefficient was used to simplify the results of the fin
calculation. The convection coefficient is the same averaged value used in the numerical
model to obtain agreement with experimental data. The maximum experimentally
obtained heat generation partitions into each body by,

qMe, = M = JhPkAceb = a/40-0.1493- 60.1 0.000588 (rt -2 6 ) = 0.4594^ -11.9 IV

Tb= 132°C [6.6]


and

q„$,„ = M = bJhPkAc6h = ^40 •0.1698 •0.243 •0.000135 (Tb - 26) = 0.0149 - ~ T b - 0.39 W

Tb= 132°C [6.6]


where:
h = 40 W/m2*K
Psteei = 0.1493 m
ksteei = 60.1 W/m*K
knyion = 0.243 W/m*K
Ob = Ty - Too
Too = 26 °C
Q stee i = Q n y io n = 50.28 W
Pthmst = 0.1698 m

The partitioning factor for the numerical model is not specified externally but, rather
determined by the model. The partitioning can be estimated using a one dimensional
approximation between the interface temperature and a temperature at a small distance on
both sides of the interface. Table 7.1 shows the partitioning for each washer.
163

Table 7.1 Heat Partitioning Values

F lash T em p eratu re:


Thrust W asher: 0.0 0 4 x Q gen
S teel W asher: 0.996 x Q gen_______
O n e D im e n sio n a l Bulk S u r fa c e T em p eratu re:
Thrust W asher: 0.08 x Q gen
S teel W asher: 0.92 x Q gen________
O n e D im e n sio n a l Fin C alcu lation :
Thrust W asher: 0.03 x Q gen
_____________ S teel W asher: 0.97 x Q gen_______
N u m erical M odel:
Thrust W asher: 0.04 x Q gen
_____________ S teel W asher: 0.96 x Q gen________

7.1.3 Geometry

Only two of the models; the fin calculation and the numerical model, take the
geometry of the contacting pieces into account. The fin calculation considers only the
contact area and the perimeter of the contact area. Since the contacting pieces are of two
different shapes the steel washer did not have the same perimeter as the thrust washer.
When solving for the heat generation rate of each material, the actual perimeters of both
pieces were used. The numerical model required an approximation of the complete
tribometer tester geometry. These issues were described in detail in Chapter 6 and the
geometries investigated are shown in Appendix B.

7.2 Comparison of Numerical Model to Experimental Data

The numerical model is the most representative model regarding boundary conditions
and geometry of the four models, since it utilizes full geometry and convection. The
164

numerical model was compared to the experimental temperature array thermocouples.


Numerical tests were run to simulate the previous analytical models; therefore, the
"A
numerical models were run with conduction only (one dimensional conduction bulk
surface temperature) and with conduction with free and forced convection (one
dimensional adapted fin calculation).

As discussed in Chapter 6, the numerical models evolved over the many test cases run,
and three geometries were used to simulate the tribometer tester such as; “small” and
“middle” geometry (See Figure 7.4) and the “actual” geometry (See Appendix B).

ro ta tin g s p e c im e n
h old er

t h r u s t washer thrust w asher

steel w asher

sta tio n a ry
specim en
"1 h o l d e r
steel
washer

Small Middle
Figure 7.4 Numerical Model of Small and Middle Geometries

Each geometry was run with three different boundary conditions. The first boundary
condition was a constant temperature boundary condition at 26°C. The second was a free
165

convection boundary condition, with a value assumed to be 13 W/m2*K. The third


boundary condition was that of forced convection, with a convection value assumed to be
40 W/m2*K, based on experimental data. It can be seen in Figure 7.5 that over the heat
generation range, differences in results were negligible for the temperature boundary
condition.

55

50 --

40 --

Small
Middle
35 -- Actual

30
0 10 20 30 40 50 60
H eat G e n e ra tio n (W)

Figure 7.5 Numerical Geometry Comparison with Conduction Only

Temperature over the complete heat generation range for all geometries with the free
convection boundary conditions can be seen in Figure 7.6. The full representation of the
tribometer yields the most accurate results in relationship to the experimental data (not
shown in Figure 7.6).
166

Temperature over the complete heat generation range was then predicted for all
geometries with the forced convection boundary condition, as shown in Figure 7.7. Once
again the full representation of the tribometer tester yields the most accurate results in
relationship to the experimental data (not shown in Figure 7.7).

1600

1400 --

1200 --

1000 --

5
1 •Sm all
M iddle
I A ctual
600 --

400 --

200 --

0 10 20 30 40 50 60
H eat G e n e ra tio n (W)

Figure 7.6 Numerical Model Geometry Comparison for Conduction


and Free Convection
167

600

500 -

400 --

Small
2 300 -- Middle
Actual

200 --

100 -

0 10 20 30 40 50 60
H e a t G e n e r a ti o n (W)

Figure 7.7 Numerical Model Geometry Comparison for Conduction


and Forced Convection

It was shown in Chapter 6 that of the three geometries, the most complete geometry is the
most accurate geometry. Since this geometry has been determined to have the best
accuracy, the boundary conditions can be compared to the experimental data (See Figure
7.8).
168

300

250 --

200 --

« 150-

Conduction
100 -- Free Convection
Forced Convection

50 --

0 10 20 30 40 50 60
H e a t G e n e ra tio n (W)

Figure 7.8 Numerical Model Boundary Condition Comparison


for Actual Geometry Representation

It would appear that the full size geometry with free and forced convection should yield
the most accurate and comparable results to that of the experimental data. Several
experiments were run using the tribometer tester fan on (forced convection), and
temperature data were taken up and down the tribometer tester shaft and at the tribometer
thermocouple (x = 0 cm). The experimental temperatures were then compared with those
predicted at the same locations by the numerical model (See Figure 7.9). The predicted
surface temperature for free convection is higher than the forced convection predicted
surface temperature due to a lower convective coefficient. Note the temperature rise
indicated at approximately 2.5 cm was due to high surface temperature measurements that
169

were not indicative of cross-sectional average temperatures in the region of the shaft with
cooling holes.

90

Experimental
o h=40
h=50
I h=60
h=70

0 --

20 --

10 --

-10 ■5 0 5 10 15 20 25 30
Distance (cm)

Figure 7.9 Comparison of Numerical Model and Experimental Results


for Forced Convection

The estimated temperatures for the numerical model are shown for each convection
coefficient, and were fitted to the experimental data by using a Least Squares fit (See
Table 7.2).
170

Table 7.2 Least Squares Fit of Temperatures for Forced Convection


Coefficients to Experimental Data

JZ
<o
o
h= 40 h= 50 h= 70 E xp erim en tal L o c a tio n (cm )

ii
25 25 24 24 25.9 -7.87
25 25 25 24 28.7 -3.18
27 26 26 25 32.5 -1.02
97 90 85 81 90 0.00
57 50 45 42 53 1.27
62 55 51 47 35.9 2.41
34 31 29 28 26.7 5.87
29 27 26 26 26.6 9.68
26 25 25 25 25.4 12.95
25 25 25 24 25.1 16.66
24 24 24 24 24.8 18.29
24 24 24 24 24.1 20.83
24 24 24 24 24.3 21.84
24 24 24 24 24.4 26.26
L east S q u ares: 851 450 383 411

It can be seen from the Least Squares fit that the best approximation for an averaged
convection coefficient is 60 W/m2*K. In Figure 7.9, it can be seen that the poorest
agreement occurred in the region where the tribometer had convective cooling holes
(between 1.5 and 2 inches). The poor agreement could be a combination of things such
as; a poor axisymmetric approximation of the convective cooling holes and a lack of
accurate temperature measurements in that area. The calculated convection coefficient
was considered very approximate, and it was determined that any further comparisons
would require a more accurate method for temperature measurement.

When the same test was run for free convection, the results were considerably worse,
since the upper and lower ends of the shaft did not reach ambient temperature due to the
higher temperatures seen with the free convection predictions. These higher temperatures
at the top and bottom of the model were the main reason the convection coefficient
significantly changes from test to test, making it difficult to find an average convection
171

coefficient for the selected test range. It was determined that this method of comparison
was not accurate enough in trying to determine a suitable convection coefficient for our
numerical model due to the inability of the numerical model to accurately predict
temperatures along the shaft compared to the experimental measurements.

Since the study being performed was focused on predicting an accurate surface
temperature, it was next decided to find an average convection coefficient that would
accurately predict temperature near the sliding surface. This was accomplished by using
only the experimental data taken by the thermocouple array for comparison with the
model.

The nine thermocouple array points were matched to the corresponding locations on
the numerical model. Thermocouples numbered 1-3 are the top row inner radius to outer
radius positions. The second row of thermocouples is numbered 4-6, from the inner
radius to the outer radius. The third row of thermocouples is numbered 7-9, from the
inner radius to the outer radius positions. These temperature points were internally
measured in the steel washer, whereas the last approach used surface measured
temperatures. The internal measurements are believed to be more accurate since they are
not directly influenced by convection. Temperatures at the nine locations were fitted by
least squares to the experimental data and a convection coefficient was estimated (See
Figure 7.10).
172

100

90 --

80 -

70 --

60 --
Temperature (C)

50 --

40 --

30 -- — Experimental
• h=30, Model
20 -- —A—h=40, Model
— h=50, Model
— h=60, Model
10 --

0 1 2 3 4 5 6 7 8 9
Thermocouple (#)

Figure 7.10 Rough Best Fit Estimate of Convection Coefficient for Forced Convection
Using the Thermocouple Array (Load = 1112 N, V = 0.0508 m/s)

The convection coefficient was narrowed down to between 30 and 40 W/m2*K and was
further refined by using a Least Squares method, which showed the convection coefficient
to equal 36 W/m2*K (See Figure 7.11).
173

89 --

88 --

87 --

r 86 -

E 85 --

84 -

Experimental
83 -- h=35, Model
h=36, Model
h=37, Model
82 --
h=39, Model

0 1 2 3 4 5 6 7 8 9
Thermocouple (#)

Figure 7.11 Refined Best Fit of Convection Coefficient for Forced Convection Using
the Thermocouple Array (Load = 1112 N, V = 0.0508 m/s)

For forced convection, a convection coefficient was determined for each test in this
manner. Over the test range the convection coefficients varied from h = 34 W/m2*K to
h = 44 W/m2*K, and an average coefficient of friction (h = 40 W/m2*K) was determined
for the test matrix (See Table 7.3).
174

Table 7.3 Averaged Convection Coefficients (W/m2*K) and PV (kPa-m/s)


for Forced Convection

Force (N)
44 88 111 222 444 1112 h(W/m2K)
0.0508 39 36
44.5 39.83
pv(175) pv(437)
(kPa-m/s) (kPa-m/s)
VELOCITY 0.254 40 41 40.5
(m/s) pv(175) pv(437)
(kPa-m/s) (kPa-m/s)
0.508 36 43 41.50
43 44
pv(175) pv(437)
(kPa-m/s) (kPa-m/s)
127 34 34
pv(437)
(kPa-m/s) h(ayg)
37.67 40 43.5 41 41.75 36 40.05

To determine sensitivity to the convection coefficient, the model was used to predict
surface temperatures, for a range of heat generation values, using h = 36 and h = 44
W/m2*K. As can be seen in Figure 7.12, the best fit convection coefficient values are
compared to the minimum and maximum convection coefficients seen throughout the test
range.
175

180

1 60 --

140 --

120 --

S
jjj 100 --

i 8 0 --

6 0 --

■ h = 3 6 (-2 s ig m a )
4 0 --
■ h = 4 4 (+2 s ig m a )
-#— h (ex p e rim e n ta i)
20 --

1 0 .1 7 1 0 .39 11.3 1 4 .6 9 2 7 .1 2 3 8 .9 8 4 2 .3 7 5 0 .2 8
H m al G é n é r a tio n (W a lla )

Figure 7.12 Comparison of hexpenmentai and hmax/min for Forced Convection

Using an overall average convection coefficient is more practical when trying to


predict the surface temperature at a given heat generation. By using an average
convection coefficient for a given heat generation range, experimental convection
coefficients do not have to be estimated for each experimental test. The difference in
surface temperature predictions between the actual individual test and average convection
coefficient can be seen in Figure 7.13.
176

180

160 --

140 --

120 --

a 100 -

40 --
h=40
h(experimental)
20 --

0 5 10 15 20 25 30 35 40 45 50 55
H eat G e n e ra tio n (W atts)

Figure 7.13 Comparison of the Experimental and Average Convection Coefficient


for Forced Convection

Comparing the model with measured data along the shaft for free convection was not
considered to be accurate enough as described earlier. However, the second method of
comparing with the data from the array of thermocouples was utilized. Using this
method, an averaged convection coefficient value for the test matrix in free convection
was h = 13 W/m2*K. The free convection study will not be explored beyond this point,
and all test cases in the remainder of this chapter will refer to forced convection.

The design of the thermocouple array in the steel shaft allowed for a more accurate
surface temperature and a better understanding of the conduction through the steel
washer. The arrangement of the thermocouples in the radial direction allowed
177

measurement of the temperature near the edges of the contact and a temperature in the
middle of the contact path. These points can be shown as rough temperature contours
(See Appendix B).

The arrangement of the array of thermocouples in the steel washer allowed for an
observation of the way the heat is pulled through the steel washer and into the rest of the
lower shaft. Data at these nine points were used to fit the numerical model to the
experimental data, as discussed earlier, and comparisons could be made between the array
thermocouples and the tribometer tester thermocouple to the numerical model
predictions. A comparison between the temperature of the center top row thermocouple
on the tribometer tester and the numerical model temperature prediction at the same
location is shown (See Figure 7.14).

160

1 4 0 --

120 --

1 00 --

6 0 --

40
N u m e ric a l M odel (h = 40)
E x p . T o p R ow C e n te r T h e rm o c o u p le
20 -

0 10 20 30 40 50 60
H e a t G e n e r a t i o n (W )

Figure 7.14 Comparison of Top Row Array Thermocouple


and Numerical Models Temperature
178

The numerical model predictions are in good agreement with the experimental data of the
top row center thermocouple location showing a maximum error of 2.5%. The
importance of the top row center thermocouple is that it is believed to provide the best
experimental approximation of the interface temperature.

The same comparison can be made between the temperature indicated by tribometer
thermocouple and the numerical models prediction at the same location (See Figure 7.15).

160

140 --

120 --

100 --

o
£
2

60 --

40 --

— Numerical Model (h = 40)


— Exp. (Tribometer Thermocouple)
20 -

0 10 20 30 40 50 60
Heat Generation (W)

Figure 7.15 Comparison of Tribometer Thermocouple


and Numerical Model Temperature
179

The numerical model does not match the experimental data point as well as it did the top
row center thermocouple data. It is important to be able to model the tribometer tester’s
data due to its use in industry, and since the tester does not have an array of
thermocouples to take measurements closer to the surface. Thus, the tribometer tester
measurement is the best bulk surface temperature approximation available
experimentally, and the numerical model will use this experimental data to fit the
temperatures at the same location on the model.

The numerical model also predicts the interface temperature, which is compared in
Figure 7.16) to the maximum array thermocouple reading (at the center location).

180

160 --

1 4 0 --

120 --

2
I

6 0 --

N u m erical S u rfa c e T e m p e r a tu r e P re d ic tio n


40 -

Exp. T o p R ow C e n te r T h e rm o c o u p le P re d ic tio n

20 --

0 10 20 30 40 50 60
H e a t G e n e r a t i o n (W )

Figure 7.16 Comparison of Maximum Array Temperature and Numerical


Model Surface Temperature
180

The top row array thermocouple is as 10-15°C difference lower than that of the predicted
sliding surface temperature. This is due to the subsurface location of the thermocouple
and the combination of conduction and convection across that subsurface area.

A comparison of the numerical model surface temperature at both the overall forced
convection average and the best fit individual test convection coefficient is shown in
Figure 7.17.

180

160 --

1 40 --

120-•

80 --

6 0 --

4 0 --
B e s t Fit In d iv id u al T e s t
O v e ra ll A v e ra g e

20 --

0 10 20 30 40 50 60
H e a t G e n e r a t i o n (W )

Figure 7.17 Numerical Model Surface Temperature for hactuai and IWrage

The difference between the best fit convection coefficient and the average coefficient is
small across the test range at a PV of 437 (kPa-m/s). This implies that the average
coefficient would provide acceptable prediction of sliding surface temperatures.
181

It was initially believed that the analysis of the best fit convection coefficient data
would show convection coefficient dependence on factors such as: speed, load,
temperature and coefficient of friction. Figure 7.18 shows the best fit convection
coefficients (shown as data points) in comparison to speed. The solid line indicates the
overall average convection coefficient, and the dotted lines represent convection
coefficient values at two standard deviations above and below the average. Figure 7.19
shows the best fit convection coefficients as a function of normal load.

50

45

40

35
♦ ♦

30

f
I 25
20

15
♦ D a ta
a v e ra g e
10 (-) 2 s ig m a (b o tto m )
(+) 2 s ig m a (top)

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Speed (m/s)

Figure 7.18 Comparison of Best Fit Forced Convection Coefficient and Speed

Figure 7.19 shows the best fit forced convection coefficients and normal load, and Figure
7.20 shows the best fit forced convection coefficients and temperature.
182

50
45

40

35

30
S
t -
20

15
D ata
10 -av e ra g e
-(-) 2 s ig m a (b o tto m )
(+) 2 s ig m a (top)
5

0
0 200 400 600 800 10 0 0 120 0
N o r m a l L o a d (N)

Figure 7.19 Comparison of Best Fit Forced Convection Coefficient and Normal Load

50

45

40

35

30
2
I 25
20

15
D ata
-av e ra g e
10
-(-) 2 s ig m a (b o tto m )
-(+ ) 2 s ig m a (top)
5

0
0 20 40 60 80 100 120 140 160
T e m p e r a t u r e (°C)

Figure 7.20 Comparison of Best Fit Forced Convection Coefficient and Temperature
183

Contrary to earlier expectations, the convection coefficient did not appear to depend on
speed, load or temperature. The previous graphs show no dependence on any of the
factors.

7.3 Comparison of Analytical Surface Temperature Predictions

Three different analytical approaches were performed in this study. The first
analytical approach was one dimensional conduction, second was one dimensional fin
calculation and third was flash temperature. These methods were compared to
predictions of the numerical model in order to determine how well they predict bulk
surface temperatures relative to the experimentally obtained data.

7.3.1 Flash Temperature

The flash temperature theory takes into account that both bodies acquire heat from the
surface; therefore, the partitioning of the two bodies will yield the interface temperature
(see eq. [3.10]). For the maximum experimentally obtained heat generation at 5 contacts
across the contact area, the flash temperature theory yields:

i = i +i =^ k +^ =m t i or25AK [310]

For the maximum experimentally obtained heat generation at 25 contacts across the
contact area, the flash temperature theory yields:
184

T =j + { = ~ d i ï + ^ k ï = ik or0S9 K [310]

The temperatures found in the flash temperature theory are higher in comparison to the
numerical and experimental data. The flash temperature calculation calculates the mean
temperature rise above the bulk surface temperature. The increase in temperature is
attributed to the heavy localized loads at the contact areas.

The flash temperatures calculations for 25 contacts provided a low estimation of the
mean temperature rise over the bulk surface temperature. Therefore, 5 contacts across the
contact area were used to estimate maximum expected flash temperature for all the tests
in the experimental matrix with much better results (See Table 7.4).

Table 7.4 Calculated Flash Temperatures Above the Bulk Surface Temperature
for 5 Contacts Across Contact Area

Normal Load (N)


44 88 111 222 444 1112
0.0508 5.1 °C 13.6 °C
5.2 °C

VELOCITY 0.254 5.6 °C 19.5 °C


(m/s)
0.508 7.1 °C 19.5 °C
7.3 °C 21.2 °C

1.27 25.4 °C
185

7.3.2 One Dimensional Conduction Analytical Calculation

In one dimensional conduction bulk surface temperature calculations an experimental


data point from below the surface is needed to apply to Fourier’s Law to predict the
interface temperature (see eq. [6.4]). Two different subsurface temperatures have been
used to estimate surface temperatures using one dimensional conduction. The first,
considered to be most accurate, is the top row center thermocouple from the array in the
steel washer. The second is the tribometer thermocouple. In both cases, the one
dimensional model surface predictions will be compared to the numerical model
predictions, since these are assumed to be the most valid (See Table 7.5).

Table 7.5 Thermocouple Temperatures for Test Matrix of Tribometer


and Array Top Row Center (°C)

0.0508 m/s 0.254 m/s


444 N 1.112 N 88 N 222 N
00 CO

T ribometer: 47.1 °C 77.5°C Tribometer: 108.4°C


tn in
P P
d

Top Row Center: 49.7°C 89.8°C Dp Row Center: 107.9°C


d

T ribometer: 41.3°C
Top Row Center: 46.7°C

0.508 m/s 1.27 m/s


44 N 111 N 44 N

Tribometer: 58.5°C 95.2°C T ribometer: 138.7°C


Top Row Center: 78.3°C 106.3°C Dp Row Center: 148.9°C

T ribometer: 49.1°C 106.5°C


Top Row Center: 56.2°C 117.1°C
186

An example of the calculated bulk surface temperature for the tribometer tester
temperature, at the maximum experimentally obtained heat generation is,

f QAx'' ( 50.28 -0.00236


7 = 7 + = 138.7 + = 151.1 °C [6.5]
\ Ak j 1.24 xlO"4 -60.1

where:
Ax = 0.00236 m
A s, e e l = 1.24X l O "4 ! ! ! 2

Tx= 138.7°C
k s te e i = 60.1 W/m*K
Q = 50.28 W

An example of the calculated bulk surface temperature for the top row array
thermocouple temperature, at the maximum experimentally obtained heat generation is,

QM ( 50.28 0.00118 )
7 = 7 + = 148.9 + = 156.9 °C [6.5]
Ak V-24 x 10™4 -60.1 )

where:
Ax = 0.00118 m
Aste=i= 1.24 x 10"4m2
Tx= 148.9°C
ksteei = 60.1 W/m*K
Q = 50.28 W
187

The bulk surface temperatures calculated for both one dimensional conduction cases are
lower than the numerical surface temperature prediction. The one dimensional calculated
interface temperature based on the top row center array thermocouple is 5°C higher than
the same calculation based on the tribometer tester thermocouple data. This provides
good reason to believe that the steel washer has significant two dimensional conduction
and convection effects. The values calculated for the bulk surface temperature method
are shown in Table 7.6.

Table 7.6 Calculated Bulk Surface Temperatures from the Tribometer


and Top Row Array Thermocouple (°C)

0.0508 m/s 0.254 m/s


444 A/ 1.112 N 88 N 222 N

T ribometer: 49.7°C 84.2°C Tribometer: 53.7°C 118°C


Top Row Center: 52.3°C 96.5°C Top Row Center: 53.6°C 117.5°C

T ribometer: 43.8°C
Top Row Center: 49.2°C

0.508 m/s 1.27 m/s


44 AZ 111 N 44 N

Tribometer: 51.5°C 104.8°C T ribometer: 151.1°C


Top Row Center: 62.2°C 115.9°C Top Row Center: 161.3°C

Tribometer: 52.7°C 116.9°C


Top Row Center: 59.8°C 127.5°C
188

7.3.3 One Dimensional Adapted Fin Calculation

The fin calculation provides a simple estimate of conduction and convection in


determining the interface temperature. The fin calculation, much like the flash
temperature theory, uses a partitioning of heat between the two bodies (see eq. [6.6]). As
described in Chapter 6, the infinite fin boundary condition provides the best
representation of the geometry of the tribometer tester.

The fin calculation required a very simplified model of the geometry. The fin equation
calls for the perimeter of each fin. In the case of the thrust and steel washer, the contact
area is the same but the perimeters are different. Three cases of different average
perimeter were used to estimate the steel washer to see what would best represent the
tribometer tester results. The “small geometry” is where the perimeters of both the thrust
and steel washer are the same size, which is the perimeter size of the thrust washer. The
“actual geometry” uses the actual perimeters of the steel washer and the thrust washer.
The “middle geometry” uses the actual perimeter of the thrust washer and an average of
the thrust washer and steel washer perimeters for the steel washer perimeter.

The fin calculation convection coefficient used the same average convection
coefficient as that was used by the numerical model, or 40 W/m2*K. This value of the
convection coefficient does not provide the best agreement of the fin calculation with the
numerical model prediction for surface temperature for this geometry, since at the slightly
lower value of the convection coefficient (h = 38 W/m2*K) the interface temperatures are
a better match. However, the improvement in agreement is small, so all calculations were
made for h = 40 W/m2*K to be consistent with the numerical model.
189

An example of the interface temperature fin calculation is shown below. Tb is the


“base temperature”, which occurs at the interface of the sliding temperatures. It can be
seen from the maximum experimentally obtained heat generation for an infinite fin and
actual geometry size, the temperature at the base (Tb) or interface is:

^ gen (7 up Q down

50.28 = (0.0149 Tb - 0.39) + (0.4594 Tb - 11.9)

Tb= 131.90C

The three geometries are compared to each other for the entire heat generation range
(See Figures 7.21). Also included is the numerical model surface prediction, since the
model was shown to be in good agreement with experimental data.

250

200 --

fT 150 -

100 --

1♦ " S m a ll
50 - - « — M iddle
—« —A ctual
N u m e ric a l S u r fa c e T e m p e r a tu r e

0 10 20 30 40 50 60
H e a t G e n e r a ti o n (W )

Figure 7.21 Fin Geometry Comparisons to Numerical Surface Temperature


190

From the previous figures, the middle geometry appears to provide the best fit. Table 7.7
shows fin calculation predictions for all tests for the three geometries and the numerical
surface predictions.

Table 7.7 Infinite Fin Calculations of Bulk Surface Temperatures (°C)


for Three Geometries

Small Middle Actual Numerical


444 N, 0.0508 m/s: 63.7 48.6 47.4 51.8
444 N, 0.0508 m/s: 64.6 49.1 47.9 54.4
1112 N, 0.0508 m/s: 131.6 91.2 83.2 97.5
88 N, 0.254 m/s: 68.3 51.4 49.8 55.8
222 N, 0.254 m/s: 179.1 121 108.2 128.8
44 N, 0.508 m/s: 80 58.8 56.1 63.3
44 N, 0.508 m/s: 81.8 59.9 57 64.8
111 N, 0.508 m/s: 179.1 121 108.2 134.8
111 N, 0.508 m/s: 192.7 129.5 115.3 139.8
44 N, 1.27 m/s: 224.4 149.4 131.9 158.8
Least Squares: 11.344 366 2.112

From the Least Squares analysis, it can be seen that the “middle geometry” yields the
most accurate results.

7.4 Comparison of Analytical and Numerical Surface Temperature Predictions

A comparison of the analytical methods to the numerical model will be made to


determine which analytical method is the most simulative (See Figure 7.22).
191

200

180 --

160 --

140 --

120 --

S 1 0 0 --

80 --

60 --
1-0 Tribometer
1-0 Top Row Center
40 - Fin Calculation
Flash Temperature
Numerical Surface Temperature
20 --

0 10 20 30 40 50 60
H eat G e n e ra tio n (W)

Figure 7.22 Comparisons of Analytical and Numerical Results

The comparison revealed some interesting results. The numerical model best
represents the best estimate of interface temperature, since the convective boundary
condition was chosen so that the model predictions would agree with the experimental
thermocouple and array data. Of the three approximate analytical models, the one
dimensional conduction model based on the top row center thermocouple data and fin
calculation agreed best with the numerical predictions. The fin calculation might be
expected to be the best representation because it attempted to model the geometry of the
tribometer tester, as well as convection, which was not attempted by the one dimensional
conduction method. The fin calculation yields reasonable results for this test range of
speeds and loads with a maximum error of 7%, however it would be expected to perform
192

more poorly in higher or lower heat generation test ranges unless either the perimeter o r.
convection coefficient were adjusted. The one dimensional top row center calculation
was based on the temperature data from the thermocouple closest to the surface. It shows
that if the distance is short enough between the measured data and the surface, then the
heat transfer near the interface is approximately one dimensional.

The one dimensional conduction bulk surface temperature was not as accurate when
using the tribometer thermocouple temperature. It is clear that the one dimensional
approximation becomes better as the subsurface data on which the calculation is based
are taken closer to the surface. At the location of the tribometer thermocouple, the error
is significant enough that the one dimensional conduction method is not a very good
means to estimate surface temperature.

The flash temperatures are much higher than any of the other interface predictions and
experimental data taken. This is due to the definition of flash temperature, which is the
mean temperature rise above the bulk surface temperature. It is difficult to know if the
flash temperature is accurate to the interface.

7.5 Numerical Model as a Predictive Tool

The results from section 7.4 show that the numerical model can accurately match the
experimental data for temperatures in the steel washer, so that it is expected to provide a
good prediction for an interface temperature. This leads us to believe that the model can
be used to predict surface temperatures for polymeric thrust washers in contact with a
wide range of materials in place of the steel washer. A graph of the surface temperature
193

as a function of the thermal conductivity of the stationary washer is shown (See Figure
7.23).

1000 --

100 --

10 --
1 (W a tts)
10 (W a tts)
3 0 (W a tts )
5 0 (W a tts )

0.1 1 10 100
T h e rm a l C o n d u c tiv ity (W/m*C)

Figure 7.23 Comparison of Surface Temperature and Thermal Conductivity


for Polymer-on-Material

It is important to note that the estimated surface temperatures above 270°C would be
truncated due to this being the melt temperature of nylon 6,6.

The numerical model was used to estimate surface temperatures over a range of
stationary washer conductivities and heat generation values for a polymer thrust washer.
The original range of thermal conductivities was a low of 0.11 W/m*K for ceramic, to a
middle range of 237 W/m*K for aluminum, and a high range of 430 W/m*K for silver.
194

The higher the thermal conductivities were, the lower the surface temperature, unless
extremely large heat generation values were used. The results shown in Figure 7.24 yield
the best approximate to obtaining a melt temperature.

■ 400-450
0350-400
■ 300-350
0250-300
■ 200-250
0150-200
10 (Watts) □ 100-150
■ 50-100
■ 0-50
Heat Generation

1 (Watts)
Thermal Conductivity
(W/m*K)

Figure 7.24 Numerical Model as a Predictive Tool for a Polymer-on-Material

It is important to note that a surface temperature can be estimate when knowing the
coefficient of friction (Q = |LiWV), or the coefficient of friction can be estimate knowing a
surface temperature.

Chapter 7 has provided results that have shown a good correlation between the
numerical results for an averaged convection coefficient and the experimental data. It
was also shown how three analytical model (one dimensional conduction, one
195

dimensional fin calculation and flash temperatures) compare to the numerical results. Of
the three analytical methods, the one dimensional conduction model using the top row
thermocouple data provided the closest results compared to the numerical results, but it
was fully dependent upon experimental data. The fin calculation also provided
reasonable results, but further studies showed that the model was not very reliable when
using heat generation rates higher or lower than our current test matrix.
196

CONCLUSIONS AND RECOMMENDATIONS

This study was conducted to gain a better understanding of the factors that dictate the
failure of a material under sliding conditions. The study also set out to develop a
modeling tool to predict the surface temperatures using either a subsurface temperature or
coefficient of friction from the experimental data. Experimental and numerical test cases
were performed using a rotating nylon 6,6 thrust washer mated against a stationary steel
washer.

The main contributions of this study were:

• accurate numerical simulation of a thrust washer sliding on a steel washer type


of system

• comparison of analytical, numerical and experimental work for this type of


sliding system

• PV limit is temperature controlled and strongly dependent upon geometry and


material properties

The experimental temperature results were measured using two separate temperature
recording devices. This was done to check the validity and accuracy of the tribometer
temperature recording and yield a better approach to measuring the bulk surface
temperature. The metallic counterface was remachined such that a comparison could be
made at the same location of the two measuring devices and at a location closer to the
sliding surface compared to the tribometer thermocouple measurement.
197

Analysis of experimental results yield the following comments:

• Temperature behavior is dependent upon the speed of the rotating thrust


washer for a given PV. In both cases the temperature and coefficient of
friction almost doubled as the speed increased but the PV remained the same.

• The largest difference in temperature between the tester and top row array
thermocouple was at slower speeds for both PV values investigated.

• The top row array thermocouple did yield more accurate bulk surface
temperatures than the tribometer tester, which shows that even in such a small
distance, the surface temperature could be as much as 10-20°C higher than the
tribometer tester measurement.

Analysis of analytical results yielded the following comments:

• Accuracy of one dimensional conduction analytical calculations for estimating


bulk surface temperature depends upon the thermocouple depth from the
sliding surface, geometry (2-D conduction) and convection effects.

• The tribometer thermocouple appears to be too far from the surface, so that the
effects of convection caused for poor surface temperature predictions.

• The top row center thermocouple appeared to be located close enough to the
contact surface that convection did not have much effect on the surface
temperature predictions in comparison to the numerical results.

• accuracy of one dimensional adapted fin calculations for estimating bulk


surface temperatures is dependent upon fin geometry and the variation of the
convection coefficient for a wide range of heat generation

The analytical methods of both the one dimensional conduction and one dimensional
fin approximations met some of the initial conditions of the study; one, they were easy to
use and two, they did not entail any additional cost over the purchase of the tribometer
tester.
198

Analysis of numerical results yielded the following comments:

• The most detailed and complete model of the actual tribometer tester geometry
yielded the most accurate results.

• The numerical model can be used as a predictive tool to predict temperatures


over a wide range of heat generation rates and thermal conductivités.

Two general approaches were made in developing this study. An experimental


investigation was performed to try to understand the underlying factors in the failure of
nylon 6,6. Also an analytical and numerical study was undertaken to investigate the
possibility of predicting surface temperatures using either a one or two dimensional
model. Now that the study has been completed, several recommendations can be made to
help achieve our original goals. Even though the experimental results and numerical data
yielded useful results, future work could be done to minimize the impact of assumptions
on error and to obtain a wider range of data.

Recommendations for future work include:

• A speed of 0.254 m/s (50 ft/min) and forces between 667 and 890 N (150-200
lb) could be used to accomplish the melting temperature of nylon 6,6 without
excessive wear occurring to the nylon 6,6 thrust washer.

• To design a tribometer tester that has a simpler geometry. Non-axisymmetric


convection cooling holes or fins off the main spindles could be eliminated.
Ideally, identical shaft geometry for the upper and lower parts would be made
in order to provide a more one dimensional like geometry.

• By minimizing the convection and lowering the thermal conductivity of the


spindle material, experimentally obtained surface temperatures should be
much higher than in the current tribometer tester.
199

• Simplifying the geometry would make it easier to numerically model the


system, therefore allowing for a more accurate model and a better
approximation of the surface temperature.

If the options mentioned above for achieving the melting temperature for nylon 6,6,
are not practical, another option would be to chose a polymer that is less robust. There
are many polymeric materials with lower melting temperatures than nylon 6,6. By
choosing a material with a lower melting temperature than nylon 6,6, such as;
polyethylene, then catastrophic failure can be obtained at much less severe test
conditions.

Another issue that needs to be addressed for the experimental work is the fragileness
of the array thermocouples. A more robust thermocouple size would be desirable for the
rigorous test cases, but the accuracy and size of the current thermocouples can not be
jeopardized. The present thermocouples could have a protective sheath applied over the
beads of the thermocouples to make them more rigid and less susceptible to breakage
during testing. As long as the sheath has a thermal conductivity comparable to that of
metal, this would ensure a comparable level of accuracy to what is obtained currently
without the sheath. It would also be helpful to apply a rigid protective layer to the
thermocouple wire at the point of contact where the thermocouples are supported to
prevent them from falling out due to vibrations, since the thermocouples had a tendency
to become fragile at this point of contact.

This investigation studied temperatures in a steady state condition both experimentally


and numerically. It would be of great interest to be able to model the transient effects of
the experimental data, but it was not within the project scope. Any future experimental
and numerical work in this area would be beneficial. Another area to be further studied
would be that of polymer-on-polymer, steel-on-polymer and steel-on-steel. It would be
200

interesting to see if the same average convective coefficient developed from the polymer-
on-steel experiments would apply to these cases.

The analytical fin calculation and the numerical model proved to represent the
experimental measurement areas well, but the assumptions in their geometries result in
inaccuracies. The numerical model has much more to offer in trying to build the most
realistic model possible. To achieve the maximum accuracy for the numerical model a
three dimensional model should be run. This would eliminate some of the geometrical
inaccuracies present with the two dimensional axial symmetrical case. In a three
dimensional model the convection cooling holes could be identically modeled along with
any other non-symmetrical feature of the tribometer tester that were either eliminated or
approximated by an axisymmetrical representation. The numerical model’s software
allows for the modeling of moving mesh in three dimensions. This moving mesh
together with a gridded flow solution can calculate the convection coefficient. Ideally,
the moving mesh and gridded flow solution would model heat transfer and fluid flow
simultaneously. Thus, allowing the numerical model to solve for the convection
coefficient so that a known convection coefficient does not have to be supplied as a
boundary condition. Modeling radiation would also add another degree of realism to the
model.
201

REFERENCES CITED

Anderson, J.C., and E. J. Robbins 1978. The Influence of Temperature Generation on


the Wear of Some Polymers. Wear of Non-Metallic Materials. Proceedings of the Leeds-
Lvon Symposium on Tribology. 3rd Meeting, England, pp. 94-98.

Archard, J.F. 1958. The Temperature of Rubbing Surfaces. Wear, Vol. 2, pp. 438-455.

Archard, I.E., and R.A. Rowntree 1988. The Temperature of Rubbing Bodies: Part 2,
The Distribution of Temperatures. Wear. Vol. 128, pp. 1-17.

Amell, R.D., P.B. Davies, J. Hailing and T.L. Whomes 1993. Tribology: Principles and
Design Applications Spinger-Verlay New York Inc., pp. 1-255.

ASTM D 3702-90 1990. Standard Test Method for Wear Rate of Materials in Self-
Lubricated Rubbing Contact Using a Thrust Washer Testing Machine. American Society
for Testing and Materials. Philadelphia, PA

Barber, J.R. 1969. Thermoelastic Instabilities in the Sliding of Conforming Solids.


Proceedings of the Royal Society of London. Ser. A, 312, pp. 381-394.

Berry, G.A. and J.R. Barber 1984. The Division of Frictional Heat—A Guide to the
Nature of Sliding Contact. Journal of Tribology. Vol. 106, July, pp. 405-415.

Blok, H. 1963. The Flash Temperature Concept. Wear. Vol. 6, pp. 483-494.

Clerico, M. 1969. A Study of the Friction and Wear of Nylon against Metal. Wear. Vol.
13, pp. 183-197.

Ettles, C M., O.S. Dine and S.J. Calabrese 1994. The Effect of Frictionally Generated
Heat on Lubricated Transition. Tribology Transactions. Vol. 37, pp. 420-424.

Ettles, C M. and Hardie, C E. 1988. The Friction of Some Polymers and Elastomers at
High Values of Pressure X Velocity. Transactions of the ASME. October, vol. 110, pp.
678-684.
202

Ettles, C.M.M. 1987. Polymer and Elastomer Friction in the Thermal Control Regime.
ASLE Transactions. Vol. 30, 2, pp. 149-159.

Floquet, A., D. Play and M. Godet 1977. Surface Temperatures in Distributed Contacts—
Application to Bearing Design. Journal of Lubrication Technology, pp. 277-283.

Fluent Inc. 1995. Theory. FLUENT User Guide. Version 4.3. Lebanon, Vol. 4, Chapter
19, March, pp. 19-1 to 19-128.

Fluke Mfg. Co. 1990. HELIOS PLUS: 2287A Data Acquisition Front End. System
Manual. Vol. 1, Section 1-4, February.

Furber, K., J.R. Atkinson and D. Dowson 1976. Wear Mechanisms for Nylon 66.
Institute of Tribology. Proceedings of the 3rd Symposium on Tribology. Leeds, England,
Sept., pp. 25-31.

Furey, M.J. 1964. Surface Temperatures in Sliding Contact. ASLE Transactions. Vol. 7,
pp. 133-146.

Gecim, B. and W.O. Winer 1986. Effect of a Surface Film on the Surface Temperature
of a Rotating Cylinder. Journal of Tribology. Vol. 108, January, pp. 92-97.

Greenwood, J.A. and J.B.P. Williamson 1966. Contact of Nominally Flat Surfaces.
Proceedings of the Royal Society of London. A, Vol. 205, pp. 300-319.

Incropera, F.P. and D P. DeWitt 1990. Fundamentals of Heat and Mass Transfer. John
Wiley & Sons, Inc. 3rd ed., New York.

Jaeger, J.C. 1942. Moving Sources of Heat and the Temperature at Sliding Contacts. L
Royal Society of N.S. Wales. Vol. 76, pp. 203-224.

Kennedy, F.E. 1981. Surface Temperatures in Sliding Systems—A Finite Element


Analysis. Journal of Lubrication Technology. Vol. 103, pp. 90-96.

Kennedy, F.E. 1982. Single Pass Rub Phenomena—Analysis and Experiment. Journal
of Lubrication Technology. Vol. 104, October, pp. 582-588.

Kennedy, F.E. 1984. Thermal and Thermomechanical Effects in Dry Sliding. Wear,
Vol. 100, pp. 453-476.
203

Kennedy, F.E., F. Colin, A. Floquet, and R. Glovsky 1983. Improved Techniques for
Finite Element Analysis of Sliding Surface Temperatures. Proceedings. 10th Leeds-Lvon
Symposium on Tribology. Butters worth, London.

Lancaster, J.K. 1971. Estimation of the Limiting PV Relations for Thermoplastic


Bearing Materials. Tribology. vol. 4(2), pp. 82-86.

Ling, F.F., and S.L. Pu(1964), “Probable Interface Temperatures of Solids in Sliding
Contact. Wear. Vol. 7, pp. 13-34.

LNP Engineering Plastics, LUBRICOMP: Internally Lubricated Reinforced


Thermoplastics and Flouropolvmer Composites Bulletin 254-691.

Ludema, K.C. 1980. Evaluation of PV Limits for Plastic Bearing Materials,” ASTM
Spec. Tech. Publication. STP 701.

Omega Engineering Inc. 1995. The Temperature Handbook Vol. 29, Section H.

Rashid, M. and A. Seireg 1987a. Heat Partition and Transient Temperature Distribution
in Layered Concentrated Contacts, Part I-Theoretical Model. Journal of Tribology. Vol.
109, July, pp. 487-495.

Rashid, M. and A. Seireg 1987b. Heat Partition and Transient Temperature Distribution
in Layered Concentrated Contacts, Part U-Dimensionless Relationships and Numerical
Results. Journal of Tribology. Vol. 109, July, pp. 496-502.

Sonn, H. W., C.G. Kim and C.S. Hong 1995. Transient Thermoelastic Analysis of
Composite Brake Disks. Journal of Reinforced Plastics and Composites. Vol. 14,
December, pp. 1337-1362.

Tanaka, K., and Y. Yamada 1984. Effect of Temperature on the Friction and Wear of
Some Heat-Resistant Polymers. Conference: Polymer Wear and Its Control. American
Chemical Society Svnoposium. Series 287, pp. 103-128.

Theberge, J.E. 1970. A Guide to the Design of Plastic Gears and Bearings. Machine
Design. February, 5, pp. 111-120.

Vick, B., L.P. Golan II, and M.J. Furey 1994. Thermal Effects due to Surface Films in
Sliding Contact. Journal of Tribology. Vol. 116, pp. 238-246.
204

Vick, B., M.J. Furey, and S.J. Foo 1991. A Boundary Element Thermal Analysis of
Sliding Contact. Numerical Heat Transfer Part A: Applications. Vol. 20, pp. 19-40.

Vroegop, P H., H.H. Vermeulen, and R. Bosma 1980. The influence of Temperature,
Speed and Roughness on Dry Sliding Friction and Wear of Nylon 6.6 Against Steel.
Friction and Traction. Proceedings of the 7th Leeds-Lvon Symposium on Tribology.
Leeds, England, Sept. 9-12.

Wantabe, M. and H. Yamaguchi 1986. The Friction and Wear Properties of Nylon.
Wear. Vol. 110, pp. 379-388.

Yamaguchi, Y. and Kashiwagi, K. 1982. The Limiting Pressure-Velocity of Plastics


Under Unlubricated Sliding. Polymer Engineering Science. Vol. 22(4), pp. 236-241.
205

APPENDIX A Variations of Tribometer Test Geometry

This appendix shows drawings of the various tribometer tester geometries investigated
for modeling purposes in Chapters 6 and 7.
206

1.8 25 all dimensions in inches

R o ta tin g S h a f t

.70

T h r u s t W asher
S t e e l W ash er

.375 .19

.575

S t a t i o n a r y Specimen Holder
.25
dia. 0.3675
6.0 dia. 0.3675

.975

Figure A .l Front View of Tribometer Tester (Not to Scale)


207

rotating s h a f t

t h r u s t w ash er

s t e e l w asher

y
y— load bearing s h a f t
(no cooling h o l e s )

Figure A.2 Axisymmetric Representation of Tribometer Tester


208

/ —b earings

y— r o t a t i n g s h a f t

t h r u s t washer

s t e e l washer

y — cooling h o le s
[—^ y — load bearing s h a f t
load p l a t e

Figure A.3 Axisymmetric Representation of Tribometer Tester


with Bearings, Cooling Holes and Load Plate
rotating s h a f t

t h r u s t w asher

s t e e l washer

cooling h oles
load bearing s h a f t
load

Figure A.4 Axisymmetric Representation of Tribometer Tester


with Cooling Holes and Load Plate
210

APPENDIX B Temperature Contours for Different Geometries at Experimental


Maximum Heat Generation from Numerical Model

This appendix shows temperature contours at the experimental maximum heat


generation, of the tribometer tester geometry, from the numerical model for the three
axisymmetric geometries shown in Appendix A.
211

I s g
Tf l— l

S
CL
<

yt/3
<9
U
II
Z
o
§
Qd I
D
O gON
E <N
z II
o
u

CN
uu 2 +
•o w ù
"C O NO
O
H I 5
Z
U SL »
D
LL

+ +
C
+
o
+
o
+ +
o
+
o
+
o
+ + +
o
+ X
U.' tu LJ US US US US US US US US US UJ US US
r- TT sC sc r~ Os
Tf
o\
r-i
00 so
rn
rr
rn
<N
rS
O
M
.

lr& 3 U * i£ >

Figure B.l Temperature Contour at Experimental Maximum Heat Generation


for Geometry with Bearings, Cooling Holes and Load Plate
4.29E+02

for Geometry with Load Plate


FLUENT Grid File /* CONFIGURATION = case3 */ Apr 24 1997
Temperature (K) Fluent 4.32

Figure B.2 Temperature Contour at Experimental Maximum Heat Generation


Max = 4.286E+02 Min = 2.990E+02 Fluent Inc.
213

| s j
3 = 5
n I - I e

ton
«3
U

Z
O
5 o(N+
ûd U
D on
O C
E o\
z <N
o II
u

— OJ
3 - W
5 | £
H 2
Z
UJ 8. il
D £ C
<L> x3
tL H S

5 5
c
UJ ?.’
U ?oU
L ?U
L
X

O
rr

Figure B.3 Temperature Contour at Experimental Maximum Heat Generation


for Geometry with Cooling Holes and Extended Load Plate
214

E s u
c
G Tf
CN gg
k.
CL
< E E

#
m
ir .


Il
Z
o
CN
5 O
or +
3 UJ
g
E oON
CN
z II
o
u C

CM
u. ?
UJ
<U nC
o E c-,
H 2 Tf
Z u II
m CL
D X
£ C3
U
u. h-

?
UJ
§
UJ
?
Lti
1LU ?
UJ
1
UJ
?
UJ
?
u
?
UJ
?
lti
?
UJ
?
UJ UJ UJ
X

<N rr rr <c >o oc On


<N Os oo wn rr q Os
rr rT r-1 rn f^i r*S rn m m ri

Figure B.4 Temperature Contour at Experimental Maximum Heat Generation


for Actual Geometry
215

*
en
jst/3
il
Z
O
CN
< O
cd +
D
w
o
O O'
On
E CN
z II
o
u C

CN

2w yj
?
<U VO
o
H
z 1 5
t i ë. 11
D
Eu 1 1

mm.

Figure B.5 Close up Temperature Contour of Contacting Surfaces at Experimental


Maximum Heat Generation for Actual Geometry
216

CN
CL

en

T3

OC

o c
UJ

Figure B.6 Temperature Contour at Melting Temperature of Nylon 6,6


for Actual Geometry
217

(N
CL

u oc
U OC

c
o

Figure B.7 Close up Temperature Contour of Contacting Surfaces at Melting


Temperature of Nylon 6,6 for Actual Geometry

You might also like