Está en la página 1de 175

EXPERIMENTAL INVESTIGATION OF TRIBOLOGICAL

CHARACTERISTICS OF WATER-LUBRICATED
BEARINGS MATERIALS ON A PIN-ON-DISK TEST RIG

Yuriy Solomonov

The University of Adelaide


School of Mechanical Engineering

Master of Philosophy Thesis

April 2014

Abstract

ABSTRACT
Friction is the most fundamental phenomenon accompanying the sliding motion of solid
bodies. Friction, vibration, and wear under conditions of contaminated water lubrication
are extremely important in many engineering applications such as water-lubricated
bearings, water pumps, and braking systems.

The aim of this project is to investigate the factors that could lead to an improvement in the
performance of water-lubricated bearings materials.

Previous studies have revealed the main factors contributing to power loss are frictioninduced vibrations, and wear in water-lubricated bearings. Those factors are the result of
contamination of the lubricant (sea water); bearing alignment (parallelism of the shaft and
shell); material characteristics; and condition of the contact (sliding) surfaces. The contact
mechanics of the water-lubricated bearings as well as the performance characteristics of
the bearings components on which friction is exerted also have a substantial influence on
the tribological characteristics of water-lubricated bearings materials.

Thus, the focus of the present study is on the effect of water contamination on the friction
coefficient, vibration, wear and the vibrationwear relationship under varying operational
conditions.

An experimental program was conducted to develop new methods and investigate the
effect of water contamination on the tribological characteristics of pairs of materials under
different operational conditions for water-lubricated bearings.

A Pin-on-Disk test rig was designed and built to adopt the operational environment of a
real water-lubricated bearing. This test rig was used to obtain experimental data regarding
the effect of water contamination on the long-term behaviour of the bearing systems, and
to investigate the friction, vibration, wear, and vibration-wear characteristics of the
materials. The effect of various parameters, such as the friction conditions, damping, and
operational environment on the behaviour of the bearing materials was also investigated.

ii

Abstract

The experiments demonstrated that all three factors, namely contamination, material
properties and surface conditions, have a significant influence on the tribological
characteristics of water-lubricated bearings. It was also demonstrated that when the
operation of the water-lubricated bearing takes place in boundary and mixed regimes, the
adhesive and abrasive mechanisms of friction are significant and contribute to the
generation of excessive wear and vibration. This is contrary to what is claimed by many
manufacturers. It was observed that the wear mechanism in the water-lubricated bearing
materials was associated with low-frequency vibrations and severe contamination of the
lubricant. Also, as expected, the vibrationwear relationship of the water-lubricated
bearing materials was significantly affected by the contamination of the lubricant and can
be changed by magnetic field damping.

The present study identified the primary mechanism responsible for the high friction
coefficient, vibration, and wear to be a three-body mechanism caused by the abrasive
nature of the water contaminant. It was found that there was a significant increase in the
friction coefficient, vibration, and specific wear rate at the slowest sliding speed of 0.393
m/s. This is due to the boundary regime of lubrication, the adhesive-abrasive wear
mechanism, and specific material properties of NF22 (Railko) material. It was also
explored and reported that for a specific applied load of 8 N, at low and high sliding
speeds, and water contamination levels, damping has a strong effect on the vibrationwear
relationship which is also dependent on sliding speed and, as a result, on the lubrication
regime.

The significance of this experimental study is to improve the selection of water-lubricated


bearings materials and as a result, improve their performances. The outcomes of this
research project are:

Analysis of the existing types of materials, and experimental models and techniques
for modelling and simulating the operational conditions of water-lubricated
bearings

Identification of the existing problems associated with the contemporary


technology of water-lubricated bearings materials

Development of an experimental methodology and technique for the application of


a Pin-on-Disk test rig and determination of the main contributing factors
iii

Abstract

Identification and analysis of various lubrication and operational conditions for


water-lubricated bearing materials and systems and development of further
recommendations for future work.

iv

Declaration

DECLARATION

I, Yuriy Solomonov, certify that this work contains no material which has been accepted
for the award of any other degree or diploma in any university or other tertiary institution
and, to the best of my knowledge and belief, contains no material previously published or
written by another person, except where due reference has been made in the text.

I give consent to this copy of my thesis, when deposited in the University Library, being
made available for loan and photocopying, subject to the provisions of the Copyright Act
1968.

I also give permission for the digital version of my thesis to be made available on the web,
via the Universitys digital research repository, the Library catalogue and also through web
search engines, unless permission has been granted by the University to restrict access for
a period of time.

Yuriy Solomonov

Acknowledgements

ACKNOWLEDGEMENTS
The following work presented in this thesis would have not been possible without the help
and invaluable support of my family, university staff, and my friends and colleagues from
ASC.

I am truly thankful to my Principal Supervisor, Mr Ian Brown, for his support, assistance
and effort in the planning, experimentation and guidance throughout this entire project. Mr
Brown, despite a heavy workload and looming deadlines, was most helpful and would
never hesitate to take the time to answer questions and responsively sort through various
ideas and problems.

Tatiana, my wife, has been extremely supportive in many ways and deserves a special
mention. Thank you for being patient with me when it counted most and encouraging me
to complete my thesis, especially during the most difficult period of my candidature.

The effort of my co-supervisor A/Prof. Reza Ghomashchi, throughout the final stage of my
project is also greatly appreciated.

The support of many people within the School of Mechanical Engineering, including
academic staff, mechanical and electronic workshop staff, postgraduate students, and
technical and administrative support staff, is also acknowledged.

I would also like to thank my friends and colleagues from ASC who have contributed in
one way or another throughout my candidature.

vi

Table of contents

TABLE OF CONTENTS
ABSTRACT ......................................................................................................................... II
DECLARATION ................................................................................................................ V
ACKNOWLEDGEMENTS ............................................................................................. VI
TABLE OF CONTENTS ................................................................................................ VII
LIST OF FIGURES .......................................................................................................... XI
LIST OF TABLES ........................................................................................................... XV
CHAPTER 1

INTRODUCTION .................................................................................. 1

1.1 Overview .......................................................................................................................... 1


1.2 Scope and objectives ........................................................................................................ 5
1.3 Thesis outline ................................................................................................................... 7
CHAPTER 2

BACKGROUND AND LITERATURE REVIEW .............................. 9

2.1 Historical overview of friction ......................................................................................... 9


2.2 Fundamentals of friction ................................................................................................ 14
2.2.1 Dry regime of friction ......................................................................................... 14
2.2.2 Hydrodynamic regime ........................................................................................ 17
2.2.3 Boundary and mixed regimes ............................................................................. 19
2.3 Friction in water-lubricated bearings ............................................................................. 20
2.3.1 Lubrication regimes ............................................................................................ 21
2.3.2 Friction-induced vibrations ................................................................................. 24
2.4 Numerical models of friction in water-lubricated bearings ........................................... 25
2.4.1 Linear models ..................................................................................................... 27
2.4.2 Non-linear models............................................................................................... 29
2.5 Vibrationwear relationship in dynamic systems .......................................................... 30
2.5.1 Problem of vibrationwear dependency ............................................................. 31
2.5.2 Effect of damping on vibration and wear ........................................................... 32
2.6 Concepts of experimental apparatus and experimental technique ................................. 34
2.6.1 Pin-on-Disk system ............................................................................................. 34
vii

Table of contents

2.6.2 Effect of lubricant contamination ....................................................................... 35


2.7 Summary and research gaps .......................................................................................... 36
CHAPTER 3

PREVIOUS EXPERIMENTAL STUDIES OF TRIBOLOGICAL

CHARACTERISTICS IN WATER-LUBRICATED BEARINGS ............................... 38


3.1 Introduction .................................................................................................................... 38
3.2 Review of wear study .................................................................................................... 38
3.3 Review of the aft bearing study ..................................................................................... 40
3.3.1 Experimental apparatus....................................................................................... 40
3.3.2 Experimental results of the friction coefficient measurements .......................... 42
3.3.3 Experimental results of the wear tests ................................................................ 44
3.3.4 Discussions and conclusions of the experimental tests ...................................... 44
3.4 Conclusions .................................................................................................................... 45
CHAPTER 4

EXPERIMENTAL APPARATUS ...................................................... 48

4.1 Introduction .................................................................................................................... 48


4.2 Design requirements ...................................................................................................... 48
4.3 Variable sliding speed and applied load ........................................................................ 51
4.3.1 Variable sliding speed ......................................................................................... 51
4.3.2 Variable applied load .......................................................................................... 53
4.4 Bending arm ................................................................................................................... 54
4.5 Load cell ........................................................................................................................ 58
4.6 Data acquisition system ................................................................................................. 60
4.7 Water supply system ...................................................................................................... 61
4.8 Experimental test rig ...................................................................................................... 62
4.9 Experimental methods ................................................................................................... 64
4.9.1 Experimental programs ....................................................................................... 64
4.9.2 Samples preparation ............................................................................................ 64
4.9.3 Validation study .................................................................................................. 65
4.9.4 Investigation of friction ...................................................................................... 66
4.9.5 Wear experiments ............................................................................................... 66
4.9.6 Vibration-wear relationship study ...................................................................... 67
4.10 Microscopy examination.............................................................................................. 68
viii

Table of contents

4.11 Conclusions .................................................................................................................. 68


CHAPTER 5

MATERIALS ........................................................................................ 69

5.1 Introduction .................................................................................................................... 69


5.2 Effect of water contamination on bearing materials ...................................................... 71
5.3 Materials for water-lubricated bearings ......................................................................... 72
5.3.1 Polymer-based thermoplastic materials .............................................................. 73
5.3.2 Carbon - fibre reinforced materials based on thermosetting materials ............... 75
5.3.3 Shaft materials .................................................................................................... 77
5.4 Conclusions .................................................................................................................... 79
CHAPTER 6

VALIDATION STUDY........................................................................ 80

6.1 Introduction .................................................................................................................... 80


6.2 Test plan and procedure ................................................................................................. 80
6.3 Results and discussions .................................................................................................. 81
6.4 Conclusions .................................................................................................................... 83
CHAPTER 7

EXPERIMENTAL INVESTIGATION OF FRICTION

CHARACTERISTICS....................................................................................................... 85
7.1 Introduction .................................................................................................................... 85
7.2 Experimental study of the effect of contamination on friction: experimental plan and
procedure ............................................................................................................................. 85
7.3 Experimental study of the effect of contamination on friction: results and discussions 87
7.3.1 Investigation of friction under water lubrication ................................................ 87
7.3.2 Investigation of the friction coefficient and the effect of contamination ........... 92
7.4 Conclusions .................................................................................................................... 96
CHAPTER 8

EXPERIMENTAL INVESTIGATION OF WEAR .......................... 99

8.1 Introduction .................................................................................................................... 99


8.2 Experimental study of wear: experimental plan and procedure .................................... 99
8.3 Experimental study of wear: results and discussions................................................... 101
8.3.1 Wear tests .......................................................................................................... 101
8.3.2 Specific wear rate calculations ......................................................................... 108
8.4 Conclusions .................................................................................................................. 112
ix

Table of contents

CHAPTER 9

EXPERIMENTAL INVESTIGATION OF VIBRATIONWEAR

RELATIONSHIP ............................................................................................................. 114


9.1 Introduction .................................................................................................................. 114
9.2 Operational conditions and experimental methods ...................................................... 115
9.3 Experimental study of effect of water contamination on vibrationwear relationship:
results and discussions ....................................................................................................... 119
9.3.1 Friction forcetime analysis.............................................................................. 119
9.3.2 Power spectral density analysis ........................................................................ 125
9.3.3 Vibrationwear analysis ................................................................................... 132
9.4 Conclusions .................................................................................................................. 140
CHAPTER 10

CONCLUSIONS AND RECOMMENDATIONS ........................... 144

10.1 Summary .................................................................................................................... 144


10.2 Conclusions ................................................................................................................ 147
10.3 Recommendations for future work ............................................................................ 149
REFERENCES................................................................................................................. 150
APPENDIX A: PIN-ON-DISK ASSEMBLY DRAWINGS ......................................... 158

List of figures

LIST OF FIGURES
Figure 1.1 Diagram showing a simple tribological system ................................................... 1
Figure 1.2 Diagram showing the complicity and complexity of tribological processes
(Materials Tribology laboratory, 2008) ................................................................................. 2
Figure 1.3 An example of an aft bearing system (Solomonov et al., 2010) .......................... 4

Figure 2.1 Friction experiments suggested by Leonardo da Vinci. (Krim, 2002) ............... 10
Figure 2.2 The original Stribeck curves obtained by Martens in 1888 (Martens, 18881889) .................................................................................................................................... 13
Figure 2.3 Elastic deformation of crystal lattices during dry sliding (Holinski, 2001) ....... 15
Figure 2.4 Tribological changes during initial sliding of two solid bodies (Holinski, 2001)
............................................................................................................................................. 16
Figure 2.5 Hydrodynamic lubrication .................................................................................. 18
Figure 2.6 The three lubrication regimes in the Stribeck curve ....................................... 19
Figure 2.7 Lubrication regimes in water-lubricated bearings, reproduced from Kotousov
(2009), p. 7, Figure 2.2.1 ..................................................................................................... 22
Figure 2.8 Boundary and hydrodynamic regimes of lubrication, reproduced from Kotousov
(2009), p. 7, Figure 2.2.2 ..................................................................................................... 23
Figure 2.9 Typical baseline frictionspeed curve (Pan et al., 1971) ................................... 28
Figure 2.10 Analytical two-degree model representing a submarine aft water-lubricated
bearing, as displayed in Simpson and Ibrahim (1996), p. 90, Figure 2 ............................... 30

Figure 3.1 Experimental wear rates for water-lubricated bearing materials (Biswell, 2007,
Cumberlidge, 2009, WRTSIL, 6/09/2007) .................................................................... 40
Figure 3.2 Scaled test rig and major components, as displayed in Kotousov (2009), p. 13,
Figure 3.2.1 .......................................................................................................................... 41
Figure 3.3 Friction curves vs. sliding speed, m/s (rotation speed, rpm), reproduced from
Kotousov (2009, p.18) ......................................................................................................... 43

Figure 4.1 Schematic diagram of a Pin-on-Disk experimental apparatus ........................... 50


Figure 4.2 Variable drive system: a) 3-phase BALDOR: MM3550C-57 motor with b)
GENESIS: NEMA-4X/IP-65 adjustable frequency drive ................................................... 52
Figure 4.3 Calibration data for disk rpm versus sliding speed for the POD test rig ............ 53
xi

Table of figures

Figure 4.4 Load cell with one additional weight block (applied force 17.5 N) ................... 54
Figure 4.5 Bending arm with attached strain gauges ........................................................... 57
Figure 4.6 Calibration data friction force versus voltage output for bending arm 1 ........... 58
Figure 4.7 Load cell with pin and displacement sensor ....................................................... 59
Figure 4.8 Calibration data friction force vs. voltage output for load cell .......................... 59
Figure 4.9 Schematic diagram of the data acquisition system used for the measurements of
pin displacement and arm forces on the POD test rig.......................................................... 60
Figure 4.10 Schematic diagram of the water supply system used on the POD test rig ....... 61
Figure 4.11 Design sketch of the test rig identifying the major components ...................... 62
Figure 4.12 Fully-equipped Pin-on-Disk test rig ................................................................. 63
Figure 5.1 Typical baseline of viscosity of water vs. temperature T, 0C, reproduced from
Ginzburg et al. (2006, p.696), Figure 2 ............................................................................... 71
Figure 5.2 Water-lubricated bearing damage (subjected to long-lasting operation which
resulted in significant wastage and associated ovalisation of the bush), where Do=initial
dimension, Dp=actual dimension and Dw=wear due to water contamination as displayed in
Litwin (2009, p.44, Figure 6 and p.48, Figure 21) .............................................................. 72
Figure 5.3 PTFE test sample used during the validation study ........................................... 74
Figure 5.4 NF22 (Railko) sample used in the experimental study ...................................... 77
Figure 5.5 AISI 440C stainless steel test disk fitted on the POD test rig ............................ 78

Figure 6.1 Coefficient of friction of the PTFE pin against a stainless steel disk for a sliding
speed of 0.32 m/s ................................................................................................................. 82

Figure 7.1 Coefficient of friction of NF22 (Railko) material against stainless steel versus
normal applied load under dry conditions ........................................................................... 88
Figure 7.2 Coefficient of friction of NF22 (Railko) material against stainless steel versus
sliding speed under dry conditions ...................................................................................... 89
Figure 7.3 Coefficient of friction of NF22 (Railko) material against stainless steel versus
normal load under clean water-lubricated conditions .......................................................... 90
Figure 7.4 Coefficient of friction of NF22 (Railko) material against stainless steel versus
sliding speed under clean water-lubricated conditions ........................................................ 91
Figure 7.5 Coefficient of friction of NF22 (Railko) against stainless steel for 1%
contaminated water lubrication ............................................................................................ 93
xii

Table of figures

Figure 7.6 Coefficient of friction of NF22 (Railko) against stainless steel for 2%
contaminated water lubrication ............................................................................................ 93
Figure 7.7 Coefficient of friction of NF22 (Railko) against stainless steel for 4%
contaminated water lubrication ............................................................................................ 94
Figure 7.8 Coefficient of friction of NF22 (Railko) against stainless steel for 6%
contaminated water lubrication ............................................................................................ 94
Figure 7.9 Coefficient of friction vs. water contamination of NF22 (Railko) material
(sliding speed=0.393 m/s) .................................................................................................... 95

Figure 8.1 Mass loss for NF22 (Railko) material at different load and speed values under
1% water contamination .................................................................................................... 102
Figure 8.2 Mass loss for NF22 (Railko) material at different load and speed values under
2% water contamination .................................................................................................... 102
Figure 8.3 Mass loss for NF22 (Railko) material at different load and speed values under
4% water contamination .................................................................................................... 103
Figure 8.4 Mass loss for NF22 (Railko) material at different load and speed values under
6% water contamination .................................................................................................... 103
Figure 8.5 Mass loss versus degree of water contamination for NF22 (Railko) material at a
sliding speed of 0.393 m/s ................................................................................................. 104
Figure 8.6 Micrograph of pins worn surface before/after a full cycle of experiments, at a
magnification of X500 ....................................................................................................... 105
Figure 8.7 Micrograph of pins worn surface after a full cycle of experiments, at a
magnification of X100 ....................................................................................................... 106
Figure 8.8 Micrograph of pins worn surface after a full cycle of experiments ................ 107
Figure 8.9 Micrograph of stainless steel disks worn surface after a full cycle of
experiments, at a magnification of X100 ........................................................................... 107
Figure 8.10 Specific wear rate for NF22 (Railko) material at different load and speed
values with 1% water contamination ................................................................................. 109
Figure 8.11 Specific wear rate for NF22 (Railko) material at different load and speed
values with 2% water contamination ................................................................................. 109
Figure 8.12 Specific wear rate for NF22 (Railko) material at different load and speed
values with 4% water contamination ................................................................................. 110
Figure 8.13 Specific wear rate for NF22 (Railko) material at different load and speed
values with 6% water contamination ................................................................................. 110
xiii

Table of figures

Figure 8.14 Specific wear rate versus degree of water contamination of NF22 (Railko)
material for a sliding speed of 0.393 m/s ........................................................................... 111

Figure 9.1 POD test rig equipped with a magnet ............................................................... 116
Figure 9.2 Calculated friction force (N) versus time for undamped and damped conditions
under load 8 N and sliding speed 0.393 m/s ...................................................................... 121
Figure 9.3 Calculated friction force (N) versus time for undamped and damped conditions
under load 8 N and sliding speed 1.557 m/s ...................................................................... 124
Figure 9.4 Welch power spectral densities for damped and undamped conditions under
load 8 N and 0.393 m/s, sliding speed ............................................................................... 128
Figure 9.5 Welch power spectral densities for undamped and damped conditions under
load 8 N and 1.557 m/s, sliding speed ............................................................................... 131
Figure 9.6 Calculated RMS values for NF22 (Railko) material at normal and damped load
(8 N) ................................................................................................................................... 134
Figure 9.7 Specific wear rate for NF22 (Railko) material for undamped and damped
conditions ........................................................................................................................... 137
Figure 9.8 Microscopy of the pins worn surface before and after a full cycle of
experiments at a high sliding speed of 1.557 m/s, at a magnification of X30 ................... 139

xiv

List of tables

LIST OF TABLES
Table 3.1 Basic parameters for the Wrtsil wear tests (Biswell, 2007) ............................. 39
Table 3.2 Basic parameters of the bearing system (Kotousov, 2009, Solomonov et al.,
2010) .................................................................................................................................... 42

Table 4.1 Design requirements for the Pin-on-Disk test rig ................................................ 51
Table 4.2 Design requirements for the bending arms used on the POD test rig .................. 55
Table 4.3 Technical characteristics of bending arms ........................................................... 57

Table 5.1 Physical properties of PTFE material .................................................................. 74


Table 5.2 Physical properties of NF22 (Railko) material (WRTSIL, 6/09/2007) ......... 76

Table 6.1 Technical parameters adopted for the validation study ....................................... 81

Table 7.1 Operational parameters adopted for experimental study ..................................... 86

Table 8.1 Parameters for wear test..................................................................................... 100

Table 9.1 Technical parameters adopted for the vibrationwear experiments .................. 117
Table 9.2 Calculated RMS acceleration values at different sliding speeds, lubrication, and
contamination conditions, under an 8 N load .................................................................... 133
Table 9.3 Average mass loss (g) at different sliding speeds, lubrication, and contamination
conditions, under an 8 N load ............................................................................................ 136

xv

CHAPTER 1 Introduction

CHAPTER 1

INTRODUCTION

1.1 Overview
Since prehistoric times, human activities have involved friction. It is due to friction that we
are able to stand, run, start a fire, or even swim. Yet while friction is useful and necessary
for many human activities, it can also create many technological problems.

Friction takes place in all mechanical applications. These can include bearings, braking
systems and transmissions that involve two interacting surfaces, resulting in energy loss
and wear. Mankind has therefore made every possible effort to defeat the negative effects
of friction during the whole of human history.

It is known that friction can be reduced by lubrication. Lubrication involves placing


another material such as water or oil between the two surfaces (refer to Figure 1.1) in order
to extend the life cycle of a machine or a mechanism. In the 1960s, the science and
technology concerning friction, wear, and lubrication were unified under the discipline
named tribology. (Hori, 2006).

Figure 1.1 Diagram showing a simple tribological system


1

CHAPTER 1 Introduction

The modelling of tribological behaviour is a complex and difficult task. It is extremely


important for all stages of the life cycle of a machine and its components (as shown in
Figure 1.2). During machine design, accurate tribological simulation allows material
performance to be predicted and the mechanical design of materials and lubricants to be
engineered in a such way that the life cycle of the machine elements can be optimised
(Nikolakopoulos and Papadopoulos, 2008).

Figure 1.2 Diagram showing the complicity and complexity of tribological processes
(Materials Tribology laboratory, 2008)
The problem of friction, vibration and wear in rotating machinery is significant in many
engineering applications, particularly in water-lubricated bearings where it is undesirable
to experience friction and vibration between moving parts (Hori, 2006). Therefore,
increasing interest has been expressed in rotor dynamics, particularly the stability problems
encountered in high speed rotating machinery supported by water-lubricated bearings.

Three types of lubrication regimes are considered in water-lubricated journal bearings:


boundary, mixed and hydrodynamic. In hydrodynamic lubrication regime, lubricant is
squeezed between a rotating shaft and its bearing. If the load is relatively low and the
speed of the rotating shaft is high enough, a complete film of lubricant is formed (UoS,
2008). In real engineering applications using this type of lubrication, such as in propulsion
shaft systems, hydroelectric turbines and water pumps, the frictional energy loss is due to
2

CHAPTER 1 Introduction

the viscous forces in the lubricant and any water contamination (UoS, 2008, Kotousov,
2009, Maru et al., 2005, Ibrahim, 1994a, Maru et al., 2007b, Meuter, 2006, Hori, 2006,
Nikolakopoulos and Papadopoulos, 2008). Viscosity cannot be completely eliminated,
however, because the separation of the surfaces is dependent upon it; as the viscosity
decreases so does the separation until the surfaces come into contact (UoS, 2008). When
the rotational speed of the shaft in water-lubricated bearing decreases, contact begins to
occur as a continuous film of fluid is being broken. At this stage, the situation is a mixture
of hydrodynamic and boundary lubrication, which is called, mixed lubrication. If the shaft
rotation is decreased further, the film of lubricant is reduced to localised patches which are
only a few molecules thick. This type of lubrication is known as boundary lubrication. In
boundary lubrication, the friction coefficient does not depend on the viscosity of the
lubricant but rather on its material properties (UoS, 2008).
A good boundary lubricant is one that will attach itself firmly to the clean surfaces formed
as the cold-welded junctions are sheared. A layer is then formed that acts as a lubricating
film. If that layer can be easily sheared, then the friction is low. Typically, when the
coefficient of friction is of the order of 0.1, the wear is insignificant (UoS, 2008).

Water-lubricated bearings are one of the most important and promising directions for
further development in the ship-building industry. Propeller shafts of ships and submarines
are supported in stern-tube bearings. These bearings are water-lubricated, but the rotational
speed of the shaft may not be high enough to achieve a hydrodynamic regime. In this case,
unlubricated frictional mechanisms would prevail at the region of the surface area in direct
contact. The overall friction torque would be considerably higher than in the hydrodynamic
regime (full separation of the sliding surfaces) (Simpson and Ibrahim, 1996). Therefore,
contact mechanics and friction phenomena in water-lubricated bearings are of interest to
contemporary researchers and engineers, especially in the field of tribology. During the last
two decades, many theoretical analyses and experimental studies have been undertaken to
investigate the operational characteristics and to identify the best materials for waterlubricated bearings (Harrison, 2008).

For marine engineering applications, two types of aft bearing systems exist. One is a waterlubricated bearing system that is mainly applied to small vessels. This is because of the low
allowable bearing pressure (Yamajo and Kikkawa, 2004). The other type of aft bearing
system is an oil-lubricated system for large vessel applications (Read and Flack, 1987).
3

CHAPTER 1 Introduction

This is because, at low speeds, oil-lubricated bearings provide better boundary lubrication
and therefore can withstand higher loads compared to water-lubricated bearings.

A water-lubricated journal bearing (a stern-tube bearing) is a simple bearing in which a


shaft ("journal") rotates in the bearing with sea water as the liquid lubricant. Journal
bearings are simple, compact, and inexpensive to manufacture and maintain. This type of
bearing is also characterised by lightweight, high load-carrying capacity and good damping
characteristics.

According to Ryadchenko (2003), the mechanics of contact in journal bearings can be


described as follows. When a journal bearing begins rotating at low speed, little lubricant
exists between the shaft and the bearing at the contact area. This is a condition when
lubrication is provided by boundary and mixed regimes of lubrication, and high friction is
generated. When the shaft rotation approaches sufficient speed, the lubricant wedges into
the contact point and hydrodynamic regime is attained. Various approaches have been
developed for keeping the hydrodynamic regime in a journal bearing. However, with the
continuous development of journal bearing applications in rotating machinery, the serious
wear, power loss, and friction-induced vibrations still exist and need to be addressed
(Ryadchenko, 2003).

As an aft bearing system normally utilises a water-lubricated bearing, the water-lubricated


bearing plays a key role in marine engineering (see Figure 1.3).

Figure 1.3 An example of an aft bearing system (Solomonov et al., 2010)


4

CHAPTER 1 Introduction

Aft shafts used in submarines, boats, and ships are supported by water-lubricated bearings
that are lubricated and cooled by sea water. One of the problems associated with this is that
during start-up and shutdown, the rotation speed of the shaft may not be high enough to
achieve full separation of the surfaces by water. In this case, unlubricated frictional
mechanisms prevail at that portion of the surface area in direct contact. In addition, the
overall friction torque is considerably higher than the normal running value (Simpson and
Ibrahim, 1996).

Another example of friction-related problems is the appearance of power loss, vibration


and excessive wear. This is due to two variables: the viscous forces in the lubricant; and
lubricant contamination by sand. Other researchers (Ledocq, 1973, Mokhiamer et al., 1999,
Tworzydlo et al., 1994, Qiao and Ibrahim, 1999) have theoretically and experimentally
investigated these negative influences of friction.

1.2 Scope and objectives


The problem of water contamination in water-lubricated bearings and its effect on bearings
materials have received limited attention from researchers (Younes, 1993, Mosleh et al.,
2002, Solomonov et al., 2010, Maru et al., 2007a, Maru et al., 2007b). According to
Simpson and Ibrahim (1996), those studies that have been conducted were dominated by
experimental tests of section scaled models that emulated the actual bearing dynamics.
However, the materials tribological behaviour was not considered. Simple analytical
models such as one- and two-degree-of-freedom models were put forward to predict and
simulate water-lubricated bearings (Kotousov, 2009, Tworzydlo et al., 1994, Younes,
1993, Mosleh et al., 2002). However, the current theoretical and experimental models are
too simplistic and inadequate (Ibrahim, 1994a, Hori, 2006, Kingsbury, 1997, Platt et al.,
1994, Younes, 1993, Hirani et al., 1997, Simpson and Ibrahim, 1996, Kotousov, 2009,
Tworzydlo et al., 1994, WRTSIL, 6/09/2007). Not considered within these
investigations is the role of nonlinearity, which is due to the frictionspeed curve, and the
effects of lubricant contamination and damping on water lubricated bearing materials
(Simpson and Ibrahim, 1996, Kotousov, 2009, Tworzydlo et al., 1994, Younes, 1993,
Mosleh et al., 2002).

CHAPTER 1 Introduction

With the growing interest in ship and submarine construction technologies in South
Australia, further experimental investigation and theoretical development of nonlinear
models that emulate the tribology of water-lubricated bearings and their materials are of
great interest to researchers and designers in marine engineering.

The significance of this research project is to extend the previous work of Kotousov (2009)
and to develop new methods for experimental investigations (Solomonov et al., 2010) of
tribological characteristics of water-lubricated bearings materials using a Pin-on Disk test
rig. The goal is to achieve a better understanding and modelling of tribological processes in
marine water-lubricated bearings. This enhanced understanding could lead to further
development of new materials and design solutions with a focus on efficiency and
noiselessness.

The expected outcomes from this research project are:

Identification of the existing problems in the contemporary technology of waterlubricated bearing materials and systems

Analysis of the existing types of materials, and analysis of experimental and


theoretical models of water-lubricated bearings

Further development and extension of the experimental methodology of


investigation of friction, vibration, and wear of water-lubricated bearings materials
using a Pin-on-Disk test rig

Identification and further development of methods to analyse friction, vibration,


wear, the vibration-wear relationship, and the effect of water contamination on the
tribological behaviour of water-lubricated bearing materials

Identification and further development of methods to analyse the effect of


contamination and damping on the vibration-wear relationship of water-lubricated
bearings materials

Development of practical recommendations to improve the materials design,


performance, selection, and overall modelling and performance of water-lubricated
frictional systems.

CHAPTER 1 Introduction

1.3 Thesis outline


In Chapter 1, the background, scope and objectives of this project are introduced. A brief
summary of this thesis is then presented.

Chapter 2 presents existing and relevant literature on water-lubricated bearings and


friction, vibration, and wear problems associated with them. The commonly used
theoretical analysis and experimental techniques regarding water-lubricated bearings are
summarised. The brief overview of theoretical and experimental methods to analyse and
model the water-lubricated bearings is described. Proposed Pin-on-Disk experimental
methods used to analyse friction, are also introduced. Literature review is summarised and
research gaps are then identified.

Chapter 3 is the review of the results from a prior theoretical analysis and experimental
studies on the effect of water contamination on a water-lubricated bearing system which
were undertaken by other researchers in the School of Mechanical Engineering, University
of Adelaide and at Wrtsil Pty Ltd, a UK-based supplier of water-lubricated bearing
materials. These previous studies became a starting point for the experimental
investigations undertaken herein.

Chapter 4 deals with the experimental Pin-on Disk test rig which is used for the
experimental study of friction, vibration and wear. The main elements of the test rig and
data acquisition system and applied methods are discussed.

Chapter 5 presents a review of contemporary polymer-based materials that are used for
water-lubricated bearings and then describes the process used in the selection of the
available materials for further experimental investigation.

Chapter 6 includes the results of a validation study program to verify the capability of the
newly-designed and fabricated test rig. It also includes a discussion on the results obtained
from the experimental program which was conducted to examine the effect of varying
levels of contaminated water lubrication on the friction coefficient of materials for waterlubricated bearings.

CHAPTER 1 Introduction

Chapter 7 discusses the experimental results which were obtained to analyse the effect of
contaminated water lubrication regimes on the mass loss and the specific wear rate of
materials used for water-lubricated bearings. This study investigated the sliding response
of NF22 (Railko) material against stainless steel.

Chapter 8 presents an experimental investigation of the effect of water contamination on


the

vibrationwear

relationship

for

water-lubricated

bearings.

The

contacting

characteristics of sliding bodies were considered in order to ascertain the existence of


correlations between vibration phenomena and tribological aspects.

Chapter 9 presents the conclusions and recommendations for future work on theoretical
and experimental investigation of friction phenomena in water-lubricated bearing systems.
The results from this experimental investigation are also discussed and summarised.

Appendix A presents assembly drawings detailing the experimental Pin-on-Disk test rig.

CHAPTER 2 Literature review

CHAPTER 2

BACKGROUND AND LITERATURE REVIEW

2.1 Historical overview of friction


Friction, the most fundamental phenomenon accompanying sliding movement of solid
bodies, is defined as the force resisting the relative lateral (tangential) motion of solid
surfaces, fluid layers, or material elements that are in contact. Friction can be subdivided
into several types (Ruina and Pratap, 2002):

Dry friction resists the motion of two solid bodies in contact. Dry friction can be
divided into two types: static and dynamic friction (also known as kinetic or sliding
friction).

Internal friction: is when the resisting force exists due to movement between the
components of a solid body while it is being deformed.

Fluid or Lubricated friction: is when two sliding solid bodies are separated by
lubricant.

Skin friction: is when a solid is dragged through a lubricant.

Several famous researchers have contributed to the understanding of friction in the past,
such as Guillaume Amontons, Leonardo da Vinci, Lohn Theophilus desaguliers, Leonard
Euler, and Charles-Augustin de Coulomb (Engineering-ABC.com, Courtel and Tichvinsky,
1964).

Leonardo Da Vinci (1452-1519) was one of the first scientists to investigate friction
(Courtel and Tichvinsky, 1964, Krim, 2002). He focused on many types of friction. He
found differences between rolling and sliding friction. Figure 2.1 presents sketches from
(Krim, 2002) showing Da Vinci experiments to investigate: a) the force of friction
between horizontal and inclined planes; b) the influence of the apparent contact area upon
the force of friction; c) the force of friction on a horizontal plane by means of a pulley and
d) the friction torque on a roller and half bearing (Krim, 2002).
Da Vinci was the first to proclaim two laws of friction. He claimed that the frictional
resistance was the same for two different objects of the same weight but making contacts

CHAPTER 2 Literature review

over different widths and lengths. He also found that the force needed to overcome
friction is doubled when the weight is doubled (Krim, 2002).

Figure 2.1 Friction experiments suggested by Leonardo da Vinci. (Krim, 2002)

Da Vinci stated these two original laws of friction 200 years before Sir Isaac Newton (He,
2010, Zhuravlev, 2010). Da Vinci made the observation that different materials move with
different degrees of ease. He concluded that this was a result of the roughness of the
material; smoother materials will have smaller frictions (Krim, 2002). Da Vinci did not
publish his theories, so he never received recognition for his ideas. The only evidence of
their existence is in his voluminous journals.

Guillaume Amontons (1663-1705) rediscovered the two original laws of friction that had
been discovered by Da Vinci. Amontons concluded that friction was predominately a
result of the work done to lift one surface over the roughness of the other, or from the
deforming or the wearing of the other surface (Werktuigbouw&Tribologie, 2010,

10

CHAPTER 2 Literature review

Zhuravlev, 2010). For several centuries after Amontons' work, scientists believed that
friction was due to the roughness of the surfaces (Werktuigbouw&Tribologie, 2010).
Leonard Euler (1707-1783), a famous mathematician, was also much concerned with
friction problems. With the use of classical dynamics, he expressed the value of the
coefficient of friction by parameters which could be easily measured. Two of his famous
works were published under the auspices of the Royal Berlin Academy, one called
Friction of Solid Bodies and the other called Decrease of Friction Resistance. He
pointed out that the friction force is always tangential to the sliding velocity, and he
indicated the conditions of constant-acceleration motion with friction for plane and
inclined surfaces (Courtel and Tichvinsky, 1964).

Charles Augustin de Coulomb (1736-1806) added to the second law of friction (Liang,
2005, Zhuravlev, 2010). He proclaimed that strength due to friction is proportional to
compressive force although for large bodies friction does not exactly follow this law
and, as a consequence, the friction force and the area of contact are not dependable (Liang,
2005, Zhuravlev, 2010). Coulomb published his study, and referred it to Amontons
previous work. The second law of friction is known as the "Amontons-Coulomb Law"
referring to work done by the two scientists in 1699 and 1785, respectively (Werktuigbouw
& Tribologie, 2010, Zhuravlev, 2010). The Amontons-Coulomb law of friction has been
true for many combinations of materials and their geometries but, unlike Newtons first
law, nothing fundamental can be derived from it (Werktuigbouw & Tribologie, 2010).

Philip Bowden and David Tabor (1950) provided a physical explanation for the laws of
friction (Werktuigbouw&Tribologie, 2010, Courtel and Tichvinsky, 1964). They found
that the area of actual contact is a very small percentage of the total area of contact. The
actual area of contact is formed by the asperities and depends on the applied load
(Werktuigbouw&Tribologie, 2010). As the load increases, more asperities contact occur
and the actual area of each asperitys contact increases (Kaarstad, 2009).

11

CHAPTER 2 Literature review

Overall, the above findings were formulated into three friction laws (Zhuravlev, 2010):

1.

The friction force is directly proportional to the applied load. (the First Law of
Amontons)

2.

The friction force is independent of the total contact area. (the Second Law of
Amontons)

3.

Dynamic friction is independent of the sliding speed. (Coulomb's Friction Law).

A method to reduce the negative influence of friction by placing a lubricant, such as


grease, water or oil, between contacting surfaces has been known for centuries. If the
lubricant is placed between the two interacting surfaces, it allows significant reduction of
the friction coefficient and, as a consequence, the friction force (Teidelt, 2012).

Further research on the relationship between friction and lubrication was undertaken and
published in the late 1870s by Dr Robert H. Thurston at the Stevens Institute of
Technology (USA) (Thurston, 1879, Thurston, 1894), and in 1885 by Professor Adolf
Martens (1850-1914). The research was later finalised by Professor Richard Stribeck
(1861-1950) at the Royal Prussian Technical Testing Institute in Germany (Stribeck,
1902a).
Based on the results obtained in the form of so-called Stribeck curves by Professor
Stribeck in the early 1920s, the friction regimes for sliding lubricated bodies were
identified as static friction, boundary friction, mixed friction and hydrodynamic friction.
The Stribeck curve is a tribological characteristic which is utilised to characterise the
dependence of friction force on sliding speed between two liquid lubricated bodies
(Harrison, 2008, Hersey., 1934).

12

CHAPTER 2 Literature review

Figure 2.2 The original Stribeck curves obtained by Martens in 1888 (Martens,
1888-1889)

The science of friction and lubrication was named tribology and introduced in 1966.

Further study of tribology became possible with the invention of the atomic force
microscope (AFM) in 1986 and was concentrated on the micro-scale mechanics of contact
established asperities with adhesion and deformation. This approach enables researchers to
investigate friction on the atomic scale (Palaci, 2007). Thus, researchers are able to
determine how the mechanisms of friction change at macroscopic scale under different
operational conditions.

Micro- and nano-tribology have been introduced in recent years (Palaci, 2007). Frictional
interactions in microscopically small components are crucial to the development of new
products in contemporary technology such as electronics, sciences, chemistry, and
microelectronics.

13

CHAPTER 2 Literature review

2.2 Fundamentals of friction


Most textbooks on engineering mechanics recognise two types of friction opposing the
relative motion of bodies in mechanical contact: dry or Coulomb friction and fluid or
viscous friction.

2.2.1 Dry regime of friction


Dry friction may affect machine behaviour in a way that reduces or even disables machine
operation capacity. Dry friction-induced vibrations (often referred to as stick-slip) lead to
noise problems such as bearing squeal or chatter.

Dry friction between two solid bodies in mechanical contact is caused by the dry friction
force of two components: the kinetic and static friction forces. These forces develop in a
perpendicular direction to the contact plane and oppose the relative motion of the contact
surfaces. A dry sliding friction between two surfaces can be modelled as elastic and plastic
deformation forces of microscopic asperities in contact (Olsson et al., 1997).
Consequently, it is possible to manipulate friction characteristics by employing surface
films of suitable materials between the bodies in contact. These surface films can also be
the result of contamination or oxidation of the bulk material or material displacement
(Nouira, 2008).

Holinski (2001) made a statement as to why dry friction occurs in composite materials.
According to his work, the reason for friction is lattice vibration. Lattice vibration occurs
during the sliding of solid surfaces when atoms of one surface make the atoms of the other
surface vibrate. Part of the mechanical energy which is required to move both surfaces
over each other is transformed into sound waves and heat (Holinski, 2001). The
tribological interactions of a solid surfaces exposed face with interfacing materials and
environment may result in a loss of material from the surface.

According to Holinski (2001), on the atomic scale, the crystal lattices of both solid
materials in contact are in a state of equilibrium. When a shear stress is applied to one
component, both lattices deform elastically. If the shear is further increased and instability
is reached, the atoms move to a new position of equilibrium. The crystal lattice vibrates

14

CHAPTER 2 Literature review

until all strain energy has been dissipated as seen in Figure 2.3, and all elastic energy has
been converted to heat.

Figure 2.3 Elastic deformation of crystal lattices during dry sliding (Holinski, 2001)

Holinski (2001) stated It has been found that during the initial sliding of the surfaces of
two solid bodies over each other, friction force is increased, as is frictional temperature. In
this phase, no material transfer from one surface to the other occurs. Only at the frictional
maximum are particles transferred, as shown in Figure 2.4 (Holinski, 2001).

15

CHAPTER 2 Literature review

Figure 2.4 Tribological changes during initial sliding of two solid bodies (Holinski,
2001)

When the transfer is initiated, the friction force decreases until a level of stability is
achieved. A transfer film depends on material properties and grows to a certain thickness.
For example, in a graphite-based composite, a homogeneous layer is not formed. However,
small graphite islands are built. The friction force decreases after formation of islands. It
has been found that during the sliding of solid bodies of different composite materials
against each other, a friction layer is formed on the surface of the harder material. Surface
materials properties are changed by heat and friction. This friction layer is responsible for
the tribological characteristics such as friction coefficient, vibration, and wear rate
16

CHAPTER 2 Literature review

(Holinski, 2001). Further investigations by Raman spectroscopy (named after Sir C. V.


Raman) and microprobe analysis show that the newly - formed transfer film has chemical
differences compared to the matrix of material. A surface layer is located on the harder
sliding composite material. Often, the total composition from the softer material is not
transferred but only some of the chemical components. Both layers are governing the
tribological characteristics of sliding systems. In some cases, chemical components are
added to one sliding material to form a particular friction layer to achieve the desired
tribological characteristics (Holinski, 2001).

In many cases, dry friction process leads not to transfer but to loss of material. This process
is known as "wear" (PML, 2009). Major types of wear include abrasion, adhesion
(friction), erosion, and corrosion. Wear can be minimised by modifying the material
properties or the surface properties of solids by surface engineering processes (also
called surface finishing) or by the use of lubricants (for frictional or adhesive wear) (PML,
2009).

According to Holinski (2001), depending on operational parameters such as pressure,


sliding speed and type of material, different frictional temperatures occur which give rise
to various power losses. Dry friction always generates heat and consequently diffusion
processes. Thus, dry sliding friction creates surface deformation which results in crystal
lattice deformation and higher concentration of dislocations. Also, chemical reactions such
as tribological oxidation and decompositions occur. All these factors finally lead to cracks,
fractures and wear.

2.2.2 Hydrodynamic regime


Rac and Vencl (2005) stated Complete separation of sliding surfaces with lubricant can be
achieved by hydrodynamic lubrication. The selection of the design and the tribological
parameters in the region of the hydrodynamic lubrication should ensure adequate thickness
of the lubricant film and temperature of the bearing (Rac and Vencl, 2005).
According to MarineDiesels (2011), Hydrodynamic lubrication was first researched by
Osborne Reynolds (1842-1912) in an experimental test rig for modelling a liquidlubricated bearing. When a lubricant was applied between the shaft and bearing, Reynolds
found that the rotating shaft pulled a converging wedge of lubricant between the shaft and
17

CHAPTER 2 Literature review

the bearing. He also noted that as the shaft gained velocity, the liquid flowed between the
two surfaces at a greater rate. The viscous lubricant produced a liquid pressure in the
lubricant wedge that was sufficient to keep the two surfaces separated. Under ideal
conditions, Reynolds showed that this liquid pressure was great enough to keep the two
bodies from having any contact and that the only friction in the system was the viscous
resistance or viscosity of the lubricant.(MarineDiesels, 2011)

Figure 2.5 Hydrodynamic lubrication

Viscosity of the lubricant is an important parameter: the higher the viscosity of the
lubricant, the higher the friction between lubricant and shaft but the thicker the
hydrodynamic film. However, friction generates heat. Heat will reduce the viscosity and,
therefore, the thickness of the film which may result in shaft-bearing contact. Reduced film
thickness occurs when a lubricant with low initial viscosity is used (MarineDiesels, 2011).

Care needs to be taken to ensure that the distance between the two surfaces is greater than
the largest surface defect. The distance between the two hydrodynamicaly sliding surfaces
decreases with higher loads on the bearing, lower speeds, and less viscous fluids. A
hydrodynamic regime is an excellent method of lubrication. This is because it is possible to
achieve coefficients of friction as low as 0.001 with no wear between the moving parts.
However, because the lubricant is heated by the frictional force and since viscosity is
temperature-dependent, special additives can be used to decrease the viscosity temperature
dependence (MarineDiesels, 2011).

18

CHAPTER 2 Literature review

2.2.3 Boundary and mixed regimes


Friction with boundary lubrication modes or in mixed (transition) operating conditions can
prevail in several practical cases.

Contact between the surfaces at a few high surface points (micro asperities) occurs during
boundary lubrication and then mixed lubrication as shown in Figure 2.6.

Mixed lubrication is the intermediate regime between boundary lubrication (where friction
is mostly due to asperity contact), and hydrodynamic lubrication (where the hydrodynamic
separation is achieved) (Qiang, 2009).

Figure 2.6 The three lubrication regimes in the Stribeck curve


For some materials, when a mixed regime occurs, the interacting surfaces contact at their
asperities developing heat from the localised loads introducing a co-called stick-slip
motion where some asperities can be broken off. When the high loads and temperature
occur, chemically active lubricant components interact with the contact surface. They form
a highly-resistant film, on the moving solid surfaces. This layer is capable of withstanding
the high loading pressure, and excessive wear or damage are eliminated (Mokhtar et al.,
1984, Nikolakopoulos and Papadopoulos, 2008).
19

CHAPTER 2 Literature review

Friction-induced vibrations in boundary and mixed regimes are of particular concern to the
designers of mechanical applications involving sliding surfaces at comparatively low
speeds (up to 5 m/s) (Simpson and Ibrahim, 1996, Tworzydlo et al., 1994). These
applications include liquid-lubricated bearings, wheel-rail systems, disk brake systems, and
machine tool-work piece systems. Numerical analyses to predict vibrations and noise such
as chatter and squeal in a mixed regime have not been developed, and no appropriate
theory exists that can be generalised to analyse these vibrations. Most research activities
are based on experimental tests of physical models. In general, friction-induced vibrations
on boundary and mixed regimes vary and depend on temperature, normal load, speed, and
other conditions (Simpson and Ibrahim, 1996).

2.3 Friction in water-lubricated bearings


The plain bearing was initially invented to reduce friction by placing animal fat between a
sliding shaft and a bore. Plain bearings are the simplest and the least expensive type of
bearings. They are also, lightweight, compact and are capable of withstanding high loads.
Historically, the plain bearings originated approximately 4,500 years ago from Phoenician
seafarers. Their plain bearings were built from lignum vitae, a type of wood with an
appropriate combination of density, strength, and toughness (Weichsel, 1994).

According to Weichsel (1994), modern plain bearings can be divided into three basic
categories based on the lubricant system required for successful operation:

Bearings that require oil, grease or some other lubricant to operate. They receive
this liquid or semi-solid lubricant from an outside source.

Bearings that contain the necessary lubricant within their walls: i.e., a plastic
bearing such as polyacetal, which utilises a silicone lubricant.

Bearings that are in and of themselves the lubricant, such as metallised carbon
graphite or have a running surface that contains Teflon.

Contemporary aft water-lubricated bearings are plain bearings that require lubrication by
sea water. The problems of power loss, vibration, and wear associated with water-

20

CHAPTER 2 Literature review

lubricated bearings are of great interest to researchers (Rac and Vencl, 2005, Hori, 2006,
Kingsbury, 1997).

2.3.1 Lubrication regimes


The theory on which water-lubricated bearings designs and calculations is based is
complex. It presupposes that the bearings operating conditions such as the load, size,
clearance, and properties of the materials and lubricant are known. Various
recommendations concerning the selection of those values, such as those proposed by Rac
and Vencl (2005), appear in the literature. For example, for a known load and sliding
speed, the preliminary sizes of the hydrodynamic sliding bearings can be selected.
Alternatively, for a selected size of bearing, the load capacity for a given speed can be
determined from the recommendations. These values are related to the steady-load
condition. If the load varies in magnitude and/or direction, the procedure for the selection
of bearing parameters is more complex and requires additional design considerations (Rac
and Vencl, 2005).
Friction in this type of water-lubricated bearing can be described by the Stribeck model
(Stribeck, 1902b). According to this model, four different frictionspeed regions exist (see
Figure 2.7):

1.

Static friction, where two static surfaces are mostly in contact with each other.

2.

Boundary lubrication, where the fluid films are negligible and sufficient asperity
contact exists.

3.

Partial fluid lubrication (transition or mixed) where two surfaces are partly
separated, partly in contact and the relative motion of the surfaces is insufficient to
generate the hydrodynamic action required to separate them completely (Ivanov
and Ivanov, 2012, Barwell, 1984).

4.

Hydrodynamic fluid lubrication, which is sufficiently developed to completely


separate the interacting surfaces (Barwell, 1984).

21

CHAPTER 2 Literature review

Friction coefficient

Region 2
Region 1
Region 3

Region 4
Velocity

Figure 2.7 Lubrication regimes in water-lubricated bearings, reproduced from


Kotousov (2009), p. 7, Figure 2.2.1

When a water-lubricated bearing starts up (Region 2), the combination of a low sliding
speed, low viscosity, and high load will produce boundary lubrication. Boundary
lubrication is characterised by the small amount of fluid in the interface and the large
surface contact. This results in extremely high friction (STLE, 2008, Solomonov et al.,
2010).

As the sliding speed and fluid viscosity increase, or the load decreases, the surfaces begin
to separate (Region 3), and a fluid film begins to form. The film is still insufficient but acts
to support more of the load. Mixed lubrication is the result of a fast drop in the friction
coefficient (STLE, 2008). This results in less surface contact and more fluid separation
(Neveu et al., 2012). Any of the following can prevent the build-up of a film thick enough
for hydrodynamic lubrication: insufficient surface area, a drop in the sliding speed of the
moving surfaces, reduced quantity of lubricant delivered to a bearing, an increase in the
bearing load; an increase in lubricant temperature resulting in a decrease in viscosity, or
contamination of the lubricant. If the film thickness is still insufficient, the highest
asperities may not be fully separated by lubricant (STLE, 2008, Solomonov et al., 2010).

The surfaces will continue to separate as the sliding speed or viscosity increase until full
fluid film separation has been achieved and no surface contact exists (Region 4). The
friction coefficient will reach its minimum and a transition to hydrodynamic lubrication
occurs. At this point, the load on the interface is entirely supported by the fluid film. There

22

CHAPTER 2 Literature review

is low friction and no wear in hydrodynamic lubrication since there is full fluid film
separation and no contact between asperities (STLE, 2008, Solomonov et al., 2010).

With an increase in the sliding speed, the friction increases further in the hydrodynamic
region due to fluid drag (friction produced by the fluid): higher speed may result in a
thicker fluid film, but it also increases the fluid drag on the moving surfaces. Thus, a
higher viscosity will increase the fluid film thickness but it will also increase the drag
(STLE, 2008, Solomonov et al., 2010).

Water-lubricated bearings experience boundary lubrication and mixed lubrication at startup and shutdown (low speeds and thin film) before the transition to hydrodynamic
lubrication at normal operating conditions (high speeds and thick film) (STLE, 2008). For
example, during start-up, as the sliding speed between the shaft and bearing increases, the
change from boundary lubrication to mixed and then to the hydrodynamic regime is not a
sudden or abrupt process (Dulias, 2002). A boundary-type and then mixed lubrication
occur first, and then as the surfaces slide faster, the hydrodynamic-type lubrication
becomes predominant with the change of the position of the journal shaft as shown in
Figure 2.8 (Kotousov, 2009, STLE, 2008, Solomonov et al., 2010).

R
P

o
hmin

Hydrodynamic Regime
(Region 4)

Boundary Regime
(Region 2)

Figure 2.8 Boundary and hydrodynamic regimes of lubrication, reproduced from


Kotousov (2009), p. 7, Figure 2.2.2

23

CHAPTER 2 Literature review

2.3.2 Friction-induced vibrations


When the friction coefficient depends on the sliding speed and has a negative slope with
respect to the velocity (as shown in Figure 2.7), the friction gives rise to different types of
friction-induced vibrations (EOV, 2004). These vibrations can include stick-slip
(Schallamach, 1963, Schallamach, 1971, Younes, 1993, Bhushan, 1980), quasi-harmonic
oscillation (Kotousov, 2009, Ibrahim, 1994a, Ibrahim, 1994b, Qiao and Ibrahim, 1999) and
surface-induced vibration (Platt et al., 1994, Ibrahim, 1994a, Ibrahim, 1994b).

A comprehensive literature review of friction-induced vibrations and related problems in


water-lubricated journal bearings was recently documented by Kotousov (2009). Several
distinct mechanisms can contribute to these different types of oscillations:

Self-excited vibrations of the stick-slip type occur at very low sliding speeds
(typically 0-0.3 m/s) that can be associated with start-up or shutdown. They are
attributed to the difference between the static and kinematic coefficient of friction
(Dautzenberg, 1986). At low sliding speeds, friction between the interacting
surfaces is a result of the bearing surface characteristics and properties of the
lubricant traces other than viscosity, such as metalliquid adhesion energy. The
adhesion energy is expressed as surface wettability, the actual process in which a
liquid spreads on a solid substrate or material (Younes, 1993).

Quasi-harmonic oscillations are believed to be connected with specific


characteristics of the friction versus sliding speed curve, in particular, with the
negative slope of this curve, which can spread to the sliding speed of 2-3 m/s
depending on the materials, surface, and lubrication conditions. This type of
vibration is significantly affected by surface roughness; however, it can also occur
under any surface roughness conditions including perfectly smooth surfaces
(Kotousov 2009).

Roughness-induced vibrations are associated with surface roughness and asperities


on the contact surfaces (Kotousov 2009).

Many attempts have been made to investigate and analyse friction-induced vibrations in
water-lubricated bearings. For example, Simpson and Ibrahim (1996) conducted a series of
experiments to examine the mechanism for generating vibrations in water-lubricated
bearing systems. Their study suggested that the presence of water-lubricated bearing
24

CHAPTER 2 Literature review

vibration is associated with both the contact mechanics of the bearing and the dynamic
characteristics of the structural components of the bearing system (Simpson and Ibrahim,
1996).

Simpson and Ibrahim (1996) considered several mechanisms which can give rise to
friction-induced vibrations in water-lubricated bearings. They stated that, when one of the
sliding surfaces is characterised by a certain degree of elastic freedom, the motion may not
be continuous, but may be intermittent and proceed as a stick-slip process. During stickslip motion, two different deformation mechanisms take place. The first is elastic
deformation, where the two contact surfaces stick and the asperities deform elastically. The
second is plastic deformation, where sliding takes place and the asperities deform
plastically (Simpson and Ibrahim, 1996). The authors concluded that the occurrence of
stick-slip is unpredictable, mainly because the slope of the frictionspeed curve is not
constant but varies randomly with contamination, surface finish, misalignment of sliding
surfaces and other factors (Simpson and Ibrahim, 1996). They experimentally simulated
bearing dynamics and derived a linear analytical model for water-lubricated bearings.
Different dynamic characteristics were identified from the numerical simulation of the
equations of motion. However, the role of nonlinearity, due to the frictionspeed curve,
was not considered (Simpson and Ibrahim, 1996).

Aronov et al. (1983) presented some experimental results of an investigation of the


interaction between friction, vibration, wear, and system rigidity. The results were obtained
from experimental investigation of a metal pin sliding on a steel disk under the action of
clean water as a lubricant. In this investigation, the load normal to the surface of the disk
was varied and the sliding speed was kept constant at 0.73 m/s. They found that severe
friction and wear are independent of system rigidity but dependent on the normal load.
However, it was also shown that mild wear rate increases with normal load and also with
system rigidity (Aronov et al., 1983).

2.4 Numerical models of friction in water-lubricated bearings


Frictional forces between interacting surfaces exist due to different and complex
mechanisms. Frictional forces can be responsible for undesirable dynamic characteristics.
Unfortunately, there is no appropriate method for determining or measuring friction,
vibration, and wear between sliding bodies under water lubrication. The modelling of the
25

CHAPTER 2 Literature review

friction in mechanical systems depends on several factors that were classified by Ibrahim
(1994). These factors include the material properties, characteristics of the sliding surfaces
geometry, surface roughness and structure, normal load, sliding velocity, and temperature.

Usually, a water-lubricated bearing is designed to satisfy technical requirements such as


the power and/or the flow rate to be delivered, the torque required, and other performance
factors which are independent of friction, vibration and wear (Read and Flack, 1987).
During the design process, the tribological characteristics of the bearing system are either
considered last or not considered at all. The simplest form of analytical design is based on
short- and long- bearing approximations that are inaccurate in most practical design ranges
(Hirani et al., 1997). If any problems appear, either in the design stages or testing stages,
redesigning the entire bearing system is complex, highly difficult, and sometimes
impossible. The designers and users of water-lubricated bearings require reliable tools for
assessment and adjustment for the tribological characteristics of different bearing systems
(Read and Flack, 1987). Thus, numerical modelling of friction in water-lubricated bearings
would be a useful tool.

The numerical modelling and evaluation of the coefficient of friction, vibration, and wear
as functions of sliding speed have been the subject of numerous studies which considered
the influence of such factors as material properties, surface roughness, load, and type of
lubrication (Younes, 1993). Younes undertook investigation to determine the dynamic
behaviour of a rigid balanced rotor with its journals rotating semi-dry in their bearings.
Their equation of motion, with the exponential coefficient of friction model, has been
solved numerically to determine the vibration responses. Conditions for stability, the
increase of the system vibrations, and the planetary motion of the journal along the bearing
surface were reported for different design parameters (Younes, 1993).

A fast analytical method for evaluating design parameters such as load capacity, maximum
pressure, flow rate, power loss, and maximum temperature in the lubricant film of liquidlubricated bearings was developed by Hirani et al. (1997). According to Hirani (1997),
these parameters are either too involved because of the mathematical complexities of
hydrodynamics or are based on simple methods and design charts. To predict the
tribological characteristics of bearings, an analytical model was proposed which provides
results comparable with time-consuming techniques such as thermo-hydrodynamic,
adiabatic, and isothermal analysis. Finally, the design methodology was arranged in
26

CHAPTER 2 Literature review

analytical expressions and tabular form for adequate and easy prediction of load capacity,
loss of power, flow rate, and maximum temperature. These approaches allowed to avoid
majority of numerical and mathematical complexities (Hirani et al., 1997). However, due
to the continuous development of rotating machinery, this approach is rarely adequate for
conditions involving contaminated lubricants because of the complexity of the friction
mechanism. Further development is required to enable adequate predictions and to make it
easy to use for all practical applications.

2.4.1 Linear models


Existing linear analytical models of water-lubricated bearings were reviewed by Simpson
and Ibrahim (1996), page 90. They stated that different dynamic characteristics are
predicted from the numerical simulation of the equations of motion. Several mechanisms
can give rise to friction-induced vibration in sliding systems. For example, in one-degreeof-freedom systems, where contacting surfaces interact against each other in the one
direction, a necessary condition that contributes to friction-induced vibration is the
negative slope of the frictionspeed curve (as shown in Figure 2.6) (Simpson and Ibrahim,
1996).

As was shown by Kotousov (2009) and Tworzydlo et al. (1994), in water-lubricated


bearings, the total friction at contact consists of two components: one is due to the asperity
contact friction, Fas, and the second is due to hydrodynamic viscous shear friction, Fv. The
relative importance of the two parts depends on the relative magnitude of the film
thickness between sliding surfaces and the surface roughness. According to Kotousov
(2009, p.8), the friction due to asperity contact is calculated as follows:

Fas f as

Aa
P
At

(2.1)

where:
fas is the asperity sliding friction coefficient;
Aa is the actual area of contact;
At is the total area of contact when the film thickness is zero;
P is the bearing pressure.

The viscous shear friction force is:


27

CHAPTER 2 Literature review

Fv o

2A t A a
V
2c
h min
2A t

(2.2)

where:

o is the viscosity of the lubricant;


h min is the minimum film thickness;

2c is the contact dimension;


V is the sliding speed.
Dividing by the bearing pressure P and summing the Equations 2.1 and 2.2, the total
friction coefficient is given by:

f f as

Aa
2A t A a
V
2c
+ o
.
At
h min P
2A t

(2.3)

According to Pan et al. (1971), viscosity and speed always appear as a product in the
elastohydrodynamic theory of lubrication, and the lift-off sliding speed is roughly
proportional to the viscosity. The speed, at which full film separation between shaft and
bearing takes place, decreases with decreasing temperature.

Based on the elastohydrodynamic theory of lubrication, Pan et al. (1971) established


representative baseline calculations, showing the dependence of the friction coefficient on
the shaft sliding speed as shown in Figure 2.9. The friction coefficient increases with the
effective surface roughness (or asperity size).
Friction coefficient
1

10-1

Negative slope

10-2
Positive slope
10-3
0

Sliding speed, m/s

Figure 2.9 Typical baseline frictionspeed curve (Pan et al., 1971)


28

CHAPTER 2 Literature review

According to Figure 2.9, the boundary and mixed regimes of lubrication for typical waterlubricated bearings are within the sliding speed range up to ~2 m/s (which is in agreement
with the Stribeck model) (Pan et al., 1971).

2.4.2 Non-linear models


Friction forces between sliding surfaces arise due to a complex mechanism, and lead to
numerical models which are strongly non-linear, discontinuous and non-smooth (Qiao and
Ibrahim, 1999). The inclusion of non-linearity in the equations of motion of a dynamic
system leads to differential inclusions in the mathematical model, adding further difficulty
to the problem (Qiao and Ibrahim, 1999, EOV, 2004). Many researchers have conducted
numerical studies of friction, vibration, and wear in water-lubricated bearings (Simpson and
Ibrahim, 1996, Qiao and Ibrahim, 1999, Mottershead et al., 1997). Their general

conclusions are:

1.

Most of the friction-induced vibrations takes place at low sliding speeds (typically
up to 2.0 m/s) where the slope of the friction coefficient with respect to the sliding
speed is negative (boundary-mixed regimes).

2.

The variation of the friction force, vibrations, and wear decrease as the speed of the
shaft increases to a critical speed above which the friction coefficient begins to
increase due to viscous shear (hydrodynamic regime).

The system model of a stern-tube bearing was considered by Simpson at al. (1996) who
developed an analytical nonlinear two-degree-of-freedom model for the water-lubricated
journal bearing, as shown in Figure 2.10. The mass, stiffness, and damping coefficient of
the tangential shear of the water-lubricated bearing are indicated by m1 , k 1 and c1
respectively. The constants ( k 2 and c 2 ) represent the stiffness and damping of the flexible
shaft which drives the disk of mass ( m 2 ) and of the mass moment of inertia ( I ) about the
shaft axis.

29

CHAPTER 2 Literature review

Bearing

Disk

k1
c1

k2
m1

m2, I

V
c2

Friction force

Figure 2.10 Analytical two-degree model representing a submarine aft waterlubricated bearing, as displayed in Simpson and Ibrahim (1996), p. 90, Figure 2

The non-linear response of this system which emulated the dynamics of a submarine
water-lubricated bearing shaft was investigated numerically. The influence of a time
variation of the sliding speed results in a time variation of the friction force. This time
variation of the friction force is found to be responsible for the occurrence of vibration and
noise (Simpson and Ibrahim, 1996).

2.5 Vibrationwear relationship in dynamic systems


In spite of the need to correct prediction and design of the different dynamic systems
(including water-lubricated bearings) and the increased interest of designers and engineers,
the problem of the vibrationwear relationship in different dynamic systems has not
received adequate attention from researchers. Vibration in different dynamic systems
(machines, bearings) is generally due to the dynamic friction forces as discussed by
Krishna Kumar et al. (1997), and the vibrationwear relationship depends on many factors.
These factors include contact materials; operational parameters (load, sliding speed, type
of lubrication and lubricant contamination); and the characteristics of the dynamic system
(natural frequencies and inertia of fixtures and components) (Krishna Kumar and
Swarnamani, 1997).

While the vibrationwear relationship in most dynamic systems is often unknown, the
assessment of wear of the contacting surfaces of many dynamic systems, such as bearings
and sliders, by vibration monitoring using various shock pulse measurements, methods and
tools, has been employed by many researchers and industries (Jonson, 2000).

30

CHAPTER 2 Literature review

A reliable and inexpensive methodology is required to simulate and assess vibrationwear


dependency for a physical water-lubricated bearing.

2.5.1 Problem of vibrationwear dependency


The influence of friction-induced vibrations on the wear that occurs in many mechanical
systems (brakes, bearings, wheel-rail contact, etc.) can lead to higher wear and the
formation of unwanted features on the contacting surfaces (corrugations, micro-cracks,
wash boarding) as well as noise and more vibration (Popov, 2010).

According to Chowdhury (2009), it was observed by several authors (Kato et al., 1982,
Maru et al., 2005, Maru et al., 2007b, Bryant and York, 2000) that the wear reduction
depends on interfacial conditions such as normal load, geometry, relative surface motion,
sliding speed, surface roughness of the contact surfaces, type of interacting materials,
system rigidity, temperature, humidity, type of lubrication, vibration and lubricant
contamination (Chowdhury and Helali, 2009).

Several studies have been conducted in which researchers have found that wear can be
reduced by vibration. Goto and Ashida (1984) found that ultrasonic vibration can
significantly reduce wear rates on Pin-on Disk (POD) experimental apparatus which uses a
steel pair of interacting materials. Bryant and York (2000) showed that micro-vibrations
(10-100 mm amplitude, 10-100 Hz) of a slider can reduce sliding wear by up to 50%,
particularly for rigid body rocking vibration. Moriwaki and Shamoto (1991) used
ultrasonic vibrations to reduce tool wear without damaging the surface finish. Kato et al.
(1982) also found that vibrations sometimes increased and sometimes decreased wear
rates, depending on the pairs of materials involved. In their experimental investigation,
Weber et al. (1984) used ultrasonic vibrations to extend carbide tool life in the machining
of glass (Chowdhury and Helali, 2009).

Chowdhury and Helali (2007) also considered the lack of correlation between wear rate
and other vibration-related operating parameters during an investigation of the wear
behaviour of mild steel under vertical vibration. Initially, the aim of their research was to
find a suitable correlation and a way of reducing wear rate, by applying a known,
controlled frequency and amplitude of vibration in a particular direction (Chowdhury and
Helali, 2009). A POD experimental test rig was used and results showed that the wear rate
31

CHAPTER 2 Literature review

at a particular amplitude decreases with increasing frequency of vibration. The authors also
found a reduction of the friction coefficient as a function of different amplitude and
frequency of vibration on a pair of materials comprised of mild steel. However, the wear
behaviour of mild steel and their dimensional analysis in relation to both frequency and
amplitude under vertical vibration required further investigation. It is expected that the
application of these results will contribute to the improvement of the performance of
different sliding mechanical systems (Chowdhury and Helali, 2009).

2.5.2 Effect of damping on vibration and wear


During the design stage of rotating machines, wear and vibration in bearings are significant
problems for designers. Several investigators have been interested in the damping
phenomenon. In previous studies on wear initiation and development, theoretical analyses
predicted that contact damping is the most effective factor for preventing wear (Suda and
Komine, 1996). It is known that increased damping reduces contact vibration, resulting
from surface roughness, although the mechanism may be different (Suda and Komine,
1996).
A few methods exist for contact damping to improve bearing performance. These methods
include (Sutherland, 2002):

Structural damping (which refers to energy dissipation within the structure by addon damping devices such as an isolator, by structural joints and supports, or by
structural member's internal damping);

Self-balancing damping (when the vibrations of rotors can be reduced or eliminated


by adding self-balancing mechanical components);

Liquid damping (when the vibrations of rotors can be reduced or eliminated by


adding self-balancing liquid);

Speed control damping (when the vibrations of rotors can be reduced or eliminated
by a change of rotation speed);

Magnetic damping (when the vibrations of rotors can be reduced or eliminated by a


magnetic field).

Structural damping is not always sufficient to limit vibration and wear to within the desired
level. In these cases, magnetic damping may provide a solution.

32

CHAPTER 2 Literature review

Padmanabhan (1992) emphasised the importance of damping at joints and experimentally


investigated the effect of damping on different tribological characteristics (such as friction
and wear) of interacting surfaces. A mathematical equation was formulated to predict the
effect of damping on lubricated and non-lubricated joints.

Verichev at al. (2010) outlined the principles for damping lateral vibrations of rotary
systems and undertook an experimental study to investigate the effect of speed control
damping on lateral vibrations in rotating machinery using motor speed modulation. This
method was based on the generation of a harmonic additive to the constant speed of
rotation that provided significant damping of lateral vibrations at critical rotation speeds.
The analytical solution and numerical calculations proved this concept and showed a
significant decrease in the amplitudes of lateral vibrations compared to those in a similar
undamped system (Verichev, 2010).

Kasadra et al. (2004) considered rotor instabilities in rotating machinery due to reexcitation of the first rotors critical speed (when a rotors radian frequency is in a state of
resonance with a first natural frequency) resulting in lateral rotor vibrations at frequencies
below the rotors operating frequencies. They found that active magnetic dampers were
very promising for reducing or even eliminating rotor vibrations regardless of the
excitation source. Experimentally, it was shown that an active magnetic damper was used
to effectively add damping to reduce a rotors vibration response. The experimental results
also demonstrated that the active magnetic damper can significantly dissipate the
vibrational energy and prevent system dynamic characteristics such as natural frequency,
which may lead to an increase in rotor vibrations (Kasarda, 2004).

The analysis of previous experimental investigations showed that magnetic damping is a


promising technology for reducing vibrations in water-lubricated bearings. However, it
may or may not be the best choice because the effect of damping on the vibrationwear
relationship is relatively untested in accepting the overall tribological characteristics and
needs to be addressed to ensure the effective use of active magnetic damping.

33

CHAPTER 2 Literature review

2.6 Concepts of experimental apparatus and experimental


technique
The problem of predicting the performance of water-lubricated bearings has led to a need
for further development of a reliable and inexpensive experimental methodology to
simulate a real water-lubricated bearing in an actual operational environment
(WRTSIL, 6/09/2007).

2.6.1 Pin-on-Disk system


In order to conduct an experimental investigation of friction, vibration, and wear in waterlubricated bearings, many experimental systems, including a Pin-on-Disk (POD)
experimental system, have been considered. The POD model has been successful for
similar experimental studies (Dweib and D'Souza, 1990, Aronov et al., 1983, Tworzydlo et
al., 1994, Ibrahim, 1994a, Ibrahim, 1994b, Mosleh et al., 2002, Qiao and Ibrahim, 1999,
Tworzydlo et al., 1999).

Experimental tests using a POD-type sliding system have indicated that the friction force,
vibration, and wear depend mostly on the normal load for a constant sliding speed (Qiao
and Ibrahim, 1999). According to Qiao (1999), depending on the value of the normal load,
four different friction regimes were observed:

1.

The steady-state friction regime where the frictional force increases linearly with
the normal load.

2.

The non-linear friction regime where the friction force increases non-linearly with
the normal load and the coefficient of friction is no longer constant but increases
with the normal load.

3.

The transient regime, characterised by intermittent variation of the friction force.


When the friction force reaches a sufficiently high value, a temporary burst of selfexcited vibrations occur and the friction force falls to a lower value.

4.

The self-excited vibration regime where the friction force drops to a low value and
is accompanied by high-amplitude periodic self-excited oscillations.

34

CHAPTER 2 Literature review

Chowdhary and Helali (2009) experimentally investigated a vertical vibrationwear


relationship using a POD test rig. The pair of materials used was mild steel. They found
that the wear rate for mild steel decreased with an increase in the amplitude and frequency
of the vertical vibrations. Therefore, maintaining a certain frequency and amplitude of
vertical vibration may keep wear to some lower value, which would improve the
tribological characteristics of different mechanical systems. Moreover, it was considered
that a lack of knowledge and available information on the vibrationwear relationship in
different mechanical systems existed. A lack of a suitable correlation method for reducing
wear rate by applying, or damping, known frequencies and amplitudes of vibration in a
particular direction also existed (Chowdhury and Helali, 2009).

2.6.2 Effect of lubricant contamination


Another problem which must be addressed during bearing design and operation is lubricant
contamination (Solomonov et al., 2010).

According to Maru et al. (2007a), in water-lubricated bearings, contamination of the


lubricant by wear debris or solid particles is one of the main reasons for early bearing
failure. In order to deal with this problem, it is important not only to use reliable techniques
for detection and removal of solid contaminants, but also to investigate the effects of
certain contaminant characteristics on bearing performance (Maru et al., 2007a).

Maru et al. (2007a and 2007b) presented the results of an investigation into the effect of
lubricant contamination by solid particles on the dynamic behaviour of bearings to
determine the trends in the amount of vibration induced by contamination of the oil and by
the bearing wear itself. Experimental tests were performed with radial ball bearings
lubricated in an oil bath. Quartz powder at three concentration levels and in different
particle sizes was used to contaminate the oil. Vibration signals were analysed in terms of
the root mean square (RMS) values. The results showed that changes in the RMS values of
vibration in the high-frequency band, from 600 to 10,000 Hz, were associated with changes
in the oil lubrication of the bearing as a result of contamination and wear damage to the
bearing surfaces. It was shown that the effect of contaminant concentration on vibration
was distinct from that of particle size. The vibration level increased with the percentage of
concentration, tending to stabilise at a particular limit. On the other hand, as the particle
size increased, the vibration level first increased and then decreased. Vibration level
35

CHAPTER 2 Literature review

increased during the test in contaminated oil only after 16 minutes of testing. Such an
increase in vibration was related to an effect produced by the wear of the bearing elements.
The bearing surfaces were reported to be severely damaged by a three-body abrasive
mechanism distributed along all the surfaces. Abrasion was also identified through
ferrography, a microscopic technique for analysing the particles present in fluids that
indicate mechanical wear. This indicated a severe wear regime, although measurements of
internal radial clearance of the bearings are reported to have shown an absence of
dimensional wear. The amount of vibration due to bearing wear was dependent on the
contamination level. A correlation was observed between the trends of the wear due to
bearing vibration and those of its overall surface damage. The vibration due to the presence
of particles was shown to be proportional to the vibration of the worn bearing as particle
concentration increased. On the other hand, when the contaminant particle size increased,
the dynamic action of the particles passing through the contact interface increased, but the
vibration level of the worn bearing was the same. The vibration due to the larger particles
is reported to be reduced due to the particle settling phenomenon (the decrease in vibration
with larger particles suggested that it was more difficult for larger particles to go into the
contact interface and, therefore, the vibration of the respective worn bearing was reduced)
(Maru et al., 2007a, Maru et al., 2005, Maru et al., 2007b).

2.7 Summary and research gaps


This literature review has presented that theoretical and experimental studies of friction,
vibration, wear, vibration-wear relationship, and bearings lubrication, which have been of
interest to researchers for various mechanical applications. However, existing theoretical
and experimental tribological models, presented in the literature, are mostly linear, simple
in nature, and do not consider damping, and contamination of the lubricant. Selection of
water-lubricated bearings materials generally depends on the designers previous
experience and is simplified. The relationship between various water-lubricated bearing
parameters, lubrication regimes (hydrodynamic, mixed film or boundary), lubricant
contamination in addition to the influence of water contamination and damping on friction,
friction-induced vibration, wear and the vibrationwear relationship on water-lubricated
bearings materials performance is very limited and requires further investigation.

36

CHAPTER 2 Literature review

Severe contamination in water-lubricated bearings can cause excessive wear of


components, surface damage, fatigue failure, and noise. The effect of water contamination
on friction, vibration, wear and the vibrationwear relationship has not previously been
considered. It needs to be further investigated in order to predict the tribological behaviour
of water-lubricated bearings materials during operation. It remains a matter of serious
concern that no research-based methods and models have been designed for its evaluation.

A Pin-on-Disk (POD) experimental apparatus has been chosen to investigate the effect of
water contamination and damping on friction, vibration, wear behaviour and the vibration
wear relationship in boundary and mixed regimes for water-lubricated bearings materials.
It is essential to qualify the effect of water contamination on the water-lubricated bearing
materials performance for developing design guidelines. At present, there are not design
guidelines available for water-lubricated bearing materials selection which experience
water contamination. It is therefore expected that this project will contribute to the
development of industry guidelines, and provide a better understanding and modelling
approach for the tribological process between materials used for water-lubricated systems.

37

CHAPTER 3 Previous experimental studies of tribological characteristics in water-lubricated bearings

CHAPTER 3 PREVIOUS EXPERIMENTAL STUDIES OF


TRIBOLOGICAL CHARACTERISTICS IN WATERLUBRICATED BEARINGS
3.1 Introduction
This chapter represents the experimental results and analysis of a wear study which was
conducted by Wrtsil Pty Ltd, a UK-based supplier of water-lubricated bearing materials
(Biswell, 2007, Cumberlidge, 2009, WRTSIL, 6/09/2007). It is followed by an aft
bearing study on the effect of water contamination on a water-lubricated bearing system
that was previously undertaken by researchers in the School of Mechanical Engineering,
University of Adelaide (Kotousov, 2009). These previous studies became a starting point
for the experimental investigations undertaken and results presented in the following
chapters of this thesis.

The aft bearing of ships and submarines is most commonly a plain journal bearing which is
lubricated and cooled with sea water. The relationship between operational parameters and
the system, the bearing lubrication regimes (hydrodynamic, mixed [transition] or
boundary) and the dynamic motions associated with the aft bearing in operation are largely
unknown (Biswell, 2007, Cumberlidge, 2009, WRTSIL, 6/09/2007).

The main objective of the studies undertaken previously was to investigate the operational
characteristics of the aft bearing assembly as a part of the propulsion shaft system under
water contamination conditions. These studies provide an understanding of the current aft
bearing design requirements and provided a framework for future investigations on similar
systems (Solomonov, 2009, Solomonov et al., 2010).

3.2 Review of wear study


A comprehensive wear study was conducted by Wrtsil Pty Ltd, a UK-based supplier of
water-lubricated bearing materials (WRTSIL, 6/09/2007, Biswell, 2007). The aim of
this study was to analyse the performance of various bearing materials in highly-abrasive
conditions against stainless steel counterface material (Biswell, 2007, Cumberlidge, 2009,
WRTSIL, 6/09/2007). According to Biswell (2007), Substitute sea water was used,
38

CHAPTER 3 Previous experimental studies of tribological characteristics in water-lubricated bearings

which had silica particles added. The grit used was equivalent in particle size and shape to
that found in the Portland area of the UK, at concentration accepted by the UK Ministry of
Defence as representing of aggressive British coastal water. To simulate the worst water
conditions, and to accelerate the test, the concentration of silica was increased by a factor
of 10. The grit was kept in suspension in the sea water by a stirrer agitating the mixed
solution in the contaminated water tank. A pump was used to deliver the contaminated sea
water to and from the bearing. The flow rate was set at 7.5 litres per minute (Biswell,
2007, Cumberlidge, 2009, WRTSIL, 6/09/2007). Table 3.1 details the test parameters.

Table 3.1 Basic parameters for the Wrtsil wear tests (Biswell, 2007)

Parameters

Metric Units

Sleeve: Stainless steel

ENISO 316

Sleeve diameter

50.8 mm

Shaft rotation

8.8 m/min (55 rpm)

Bearing load

2500 N

Bearing pressure

4.8 kg/cm2

Pressure velocity (PV) rating

42 kg/cm2.m/min

Lubricant flow rate

7.5 litres/min

Lubricant tank capacity

88 litres, agitated

The initial testing comprised running each material under the adopted conditions for a
duration of 100 hours, where the wear rate was measured at 20-hour intervals (Biswell,
2007, Cumberlidge, 2009, WRTSIL, 6/09/2007). To keep experimental results
consistent, all bearings were tested with the same geometry and configuration. This was
done to ensure that the performance and not the design of the bearing material was tested.
The results from the experimental wear tests for the different bearing materials are shown
in Figure 3.1.

39

CHAPTER 3 Previous experimental studies of tribological characteristics in water-lubricated bearings

Figure 3.1 Experimental wear rates for water-lubricated bearing materials (Biswell,
2007, Cumberlidge, 2009, WRTSIL, 6/09/2007)
As can be seen from the Figure 3.1, although all the materials were tested under the same
conditions, a difference in wear performance noticeable. Most materials performed well
during this period of time, except the elastomeric material, which showed significant
bearing wear. Smearing of the bearing material on the shaft and scoring on the shaft liner
were noted (Biswell, 2007, Cumberlidge, 2009, WRTSIL, 6/09/2007).

3.3 Review of the aft bearing study


According to Kotousov (2009, p.12), the physical (scaled test rig) modeling approach was
used by researchers from the School of Mechanical Engineering, at the University of
Adelaide, to simulate a water-lubricated aft bearing system. The friction conditions of the
scaled physical model (test rig) were simulated as close as possible to the actual system.
This is to ensure the similarity of the dynamic behaviour between the scaled model and
actual aft bearing systems behaviour.

3.3.1 Experimental apparatus


The test rigs, presented in Figure 3.2, were built to investigate the effect of water
contamination by applying the following aft bearing system design control parameters
(Kotousov, 2009):

Lubrication contamination;

Radial clearance;
40

CHAPTER 3 Previous experimental studies of tribological characteristics in water-lubricated bearings

Sliding speed;

Bearing pressure;

Coolant flow.

A total of three test rigs were manufactured and two types of lubrication conditions were
used: clean tap water, and contaminated tap water with added sand particles to simulate
contamination. The water was maintained at room temperature.
According to Kotousov (2009), the contaminated conditions were not specified but the
adding of sand particles seemed to be an appropriate way to create the contaminated
environment and investigate the qualitative response of the system to the contamination.
The duration of the long-term testing was approximately four months. It allowed the
achievement of a similar wear pattern as in an operational aft bearing system. This time
was sufficient to achieve a steady-state wear rate and to enable theoretical methods to be
applied to extrapolate the wear process for longer time periods.

Arch Support
Sleeve Housing
Assembly

Variable Speed Motor


Mounting Base

Shaft
Assembly

Load Cell

Figure 3.2 Scaled test rig and major components, as displayed in Kotousov (2009), p.
13, Figure 3.2.1
The first test rig simulated the long-term operation of the aft bearing system in clean water
lubrication conditions. The second test rig aimed to simulate the long-term operation of the
aft bearing system in a contaminated water regime. The third test rig was used to
investigate the effects of alignment, lubrication contamination, radial clearance, sliding
41

CHAPTER 3 Previous experimental studies of tribological characteristics in water-lubricated bearings

speed, and roughness of the sliding surfaces on the tribological behaviour of the whole
system.

3.3.2 Experimental results of the friction coefficient measurements


The value of the friction coefficient in the system was investigated using a radial clearance
of 1 mm and high flow rates. Smooth, quiet and oscillation-free regimes were easily
achievable for large radial clearance.

Measurements of friction were conducted using strain gauges mounted on the vertical arm
supporting the sleeve housing. Two sets of measurements were taken at different points on
the arm to reduce errors, and also to demonstrate the consistency of the measurement
technique.

Table 3.2 provides a summary of the parameters of the bearing system used to estimate the
thermal regime and power losses due to friction and vibrations.

Table 3.2 Basic parameters of the bearing system (Kotousov, 2009, Solomonov et al.,
2010)
Parameters

Experimental Rig

Bearing pressure

0-0.6 MPa

Actual flow rate

0-1.0 l/s

Angular speed range

200-3000 rad/s

Shaft diameter

30 mm

Figure 3.3 presents the results of the experimental investigation of the coefficient of
friction against the sliding speed/shaft rotation speed under clean water lubrication.

42

CHAPTER 3 Previous experimental studies of tribological characteristics in water-lubricated bearings

0.03

Friction coefficient

Elastohydrodynamic theory of lubrication, Pan et


al. (1971)
0.02

Trend line
0.01

0
0

700

500

1000

1500

Shaft rotation frequency, rpm

0.79

1.1

1.6

2.4

Sliding velocity, m/s

Figure 3.3 Friction curves vs. sliding speed, m/s (rotation speed, rpm), reproduced
from Kotousov (2009, p.18)

These results indicated that the friction coefficient is significantly affected by the sliding
speed. It changes from 0.03 at low speeds to 0.002 at high speeds of rotation (high sliding
speeds). It is interesting to note that these tendencies are in agreement with the theoretical
predictions made by Pan et al. (1971), especially for high sliding speeds. This boosts
confidence in the experimental technique adopted in this study. These results provide
information for the determination of the operating and temperature conditions as well as
for estimating the power losses in the oscillation-free regime.

Kotousov (2009) reported that the level of water contamination in the aft bearing system
was unknown and the results presented were aimed at providing a qualitative assessment of
the effect of the contaminated lubricant on friction, vibration, and noise characteristics of
the water-lubricated bearing system. Moreover, the water contamination was not constant
with time and it was extremely difficult to control and characterise the level of
contamination.

43

CHAPTER 3 Previous experimental studies of tribological characteristics in water-lubricated bearings

According to Kotousov (2009), the difference in the friction coefficient between the rig
with contaminated lubricant and the rig with clean lubricant operating in the mixed
lubrication regime was reported to be an order of magnitude. At the same time, the
difference between friction coefficients of the two test rigs operating within the boundary
regime is small, reaching 30-40%. Therefore, for a bearing intended to work in mixed and
hydrodynamic regimes, contamination of the lubricant is one of the major issues. However,
if the bearing is intended to operate in the boundary regime, the level of water
contamination does not appear to exert significant influence on the friction properties of
the system.

3.3.3 Experimental results of the wear tests


The rig for long-run investigations was operated in contaminated and clean lubrication
conditions at a sliding speed V ~1.57 m/s and bearing pressure P = 0.47 MPa. Periodic
measurements of the wear of the bearing shell and shaft were recorded. These results
demonstrated that wear occurred in the bottom part of the shell (Kotousov, 2009).

The wear rate due to water contamination was found to be five times higher than that for
the clean water. The friction in the clean lubricant environment left a smooth shiny surface,
whereas in the contaminated environment, the shell showed signs of heavy damage such as
deep footprints of abrasive damage.

3.3.4 Discussions and conclusions of the experimental tests


The effect of water contamination on values of the friction coefficient and wear were
experimentally investigated and the following conclusions were drawn by Kotousov
(2009):

1.

The characteristics of water-lubricated bearings depend on the operational


conditions of the aft bearing system as well as the material properties. For boundary
and mixed lubrication regimes, complex design considerations need to be
undertaken such as: careful selection of bearing materials, filtering the cooling
water, proper alignment and smooth contact surfaces. None of these steps on its
own is able to significantly improve the water-lubricated bearings performance. If
all steps are implemented, it seems that the rate of water cooling is not a critical
44

CHAPTER 3 Previous experimental studies of tribological characteristics in water-lubricated bearings

factor as, at very low friction conditions; the thermal regime does not require higher
flow rates than are currently used in the aft bearing system. However, this set of
improvements will work if one maintains the operational rotational speed of the aft
bearing above 40 rpm (sliding speed ~0.06 m/s). Below this critical speed of
rotation, the friction versus sliding speed curve for the tested materials has a large
negative slope which will lead and feed the friction-induced vibrations and
excessive wear. Consequently, additional measures are required if the bearing is
intended to operate below 40 rpm.

2.

For the boundary regime of operation, structural modifications to reduce the effect
of water contamination on the bearing assembly are required. In this case, the flow
rates should be much higher than those currently used to avoid overheating and
reduce the wear rates. The water-lubricated bearing system would also benefit from
filtering of the lubricant. This will significantly reduce the wear rates and keep the
surface in good condition.

3.

Other recommendations can be directed to the redesign of water-lubricated bearing


systems and the use of new solutions, such as:

a separate cooling system to keep the temperature of the lubricant low (low
temperature of the lubricant increases its viscosity),

new composite materials with low friction properties, specifically at the


required low operating speeds,

design of the additional lift pressure system to create a film layer between the
shaft and shell, etc.

3.4 Conclusions
The aim of the experimental studies (Biswell, 2007, Cumberlidge, 2009, Kotousov, 2009,
WRTSIL, 6/09/2007) was to provide a series of scaled model tests demonstrating the
long-term running and wear characteristics and the tribological behaviour of waterlubricated bearings under clean and contaminated water-lubricated regimes. The test rigs
simulated the lateral loading associated with the aft bearing system of a drive shaft used in
large boats and ships.

45

CHAPTER 3 Previous experimental studies of tribological characteristics in water-lubricated bearings

Their primary goal was to increase the efficiency and maintenance intervals of waterlubricated bearing systems.

The experimental programs (Biswell, 2007, Cumberlidge, 2009, Kotousov, 2009,


WRTSIL, 6/09/2007) were undertaken to identify and provide understanding of the
general problems of the aft bearing system currently used in ship and submarine
applications. However, the fundamental sources of those problems, such as the effects of
water contamination and damping on the tribological behaviour of water lubricated
bearings materials, were not investigated due to the limitations of the scaled model of
water lubricated bearing test rig capability.

Further extensive experimental and theoretical studies are therefore required to properly
investigate and understand the tribological behaviour and the physical consequences of a
variation in the material properties of composite materials used for water-lubricated
bearings.
The control of friction, wear, and vibrations in moving machine parts is a major problem
for many manufacturers and designers. It has become crucial to have quantitative data
obtained at varying operational conditions and in the presence of lubrication and
contamination (AIphaISS, 2009). POD experimental tribometers have proven their
reliability in many laboratories worldwide (AIphaISS, 2009, Dweib and D'Souza, 1990,
Aronov et al., 1983, Tworzydlo et al., 1994, Ibrahim, 1994a, Ibrahim, 1994b, Mosleh et al.,
2002, Qiao and Ibrahim, 1999, Tworzydlo et al., 1999, Chowdhury and Helali, 2007) for
studying new materials (ceramics, metals, polymers, composites); lubricants and
contaminant additives; self-lubricating systems; and quality assurance.
Compared to the scaled model of water-lubricated bearing test rig, a Pin-on-Disk
experimental apparatus can provide the following features (AIphaISS, 2009):

Precisely calibrated friction, vibration, and wear measurements

A pin is loaded onto the test disk with a more precisely known force

Stable contact point and no parasitic friction

Variable sample materials, sizes, and shapes

Tests conducted in dry, liquid-lubricated, and controlled contamination

46

CHAPTER 3 Previous experimental studies of tribological characteristics in water-lubricated bearings

PC-controlled data acquisition and instrument control, featuring real-time graphical


data display, friction coefficients, wear and vibration data, sliding lifetime, etc.

Based on the analysis of the previous experimental approaches and, together with the
conclusions outlined in within this chapter, the further POD experimental program was
developed and undertaken to investigate the effect of water contamination on the friction
coefficient, vibration, wear and the vibration-wear relationship under varying operational
conditions (Solomonov, 2009, Solomonov et al., 2010). The results and the discussion of
the POD experimental study are presented in the following chapters of this thesis.

47

CHAPTER 4 Experimental apparatus

CHAPTER 4

EXPERIMENTAL APPARATUS

4.1 Introduction
A thorough understanding of the tribological behaviour of water-lubricated contact
surfaces during the design and development of new engineering applications is vital.
According to the literature review, the boundary and mixed regimes cause many
tribological problems. These problems can include vibration, power loss and excessive
wear. The simulation of friction in boundary and mixed regimes is difficult. To obtain a
good prediction of the overall performance of water-lubricated bearings, the precise
measurements of friction, vibration, and wear in the boundary and mixed lubrication
regimes are critical. The friction coefficient, vibration, and specific wear rate for a range of
sliding speeds, applied normal loads, and contaminated lubricants need to be collected and
analysed. In order to use this data in simulations and obtain a better understanding of the
tribological behaviour of sliding bodies, it is preferable to conduct these measurements on
small test samples where different operational conditions can be applied.

This chapter presents the test rig used for the experimental study of friction. Only the main
elements of the test rig and data acquisition methods are discussed.

4.2 Design requirements


A measurement method based on the use of a Pin-on-Disk (POD) technique can fulfil the
previously-mentioned demands such as the precise investigation of the dependence of
friction, vibration, and wear on sliding speeds, applied loads, water contamination, etc.
This method can yield a better understanding of tribological behaviour under waterlubricated conditions (Tworzydlo et al., 1999, Godfrey, 1995, Marklund and Larsson,
2008).

POD tribometer is able to precisely measure the magnitude of friction, vibration, and wear
between two interacting surfaces (Dweib and D'Souza, 1990, Aronov et al., 1983,
Tworzydlo et al., 1994, Ibrahim, 1994a, Ibrahim, 1994b, Mosleh et al., 2002, Qiao and
Ibrahim, 1999, Tworzydlo et al., 1999, Chowdhury and Helali, 2007, CSM-Instriments,
2013). In one measurement, a flat or a spherical sample (pin) is placed on the rotating test
48

CHAPTER 4 Experimental apparatus

disk and loaded with a precisely known weight. The resulting frictional forces and
vibrations between the pin and the disk are measured and recorded. Additionally, the wear
of the sample is measured and calculated from the volume of the material lost during the
test (CSM-Instruments, 2013).

General requirements for a POD test rig are as following (CSM-Instriments, 2013):

High resolution obtained with unique frictional force measurement system design

Easy calibration processes and procedures

High-precision feedback controlled motor

Easy sample replacement and adjustment

Environmental (operational) configuration (dry, lubrication etc.)

Tests in dry and lubricated regimes, controlled contamination, and damping

Tribological investigation in accordance with the ASTM G 99-04 Standard test


Method for wear testing with a pin-on-disk apparatus was used to collect wear data
(ASTM, G 99 04).

A special POD test rig was designed and built to enable the precise measurements of
friction, vibration and wear for water-lubricated bearings materials. This POD test rig has
the capacity to be used for the study of friction, vibration and wear behaviour of any solid
combination of materials, for varying times, applied loads, sliding speeds, damping,
lubricants, contaminants, and fluids.

In POD tests, a load cell with a pin is held by an arm and loaded axially so that the pin is in
contact with a rotating disk, as shown in the schematic diagram in Figure 4.1.

49

CHAPTER 4 Experimental apparatus

1. Arm
2. Disk
3. Load cell
4. Pin
5. Strain gauge sensor
6. Displacement gauge
sensor

Figure 4.1 Schematic diagram of a Pin-on-Disk experimental apparatus

The initial stages of this project involved the design, building, and commissioning of a
POD test rig to study the effect of contaminated lubrication on friction, vibration, and wear
with variable sliding speeds and loads. The test rig needed to be robust and reliable with all
major control parameters such as sliding speed, applied normal load, water lubrication and
contamination, and friction force monitored and recorded using a computer-based data
acquisition device. Appropriate samples of available materials, operational conditions
including a range of sliding speeds, applied loads and levels of water contamination needed
to be chosen before commencement of the design process.

In accordance with the deficiencies discussed in Chapter 3, design requirements were


selected. The major design requirements are shown in Table 4.1.

50

CHAPTER 4 Experimental apparatus

Table 4.1 Design requirements for the Pin-on-Disk test rig


Requirements

Range

Load force, N

8-50

Coefficient of friction,

0.1-1

Sliding speed, m/s

0.3-4.5

Lubrication regimes:

Dry
Clean water
Contaminated water
(with 1%, 2%, 4% and 6% levels of
contamination)

Pin diameter, m

0.01

Disk diameter, m

0.3

Variable speed motor available

BALDOR: MM3550C-57

Adjustable frequency drive

GENESIS: NEMA-4X/IP-65

4.3 Variable sliding speed and applied load


4.3.1 Variable sliding speed
Employing a range of sliding speeds during the experimental work was essential for
obtaining frictionspeed, vibrationspeed and specific wear ratespeed curves.
Adjustments were made using a variable-speed drive system. The main purpose of the
variable-speed drive system was to enable the rotating disk to rotate at an appropriate
speed during the experiments. The variable drive system consisted of a three-phase AC
motor (BALDOR: MM3550C-57) with an adjustable frequency drive (GENESIS: NEMA4X/IP-65). The variable drive system mounted on the test rig is shown in Figure 4.2.

51

CHAPTER 4 Experimental apparatus

a) Motor

b) Frequency drive

Figure 4.2 Variable drive system: a) 3-phase BALDOR: MM3550C-57 motor with b)
GENESIS: NEMA-4X/IP-65 adjustable frequency drive

The AC motor had the following major parameters:

230/415 volt/50 Hz/3 phase

Fully reversible

Rolled steel stator frame

Met the requirements of Australian Standard AS1359, including MEPS Minimum


Efficiency.

The AC motor was equipped with the programmable, adjustable frequency drive
GENESIS: NEMA-4X/IP-65 to set the appropriate rpm speed.

The adjustable frequency drive (NEMA-4X/IP-65) had the following major parameters:

Multifunction keypad with 4-digit LED display

Simplified group programming

8 LED status indicator

Rated for 208-230/400-460 volt/50 and 60 Hz.

To obtain an appropriate sliding speed on the POD test rig, the AC motor was mounted on
the same base as the test rig and connected to the disk assembly using a pulley system. The
disk to motor pulley ratio was 13:1. This enabled a sliding speed range of 0.3 - 3.69 m/s,
which was required for the current experimental work. A digital tachometer model DT52

CHAPTER 4 Experimental apparatus

1236L was used for the calibration of the rotational speed. The calibration data of rpm
(LED display) versus actual sliding speed (at the point of pin-disk contact measured using
the DT-1236L) are shown in Figure 4.3. The calibration was conducted in accordance with
ASTM E74-06 Standard Practice of Calibration of Force-Measuring Instruments for
Verifying the Force Indication of Testing Machines (ASTM, 2006).

Figure 4.3 Calibration data for disk rpm versus sliding speed for the POD test rig

4.3.2 Variable applied load


To investigate friction and wear within the required range of applied loads, it was
necessary to make the applied load variable. The applied load was achieved by sliding
along the central axis of the pin, which was perpendicular to the rotating disk surface.

The load is applied using a series of dead weights together with a load cell assembly that
can slide vertically on the arm and produce a normal force of 8 N. To increase the value of
applied normal load, blocks each with weights of 0.97 kg (applied normal force 9.5 N)
were bolted to the top of the load cell an M10 screw. An example of a normal load of 17.5
N is shown in Figure 4.4.
53

CHAPTER 4 Experimental apparatus

Figure 4.4 Load cell with one additional weight block (applied force 17.5 N)

4.4 Bending arm


The bending arm has two main functions: to hold the load cell in an appropriate position of
contact, and to collect data on the friction force as voltage outputs using its effective
length. Effective length of the bending arm a length of the thin and relatively flexible
part of the bending arm with strain gauges attached, as shown in Figure 4.5.

ASTM G 115-04 Standard Guide for Measuring and Reporting friction Coefficients
(ASTM, 2004b) was used for the bending arms design. According to this guide, the
measurement of dynamic friction force using elastic beams (bending arms) is one of the
recommended methods. To meet this standard guidance, the bending arm as a major
component of the friction force measuring system of the POD test rig weas designed and
fabricated to be elastic enough to measure the dynamic frictional force and stiff enough to
avoid the influence of negative bending arms dynamic motions, which could lead to stick slip behaviour or resonance. The major design requirements for the bending arms are
shown in Table 4.2.
54

CHAPTER 4 Experimental apparatus

Table 4.2 Design requirements for the bending arms used on the POD test rig
Requirements

Range

Load force, N

8-50

Coefficient of friction,

0.1-1

Safety factor, F.S.

where:
F.S. = (y/ n)
n allowable normal stress, N/m2
y - yield stress, N/m2
Allowable strain range, strain

(1 10)10-3

Cross section factor, a=h/b,

where:
h hight of cross section area, m
b wigth of cross section area, m

Resonance condition

n >> max

(This condition is to avoid an effect of

where:

resonance on vibration measurement).

n natural frequency
max maximum frequency of disk
rotation

Stiffness condition

KS >> Ke

(This condition is to separate the

where:

supporting part and the effective part of an KS stiffness of supporting part of


arm and avoid any effect of supporting

bending arm

part on stiffness of effective part and

Ke stiffness of effective part of bending

vibration measurement).

arm and n is a number of bending arm (n1,2,3 or 4)

To meet these requirements, a maximum allowable normal stress, a natural frequency, and
stiffness for each arm needed to be calculated. The one degree of freedom cantilever beam
55

CHAPTER 4 Experimental apparatus

model with one fixed and one free end was used for these calculations (Sundararajan,
2009, Bhattacharjee, 2013, Gere, 2002, Feodosiev, 1974).

The allowable normal stress is calculated using:


(N/m2),

n= y/ F.S

(4.1)

where:
y is the yield stress (N/m2);
F.S. is the safety factor.

The cross sectional moment of inertia for a rectangular cross section area is given by:
I=(bh3)/12

(m4),

(4.2)

where:
b is the width of cross section area (m);
h is the height of cross section area (m).

The stiffness is:


Kn=3EI/L3

(N/m),

(4.3)

where:
E is the modulus of elasticity (N/m2);
L is the effective length of arm (m).

The first natural frequency is:

(Hz),

(4.4)

where:
is the density of the arm (kg/m3);
A is the cross section area of the arm (m2).

The arm can be clamped at any one of nine positions. This allows the changing of the
contact point of the pin on the disk without changing the effective arm length as shown in
Figure 4.5b. The natural frequencies in the frictional direction corresponding to the four
values of stiffness, K1, K2, K3, and K4, were calculated. Table 4.2 presents the main
technical parameters of each bending arm.

56

CHAPTER 4 Experimental apparatus

Table 4.3 Technical characteristics of bending arms


Arm 1

Arm 2

Arm 3

Arm 4

Stainless steel

Stainless steel

Aluminium

Aluminium

18 x 4.5

12 x 3

18 x 4.5

12 x 3

431.26

234.14

255.14

138.88

Stiffness, kN/m

117.1

23.14

40.3

7.96

Effective length, m

0.225

0.225

0.225

0.225

Material
Cross-section area,
h, mm x b, mm
Natural frequency,
Hz

The arms were manufactured from stainless steel or aluminium bar with a rectangular
cross-sectional area. To cover all ranges of applied forces, four bending arms were
fabricated.

To collect friction force data, four strain gauges were mounted to each arm. One of the
bending arms equipped with strain gauges is shown in Figure 4.5.

a)

b)

Figure 4.5 Bending arm with attached strain gauges

To measure the actual friction force, all arms were calibrated, friction force versus voltage
output, using a Digital Force Gauge, model 475040. For further experimental work,
bending arm 1 was chosen. The calibration data for bending arm 1 is shown in Figure 4.6.
The calibration was conducted in accordance with ASTM E74-06 Standard Practice of
57

CHAPTER 4 Experimental apparatus

Calibration of Force-Measuring Instruments for Verifying the Force Indication of Testing


Machines (ASTM, 2006) and Digital Force Gauge, model 475040 operation technical
manual.

Figure 4.6 Calibration data friction force versus voltage output for bending arm 1

4.5 Load cell


According to the test rig design requirements, the load should be applied along the central
axis of the pin which is perpendicular to the rotating disk surface. For this purpose, a load
cell was designed and built.

The load cell can slide vertically on the bending arm. The arm can only bend horizontally,
providing a measurement of the actual friction force. At the same time, the load cell allows
the pin holder with the pin to slide within the load cell in a horizontal direction if
necessary.

Friction force can be measured by a displacement accelerometer which is mounted on the


load cell in a position perpendicular to the pin and parallel to the disk surface. The
accelerometer is held in position by two springs placed inside as shown in Figure 4.7. The
58

CHAPTER 4 Experimental apparatus

springs are flexible in the direction tangential to the disk assembly. This allows measuring
of a sliding displacement of the pin within the load cell. The load cell is equipped with a
displacement sensor as shown in Figure 4.7.

a)

b)

Figure 4.7 Load cell with pin and displacement sensor


For precise measurement of the low range of friction forces (0-10 N), where the bending
arm is not sensitive to low friction force, the load cell displacement sensor was calibrated
using a Digital Force Gauge, model 475040. Calibration data for the load cell are shown in
Figure 4.8. The calibration was conducted in accordance with ASTM E74-06 Standard
Practice of Calibration of Force-Measuring Instruments for Verifying the Force Indication
of Testing Machines (ASTM, 2006) and Digital Force Gauge, model 475040 operation
technical manual.

Figure 4.8 Calibration data friction force vs. voltage output for load cell
59

CHAPTER 4 Experimental apparatus

4.6 Data acquisition system


In order to monitor experimental work and collect experimental data for further analysis
and calculations, a data acquisition (DAQ) system was designed and built. The initial
requirement for the DAQ system was to record the voltage outputs due to friction forces
from the displacement sensor and the bending arm separately during friction tests.

A computer fitted with a DAQ card was used for recording the voltage outputs from the
displacement sensor and strain gauges. The USB-1408FS DAQ card was manufactured by
Measurement Computing Corporation. To acquire and log data from the DAQ card, the
applications InstaCal and TracerDAQ strip chart/data logger were installed on the desktop
computer. This allowed to record and analyse data received from the displacement sensor
and strain gauges separately or simultaneously. A schematic DAQ system diagram is
shown in Figure 4.9.

Figure 4.9 Schematic diagram of the data acquisition system used for the
measurements of pin displacement and arm forces on the POD test rig

The two amplifiers were used to take the input charge from the load cell accelerometer and
arm strain gauge to provide an output in the form of acceleration. The DAQ device
incorporates built-in integrators which provide displacement outputs.

The minimum computer system requirements were:

Windows 2000 or XP

Microsoft. NET framework v1.1 or 2.0

Adobe Reader

Pentium, 90 MHz
60

CHAPTER 4 Experimental apparatus

RAM 96 MB

Microsoft mouse or compatible pointing device.

4.7 Water supply system


In order to investigate the different lubrication regimes, a simple water supply system was
designed and built. The water supply system contains the following components:

Water tank with tap, capacity of 3 litres

Laboratory magnetic variable stirrer

PVC pipe, diameter 5 mm and length 1.5 m

Water nozzle attached to the load cell

Water collector for water removal

PVC pipe of diameter 10 mm and length 1.5 m which is connected to the test rig
water collector to drain contaminated water.

To simulate the worst of the water-lubricated conditions, and to accelerate the comparative
friction and wear tests, the maximum percentage of contamination was increased by a
factor of 10 with respect to the normal concentration of sand in sea water (Polan et al.,
1981). The grit was kept in suspension in the water using a magnetic stirrer agitating the
water and silica in the supply tank. The schematic diagram of the water supply system is
shown in Figure 4.10.

Figure 4.10 Schematic diagram of the water supply system used on the POD test rig
61

CHAPTER 4 Experimental apparatus

The lubrication supply was uniform and thick enough to cover the Pin-Disk contact area.
Ordinary tap water was used as a lubricant because of its low viscosity. A rate of one drop
per second was applied to the disk track at a distance, approximately one cm ahead of the
pin-disk contact point. Visual control during experimental study confirmed that this water
rate was uniform and thick enough to create full water coverage in the Pin-Disk contact
area. The silica sand was introduced as a contaminant. The particle size range of the sand
was measured to be 53-106 m. The concentration range of the contamination was set at
1%-6%. Visual control (followed by microscopy analysis) confirmed that the sand particles
were distributed evenly under the pin in the Pin-Disk contact area.

4.8 Experimental test rig


A design sketch of the test apparatus is shown in Figure 4.11.

Figure 4.11 Design sketch of the test rig identifying the major components

The test rig compromises a horizontally-situated massive base with a vertically-mounted


variable speed AC motor. The motor is connected to a disk assembly through a pulley
system which has a ratio of 13:1. To significantly reduce to acceptable level any influence
62

CHAPTER 4 Experimental apparatus

from the motor vibration to the DAQ system, the motor and the arm are isolated with
rubber pads.

The disk assembly consists a base disk with a test disk that is clamped on top. The test disk
can be changed using a clamping system above the base disk.
The experimental POD test rig, shown in Figure 4.12, was designed, built and
instrumented to investigate the friction between a pair of composite material-steel disk
under dry and water-lubricated conditions.

Figure 4.12 Fully-equipped Pin-on-Disk test rig

63

CHAPTER 4 Experimental apparatus

4.9 Experimental methods


In accordance with the design requirements, the following variables were measured:

Friction force voltage outputs

Mass loss

Speed of rotation

Applied load

Water contamination.

4.9.1 Experimental programs


To investigate the tribological behaviour of water-lubricated bearings materials using the
POD test rig within the required range of applied loads and contamination, the
experimental program was developed. According to the research objectives, this
experimental program consisted of the following experimental investigations:

Validation study

Friction study

Wear study

Vibration-wear study.

Material samples were tested in pairs under dry and water-lubricated conditions. According
to the design requirements and scope of the project, the samples were pins fabricated from
selected material, which is commonly used in water-lubricated bearings and disk fabricated
from selected shaft material. The pin is pressed against the disk at a specified load using
dead weights.

4.9.2 Samples preparation


The POD test required the selection of materials for the pin and the disk. According to the
design requirements and scope of the project, the pin was fabricated from selected waterlubricated bearings material. The pin was fabricated with a flatted tip, and positioned
perpendicular to a flat circular disk. The disk was manufactured from a selected shaft
material.

64

CHAPTER 4 Experimental apparatus

The pins used for experimental runs were cylindrical and had a diameter of 10 mm and a
length of 15-20 mm. The test disk was 300 mm in diameter and had a thickness of 6 mm.
The surface roughness of the pin and the disk were measured to be 0.8 m Ra.

Before each experimental run, the following surfaces preparation procedures were used:

Both, the pin and disk surfaces were cleaned using ethanol and then gradually
polished using progressively finer wet and dry abrasive papers up to extra-fine 1200
grit

The pin was cleaned ultrasonically in an ethanol bath for one minute

The pin after removal from the ultrasonic cleaner ethanol bath and dried using high
pressure air, was weighed to the nearest 0.0001 g, and then carefully placed into the
pin holder of the POD test rig

The disk surface was also cleaned using ethanol and then dried using high pressure
air.

4.9.3 Validation study


The surfaces of the Polytetrafluorethylene (PTFE) pin and the stainless steel disk were
prepared using the same preparation methods discussed in section 4.9.2 before each run.
After sample preparation was completed, the pin was brought into contact with the rotating
disk under a predetermined light normal load. The disk rpm was set corresponding to a
sliding speed of 0.32 m/s. This state was continued for an hour to ensure that there was full
contact between the pin and the disk surfaces. The full experimental load was then applied.

For each load, the friction force (Fx), normal force (Fy) and sliding speed (Vx) were
recorded. A trace of the friction force variation was also measured and recorded. The
friction force readings were taken at a rate of 100 samples per second for the one hour test
period.

The friction force was measured using the voltage outputs from the strain gauges on the
arm, and from these the friction coefficients were calculated. This procedure was repeated
three times. Friction force calculations were verified using a digital forcemeter in
accordance with ASTM E4 - 03 Standard Practise for Force Verification of Testing
Machines. The limit error was calculated to be +1.0 %, which was acceptable in
accordance with ASTM E4 03 (ASTM, 2003).
65

CHAPTER 4 Experimental apparatus

4.9.4 Investigation of friction


The pin (Railko NF22) and stainless steel disk surfaces were prepared, using the methods
described in section 4.9.2, before each run. After sample preparation was completed and
the lubricant introduced, the pin was brought into contact with the rotating disk under a
predetermined light normal load. The disk rpm was set corresponding to a sliding speed
range of 0.3-4.5 m/s. If contamination was to be used, then the range was set to be between
0-6 %. This state was continued for an hour to ensure that there was full contact between
the pin and the disk. The full experimental load in the range of 8-46 N was then applied.

For each load, sliding speed and percentage of contamination, the friction force (Fx),
normal force (Fy), contamination (%) and sliding speed (Vx) were recorded. A trace of the
friction force variation was also obtained and recorded. During the test, friction force
voltage outputs were measured by strain gauges mounted on the loading arm and by the
displacement accelerometer mounted in the load cell. From these the friction force was
calculated. The friction force voltage outputs readings were taken at 100 samples per
second for the period of 60 minutes. For this purpose, a USB-1408FS microprocessorcontrolled data acquisition system was used. Each experimental run was repeated three
times and then the average of the measurements was recorded.

4.9.5 Wear experiments


The wear experiments were conducted in accordance with ASTM G99-04 Standard Test
Method for Wear Testing with Pin-on-Disk Apparatus for pair of Railko NF 22 - Stainless
steel materials (ASTM, 2004a). The pair of materials tested were Railko NF 22 as the pin
and stainless steel as the disk.

At the beginning of each experiment, the pin and the stainless steel disk were polished and
cleaned using the specimen preparation procedure outlined previously. The lubricant was
introduced and the pin and the disk were brought into contact for sufficient time, usually
one hour, to ensure full contact between the pin and the disk surface. The pin was then
removed, the mass measured accurately (up to four decimal places of a gram) and the pin
was placed in the pin holder at its original position. After wear-testing the pin over a
sliding distance of 15,000-16,000 m, the mass of the pin was remeasured using digital
scales and the difference and the mass loss determined. From this the specific wear rate
66

CHAPTER 4 Experimental apparatus

was calculated. Wear tests were repeated three times for each selected sliding distance,
sliding speed, applied load, and contamination level.

Wear tests results were recorded as mass loss in grams and specific wear rate in square
meters per Newton for the pin and plotted as mass loss/specific wear rate versus sliding
speed, applied load, and percentage of contamination.

4.9.6 Vibration-wear relationship study


Vibrationwear tests were conducted under undamped and damped conditions.

The pin (Railko NF22) and the stainless steel disk surfaces were prepared according to
section 4.9.2 before each run. After the lubricant was introduced and the pin was brought
into contact with the rotating disk under a light predetermined normal load. The disk rpm
was set corresponding to a minimum (0.393 m/s) and maximum (1.557 m/s) sliding speeds
and contamination range of 0 - 2 %. After sufficient time for the pin and the disk to
achieve full contact, usually one hour, the pin was removed, the mass was measured
accurately (up to four decimal places of a gram) and then the pin was placed into the pin
holder at its original position. An8 N normal load was then applied for the tests.

For each condition (damped and undamped), sliding speed and contamination, the friction
force (Fx), normal force (Fy), contamination (%), sliding speed (Vx) and mass loss were
recorded. A trace of the friction force variation was also obtained and recorded. During the
test, the friction force voltage outputs were measured by strain gauges mounted on the
loading arm and by a displacement accelerometer mounted in the load cell. The friction
force voltage outputs readings were taken at 100 samples per second for the period of 60
minutes. For this purpose, the USB-1408FS microprocessor-controlled data acquisition
system was used. After testing, the pin mass was measured by the digital scales and the
difference was calculated as mass loss and specific wear rate calculated. Vibration-wear
tests were repeated three times for selected sliding distance, sliding speed, condition
(damped/undamped) and lubrication and the average of mass loss measurements was
recorded.

67

CHAPTER 4 Experimental apparatus

4.10 Microscopy examination


The specimen surfaces were examined using a scanning electron microscope (SEM),
QUANTA 450, and a light microscope, Olympus BX60m, with an Olympus DP21 digital
camera and an Olympus SZH stereomicroscope. The surfaces were photographed and
analysed.

The specimens were cleaned prior to examination with alcohol in an ultrasonic bath for one
minute followed by drying in compressed air.

Using these methods of examination allowed the investigation of the effect of water
contamination on the pin and disk materials microstructures before and after wear tests to
reveal any possible surface changes.

4.11 Conclusions
The POD test rig was designed and built in accordance with the initial design requirements
with some minor adjustments. An experimental methodology was developed.

The purpose of the POD test rig and the developed experimental technique was to analyse
the performance of various bearing materials in dry, clean water-lubricated and highlycontaminated water conditions against a stainless steel counterface material. To try to
simulate the worst of the water-lubricated conditions, and to accelerate the comparative
friction and wear tests, the maximum percentage of contamination was increased by a
factor of 10 with respect to the normal concentration of silica in sea water. A PVC pipe
was used to deliver the gritted solution to the pin-disk contact area.

Using these experimental methods, the following experiments were conducted and the
results are presented in this thesis:

Validation study (Chapter 6)

Experimental investigation of friction characteristics (Chapter 7)

Experimental investigation of wear (Chapter 8)

Experimental investigation of vibration-wear relationship (Chapter 9).

68

CHAPTER 5 Materials

CHAPTER 5

MATERIALS

5.1 Introduction
The metallic antifriction materials (such as gun metal, steel and aluminium alloys) that are
able to work with oil lubrication are conventionally used in liquid-lubricated bearings
designs. In recent years, increasingly close attention has been given to environmentallyfriendly water-lubricated bearings as a cheap and reliable alternative. According to Litwin
(2009), bearings made from polymer-based composites are often used when aggressive
environments prevent the application of other materials (Litwin, 2009).

Another substantial difference between the design of the water-lubricated bearings and
many other bearing systems is the selection process of the materials. The problems with
high performance materials for water-lubricated bearings are particularly relevant to ship
and submarine engineering applications.

According to Rac and Vencl (2005, p.15), apart from adequate strength, the materials for
water-lubricated bearings must also have other crucial characteristics. This is to satisfy the
requirements for reliable and long-lasting operation. These characteristics include low
friction, vibration, good thermal resistance and wear resistance. Rac and Vencl (2005,
p.15) also reported that the mechanical loading is a function of strength of the bearing
materials, while the limits of the thermal loading are determined by the thermal stability of
the selected material (Rac and Vencl, 2005).

Rac and Vencl (2005, p.15) discussed that, with water-lubricated bearings, there is no
direct relationship between the physical and mechanical properties of the materials and the
water-lubricated bearing performance. Parameters, such as the thickness of the lubricant
film, the applied load, and the temperature of the water-lubricated bearing do not depend
on the type of materials used. However, they do have an influence on the materials
behaviour and, as a consequence, on the selection of the suitable materials (Rac and Vencl,
2005).

Rac and Vencl (2005, p.15) also suggested that water-lubricated bearings materials must
also possess a series of other characteristics that are related to the wear resistance and the
69

CHAPTER 5 Materials

surface layer properties. Those characteristics are identified as tribological and include
conformability, embeddability, compatibility, deformation, wear, corrosion, and fatigue
resistance (Rac and Vencl, 2005).
Rac and Vencl (2005, p.15) stated that the tribological behaviour is not just a function of
the bearing material, but also of the surface finish, the lubricant, the design, and the
conditions of the environment in which the bearing operates. The complexity of the
tribological properties of bearing materials and their strong system-dependent properties is
thus explained (Rac and Vencl, 2005).

It was decided that the proposed experimental work to investigate friction, vibration, and
wear using a POD test rig would include wear modes including fracture, tribochemical
effects, and material loss. According to Unal et al. (2004), transitions between regions
dominated by each of these damage modes can give rise to changes in the friction
coefficient, amplitude of vibration, and wear rate with load, sliding speed, and
contamination (Unal et al., 2004). Therefore, the proper selection of materials is essential.

This chapter presents an analysis of contemporary polymer-based materials for waterlubricated bearings and of the selection of the materials available for further experimental
work.

70

CHAPTER 5 Materials

5.2 Effect of water contamination on bearing materials


The viscosity of water is more than 100 times lower than that of oil (Ginzburg et al., 2006).
Figure 5.1 presents the dependence of water viscosity on temperature.

Figure 5.1 Typical baseline of viscosity of water vs. temperature T, 0C, reproduced
from Ginzburg et al. (2006, p.696), Figure 2

The low viscosity of water means that the water-lubricated bearings operate in boundary
and mixed regimes during initial start-up and shutdown. The effects of boundary and
mixed lubrication are determined by the properties of the contacting materials and their
surfaces.

In the case of water contamination, preliminary filters or other devices are commonly used
to remove the contaminants. However, these measures frequently fail to prevent ingress of
soil, sand, clay, or other impurities into the gap between the contacting surfaces. Figure 5.2
provides an example of the severe effect of abrasive wear due to water contamination in
water-lubricated bearings, as reported by Litwin (2009).

71

CHAPTER 5 Materials

a)

b)

Figure 5.2 Water-lubricated bearing damage (subjected to long-lasting operation


which resulted in significant wastage and associated ovalisation of the bush), where
Do=initial dimension, Dp=actual dimension and Dw=wear due to water contamination
as displayed in Litwin (2009, p.44, Figure 6 and p.48, Figure 21)

Polan et al (1981) stated that the abrasive nature of wear due to water contamination needs
to be taken into account when choosing the materials for experimental work. The extent of
this wear depends on the operation regime in the Stribeck curve (Polan et al., 1981).

As contaminants find their own way into the clearance, the nature and extent of subsequent
damage will be determined by the properties of the materials. If the contaminants are softer
than these materials, the damage will be insignificant. If, however, the contaminant
particles are harder than the contacting materials, the softer surface will suffer scuffing. As
was shown in the literature review in Chapter 2, little or no information is available on
abrasive wear resistance when operating with contaminated water.

5.3 Materials for water-lubricated bearings


The choice of materials to be lubricated with water cannot be based on the tribological
behaviour under dry sliding friction. This selection would ignore the chemical aspects of
the interaction between water, the contaminant, and the contacting bodies.

Many friction applications operating in the boundary and mixed regimes of lubrication are
characterised by a low range of sliding speeds, frequent surface interruptions, a wide range
of loads, and a comparatively short sliding distance during the lifecycle.
72

CHAPTER 5 Materials

The main materials used for water-lubricated applications are rubber, ceramics, carbon,
polymer-based materials, and composites of these materials.

5.3.1 Polymer-based thermoplastic materials


Many commercial materials are used for water-lubricated applications. They can be
classified into the following (Ginzburg et al., 2006):

polytetrafluoroethylene (PTFE) and its copolymer materials (Teflon, Polyflon, etc.)

polyamides (PA) (Kapron, Nylon, Rilsan, etc.)

polyformaldehydes (Delrin, Hostaform, etc.)

polycarbonates (Diflon, Lexan, etc.)

polyphenylene oxides (Ariloks, Noryl, etc.)

polyurethanes (Thordon SXL, etc.).

According to Thordon Bearings Inc., comparison tests of different classes of polymer


materials demonstrate that the presence of water as a lubricant in tribocontact reduces the
friction coefficient when compared with the dry friction regime. However, the wear rate
does not directly depend on the friction coefficient, but rather depends on the properties of
the material. The polymer-based thermoplastic materials available for experimental
friction, vibration, and wear tests were PTFE and polyurethane (Thordon SXL).

For the proposed validation study, PTFE was chosen because many previous experimental
studies have been undertaken that are considered PTFE material and its composites for
sliding applications, and its tribological behaviour is well known (Bayer, 2002, Ginzburg
et al., 2006, Ledocq, 1973, Yamajo and Kikkawa, 2004). Typical physical properties of
PTFE are shown in Table 5.1.

73

CHAPTER 5 Materials

Table 5.1 Physical properties of PTFE material

Properties

Value

Density, g/cm3

2.2

Melting point, C

327

Young's modulus, MPa

500

Yield strength, MPa

23

Coefficient of friction

0.05 - 0.1

Hardness, Dur meter D

50

Water absorption, %

0.001

A typical PTFE sample used for the validation study presented in this thesis is shown in
Figure 5.3. In accordance with the initial requirements and test rig design specifications,
this sample was fabricated to have a diameter of 10 mm and a length of 20 mm. It also
contains an M8 threaded section for securing the pin into the pin holder during the tests.

Figure 5.3 PTFE test sample used during the validation study
74

CHAPTER 5 Materials

5.3.2 Carbon - fibre reinforced materials based on thermosetting


materials
The strength and rigidity of polymeric materials are higher if they are thermosetting and
reinforced by fibres or textolites (Latin textus a cloth and Greek lithos stone), materials
which consist of several layers of fabric (filler) connected by synthetic resin.

Phenol-formaldehyde resins have been used for quite some time as a matrix with cotton
cloth or synthetic fibres as the reinforcing materials. A group of materials suitable for
reliable water-lubricated applications is carbon-fibre reinforced composite with
thermosetting matrices.

One of these materials (NF22 [Railko]) was available for this current experimental work.
Railko NF22 material was developed in the early 1980s for water-lubricated and dry
bearing systems and according to Cumberlidge (2009), this bearing material has been
especially designed to cope with extreme operational conditions such as loads, speeds,
temperature fluctuations, water contamination, etc. It is reported by Cumberlidge (2009)
that this material is used by more than 30 Navies around the world for water - lubricated
bearings on ships, ferries and submarines (Cumberlidge, 2009).

Depending on the application and grade, NF22 (Railko) can operate dry, partially
lubricated, or fully lubricated in oil or sea water (Cumberlidge, 2009). Typical physical
properties for the NF22 (Railko) material are shown in Table 5.2.

75

CHAPTER 5 Materials

Table 5.2 Physical properties of NF22 (Railko) material (WRTSIL, 6/09/2007)

Properties

Value

Density, g/cm3

1.64

Ultimate compressive strength, kg/cm2

normal to laminate (radial)

1800

parallel to laminate (axial)

1000

Max. working compressive strength, kg/cm2

radial

450

axial

250

Compressive modulus (radial), kg/cm2

41000

Ultimate tensile strength, kg/cm2

310

Tensile modulus (axial), kg/cm2

310

Normal working pressure, MPa

55

Max. operating temperature, oC

continuous

100

occasional

120

Impact strength, KJ/m


Charpy

35

Izod unnotched

75

Shear strength, MPa

41

Brinell hardness, HB

29

% swell in water, at 20C

<1

Coefficient of friction (dry)

0.25-0.4

Coefficient of thermal expansion, 10-6/C (normal)

radial

60

axial

40

Hardness, KJ/m

85

Thermal conductivity at 50C, W/m.k

0.74

Fire rating per NF F-16-101

F1.I3

NF22 (Railko) material was chosen as the primary material for the experimental
investigation of friction, vibration, and wear because, according to the Wrtsil manual
76

CHAPTER 5 Materials

(WRTSIL, 6/09/2007), the NF22 (Railko) grade of material comprises an anti-scuffing


thermosetting resin reinforced by organic fibres and enhanced with dry lubricants.

An NF22 (Railko) sample used for the experimental study is shown in Figure 5.4. In
accordance with the initial requirements and test rig design specifications, this sample was
fabricated to have a diameter of 10 mm and a length of 15 mm. It also features an M8
threaded section to secure it in the pin holder during tests.

Figure 5.4 NF22 (Railko) sample used in the experimental study

5.3.3 Shaft materials


The shafts of water-lubricated bearings for the marine environment are manufactured from
gun metal (e.g. BS 1400 LG4 or similar), stainless steel (e.g. AISI 440C, AISI 316 or
similar) or a combination of both materials (WRTSIL, 6/09/2007, Yamajo and
Kikkawa, 2004).

To simulate water-lubricated bearing shafts, an AISI 440C stainless steel material was
chosen as the disk material for the experimental studies presented in this thesis. The typical
chemical composition in weight percent (%) is:

Carbon - 1.08;

Silicon - Max 1.0;

Manganese - Max 1.0;


77

CHAPTER 5 Materials

Chromium - 17.0

Molybdenum - 0.75

Phosphorus 0.04

Sulphur 0.03

Iron - Balance.

The test disk was fabricated with the following specifications:

Disk diameter, 300 mm

Disk thickness, 6 mm

Surface roughness of the disk, 0.8 m Ra.

A fabricated stainless steel test disk, fitted on the POD test rig for experimental study, is
shown in Figure 5.5. It has eight clamp holes to secure it to a base disk during the
experimental work.

Figure 5.5 AISI 440C stainless steel test disk fitted on the POD test rig

78

CHAPTER 5 Materials

5.4 Conclusions
In accordance with the materials requirements, the waterlubricated bearings materials
were selected so as to investigate the process of friction, vibration, and wear with the
different load (0-50 N) and sliding speed (0.3-4.5 m/s) ranges in dry, boundary, and mixed
lubrication regimes.

The test disks and sample pins were fabricated from the same materials as used in
commercial water-lubricated bearings, e.g.:

For the validation study: PTFE (pin), stainless steel (disk)

For the experimental study: NF22 (Railko) (pin), stainless steel (disk).

The cylindrical pin specimens, 10 mm in diameter and 15-20 mm in length, were tested
against an AISI 440C stainless steel disk. The surface roughness, prior to the disk testing,
was measured to be 0.8 m Ra.

The aim of this chapter was to analyse the performance of various water-lubricated
bearings materials in dry, water-lubricated and highly abrasive conditions, against a
stainless steel counterface material.
This chapter presented a review of contemporary water-lubricated bearings materials, their
analysis and selection of available materials for the following validation study and
experimental investigations, which are presented in the next chapters of this thesis

79

CHAPTER 6 Validation study

CHAPTER 6

VALIDATION STUDY

6.1 Introduction
A POD test rig validation study program was set up to verify the capability of the newlydesigned and fabricated test rig. These validation friction tests were conducted under dry
lubrication at a sliding speed of 0.32 m/s and a normal applied load range of 8 - 36.5 N.
The results were compared with the published results (Godfrey, 1995, Unal et al., 2004,
Yamajo and Kikkawa, 2004).

6.2 Test plan and procedure


A validation study was conducted to assess the new experimental POD test rig, described
in Chapter 4, against the design requirements and to assess whether a POD test rig was
appropriate for this use. The validation study was conducted in accordance with ASTM
E4-03 Standard Practise for Force Verification of Testing Machines (ASTM, 2003) and
ASTM G115-04 Standard Guide for Measuring and Reporting Friction Coefficients
(ASTM, 2004b).

The purposes of the validation study were:

To quantifiably characterise the POD test rig performance

To assess the potential for errors

To identify method-to-method differences

To determine if the experimental test rig would meet existing regulations (ASTM
E4-03 Standard Practise for Force Verification of Testing Machines and ASTM
G115-04 Standard Guide for Measuring and Reporting Friction Coefficients).

The technical parameters, including the specific test conditions and experimental samples
adopted for the validation study are provided in Table 6.1.

80

CHAPTER 6 Validation study

Table 6.1 Technical parameters adopted for the validation study

Parameters

Experimental Rig

Applied normal force

8N
17.5 N
27 N
36.5 N

Lubrication

Dry

Sliding speed

0.32 m/s

Materials:
Disk
Pin

Stainless steel (AISI 440C)


Polytetrafluorethylene (PTFE)

A number of POD experimental studies have been conducted to examine the influence of
test speed and load values on the friction and wear behaviour of pure PTFE, glass fibre
reinforced PTFE, bronze and carbon filled PTFE polymers (Godfrey, 1995, Unal et al.,
2004, Yamajo and Kikkawa, 2004).

Unal et al (2004) carried out the friction and wear tests versus AISI 440C stainless steel
disk were conducted on a Pin-on-Disk test rig at a dry condition. Tribological tests were at
room temperature, using 5 N, 10 N, 20 N, and 30 N loads and at 0.32 m/s, 0.64 m/s, 0.96
m/s and 1.28 m/s speeds. These experimental results were used to validate a newly built
POD test rig (Unal et al., 2004).

The current validation study, presented in this Chapter, was conducted at room
temperature. To measure the friction force under dry regime, a constant sliding speed of
0.32 m/s was selected. The load was incrementally increased from 8 N to 36.5 N.

6.3 Results and discussions


Figure 6.1 presents variations of the friction coefficient values (between each run) for the
PTFE pin against the stainless steel disk tested at room temperature under dry wear
conditions at 0.32 m/s sliding speed and at 8 N, 17.5 N, 27 N and 36.5 N normal loads.

81

CHAPTER 6 Validation study

Each run was repeated three times. The error bars represent the variation of the friction
coefficients calculated from measured friction forces.

Figure 6.1 Coefficient of friction of the PTFE pin against a stainless steel disk for a
sliding speed of 0.32 m/s

The experimental measurements of the friction forces and calculated friction coefficients
were in good agreement with experimental results reported in the literature (Unal et al.,
2004, Godfrey, 1995, Yamajo and Kikkawa, 2004).

The experimental results of the validation study show that for the PTFE material, the
coefficient of friction decreases with an increase in applied load. This effect is known for
polymers that are visco-elastic materials and their deformation under applied load is viscoelastic. Thus, the variation of the friction coefficient with the load for visco-elastic
materials is given by the following equation (Unal et al, 2004):
=kN(n-1)

(6.1)

where:
is the coefficient of friction;
82

CHAPTER 6 Validation study

N is the normal load;


k is a constant;
n is a constant, its value 2/3 <n <1.
Unal et al (2004) stated that according to this equation, the friction coefficient decreases
with an increasing normal load due to the visco-elastic deformations. In this case, when the
load increases to the load limit values (just before yield) of the polymer, the friction and
wear will decrease due to the critical surface energy of the polymer. Unal et al (2004)
reported that this is due to frictional heat increasing the temperature of the sliding surfaces
which leads to the relaxation of the polymer chains. Therefore, molecules at the polymer
surfaces are compressed, drawn and sheared. Highly active radicals then react with
unbroken chains and give rise to a series of new chains (Unal et al., 2004).

6.4 Conclusions
A newly built POD experimental test rig has been used for a validation study of the effect
of an applied normal load on the friction coefficient under dry friction conditions.

The experimental results of the validation study show that for the PTFE - stainless steel
materials the coefficient of friction decreases with an increase in normal load due to viscoelastic deformations.

It was investigated and confirmed that the obtained results of validation study are in a good
agreement with published experimental results.

Wear and vibration-wear validation is not required for the POD test rig in accordance with
ASTM E4-03 Standard Practise for Force Verification of Testing Machines where A
testing machine shall be verified as a system with the force sensing and indicating devices
in place and operating as in actual use (ASTM, 2003). Digital scales were used to
measure the mass loss of the pin used in these wear and vibration-wear tests. These digital
scales are not a part of POD arrangement; they were verified apart from this project as
laboratory equipment.

Using this information, the following chapter presents experimental investigation of


friction characteristics of NF22 (Railko)-AISI 440C materials pair. This is used to
83

CHAPTER 6 Validation study

experimentally examine the effect of water contamination on the friction coefficient of


water-lubricated bearing materials.

84

CHAPTER 7 Experimental investigation of friction characteristics

CHAPTER 7 EXPERIMENTAL INVESTIGATION OF


FRICTION CHARACTERISTICS
7.1 Introduction
The experimental investigation of friction in mechanical systems depends on many factors.
According to Pilipchuk (2002) some of these factors include the material properties and
geometry of sliding bodies, surface finish, surface chemistry, sliding speed, temperature,
normal load, and type of lubrication (Pilipchuk, 2002).

An experimental program was conducted to examine the effect of varying levels of


contamination of water lubricant on the friction coefficient of materials for waterlubricated bearings. Experiments were conducted at room temperature for the following
ranges:

Sliding speeds, 0.393-1.557 m/s

Applied normal loads, 8-46 N

Contamination, 0-6%.

The experimental study was conducted in accordance with ASTM G115-04 Standard
Guide for Measuring and Reporting Friction Coefficients (ASTM, 2004b).

7.2 Experimental study of the effect of contamination on friction:


experimental plan and procedure
This experimental study of friction was conducted to examine the effect of varying
contaminated lubrication regimes on the friction coefficient of materials for waterlubricated bearings within boundary and mixed lubrication regimes.

The purposes of this experimental study of were to:

Investigate the effect of water contamination on the friction coefficient of materials


used for water-lubricated bearings

Assess material performance under dry and water-lubricated conditions


85

CHAPTER 7 Experimental investigation of friction characteristics

Identify the friction mechanism under dry and water-lubricated conditions.

The selected test parameters for the experimental study of friction characteristics are given
in Table 7.1.

Table 7.1 Operational parameters adopted for experimental study


Parameters

Experimental Rig

Applied normal force

8N
17.5 N
26 N
36.5 N
46 N

Lubrication

Water-lubricated

Water contamination

0%
1%
2%
4%
6%

Sliding speeds

0.393 m/s
0.767 m/s
1.158 m/s
1.557 m/s

Test duration

1 hour

Disk diameter

0.3 m

Pin diameter

0.01 m

Sand particles size

53 106 m

Materials:

Disk

Pin

Stainless steel (AISI 440C)


NF22 (Railko)

86

CHAPTER 7 Experimental investigation of friction characteristics

7.3 Experimental study of the effect of contamination on friction:


results and discussions
7.3.1 Investigation of friction under water lubrication
Two sets of experiments were conducted for both the dry and the clean water-lubricated
conditions at room temperature, namely:

Friction coefficient versus normal load

Friction coefficient versus sliding speed.

The first set of experiments was conducted at room temperature under dry and clean water
lubrication conditions. A constant sliding speed (0.393 m/s) was selected and the load was
incrementally increased from 8 N to 46 N. The friction force was recorded by measuring
voltage outputs from the strain gauges on the arm, and from this data the friction
coefficients were calculated. This procedure was repeated three times for each sliding
speed of 0.767 m/s, 1.158 m/s and 1.557 m/s under both dry and water lubrication
conditions. The purpose of these tests was to collect base data on the variation of the
friction coefficient with changing normal load under dry and clean water lubrication and to
identify the boundary and mixed lubrication regimes for clean water.

The second set of experiments was also conducted under dry and clean water lubrication
conditions. The sliding speed was increased gradually from 0 m/s to the maximum possible
sliding speed of 3.69 m/s at each load of 8 N, 17.5 N, 27 N, 36.6 N and 46 N, respectively.
The friction force voltage outputs were measured and recorded, and friction coefficients
were calculated. This procedure was repeated for all sliding speeds (0.767 m/s, 1.158 m/s
and 1.557 m/s). The purpose of this second set was to investigate the effect of sliding speed
on the friction coefficient and to enable the identification of the boundary and mixed
lubrication regimes with clean water lubrication.

Figures 7.1 and 7.2 illustrate the experimental results for variations of the friction
coefficient values for the NF22 (Railko) pin against the stainless steel disk. The tests were

87

CHAPTER 7 Experimental investigation of friction characteristics

conducted at room temperature under dry lubrication conditions, at 8 N, 17.5 N, 27 N,


36.5 N, and 46 N loads, and at various sliding speeds.

Figure 7.1 Coefficient of friction of NF22 (Railko) material against stainless steel
versus normal applied load under dry conditions

The dependence of the friction coefficient on the relative velocity between sliding surfaces
is complex and depends on the material properties at the interacting surfaces. In dry sliding
contact between interacting surfaces, friction can be modelled as elastic and plastic
deformation forcing microscopic asperities into contact. The friction force can be described
by the following equation reproduced here from Ibrahim (1994a, p. 218, Eq. 15):

F1(Pn)=fas (Aas/At) Pn

(7.1)

where:
F1(Pn) is the friction force due to asperity contact (N);
fas is the asperity sliding friction coefficient;
Aas is the real area of contact (m2);
At is the total area of contact (m2);
Pn is the contact load (N).
88

CHAPTER 7 Experimental investigation of friction characteristics

According to this equation (Ibrahim, 1994a), increasing the load leads to greater contact
between the asperities and therefore, a higher friction coefficient.

Figure 7.2 Coefficient of friction of NF22 (Railko) material against stainless steel
versus sliding speed under dry conditions

For dry conditions, the friction coefficient increases with increasing load and sliding speed.
The frictionspeed curves are non-linear and are dependent on three factors; the material
properties of the sliding surfaces, the normal load, and the medium which occupies the gap
between them. The NF22 (Railko) material has three phases (matrix and organic fibres)
and a solid lubricant. According to Mosleh et al (2002), increasing the applied normal load
increases the friction force due to increasing the area of asperity contact (Mosleh et al.,
2002). The nonlinearity of the friction coefficientspeed curve dependence is in strong
agreement with the literature and is attributed to the creep deformation of the interface
asperities (Martins, 1990). More advanced examples of dry friction models of this
mechanism were described in the literature by Marklund and Larsson (2008) where the

89

CHAPTER 7 Experimental investigation of friction characteristics

friction coefficient was found to be a function of speed and temperature in addition to the
normal load (Marklund and Larsson, 2008).

Figures 7.3 and 7.4 illustrate the experimental results of variations in the friction
coefficient values for an NF22 (Railko) pin against a stainless steel disk tested at room
temperature using clean water lubrication, at 8 N, 17.5 N, 27 N, 36.5 N, and 46 N loads,
and at various sliding speeds.

Figure 7.3 Coefficient of friction of NF22 (Railko) material against stainless steel
versus normal load under clean water-lubricated conditions

It can be seen from Figure 7.3 that the value of the friction coefficient decreases with
increasing load and sliding speed. This is the opposite effect to that found under dry
lubrication conditions. The normal load exerts slightly less influence on the friction
coefficient than sliding speed, particularly at high speeds. Marklund and Larsson also
reported that the friction coefficient was not particularly load dependent (Marklund and
Larsson, 2008).

90

CHAPTER 7 Experimental investigation of friction characteristics

Figure 7.4 better illustrates the relatively small influence that the normal load has on the
coefficient of friction, particularly at higher speeds. The largest difference in the value of
the coefficient of friction for different normal loads is approximately 0.18 which is at the
lowest sliding speed of 0.393 m/s.

Figure 7.4 Coefficient of friction of NF22 (Railko) material against stainless steel
versus sliding speed under clean water-lubricated conditions

This type of friction behaviour (NF22 (Railko) material against stainless steel) was
explained in the literature by the elastohydrodynamic theory of lubrication and theory of
asperity contact (Unal et al., 2004, Godfrey, 1995, Ibrahim, 1994a). In the zones of real
contact, all types of deformation (elastic, elastoplastic, and plastic) can occur. The friction
coefficient for materials with a high modulus of elasticity changes during plastic contact as
the load increases, as shown in Figure 7.4. This change is due to the interaction between
asperities and deformation processes in the zones of real contact. The maximum friction
coefficient conforms to elastic asperities contact and the transition to a state where the
interaction between irregularities begins to affect the deformation processes in the zones of
real contact. The decrease of the friction coefficient due to increasing load is a result of
transition from elastic asperity contact to plastic deformation at the points of contact. The
91

CHAPTER 7 Experimental investigation of friction characteristics

increasing load makes asperities in contact plastically deform which may release the solid
lubricant; in conjunction with water, this provides better lubrication by creating a wedge of
lubricant and reducing the friction coefficient as stated in Unal et al (2004).

It is evident that the friction coefficient of NF22 (Railko) material against stainless steel
under clean water lubrication conditions is much more dependent on the sliding speed than
on the normal applied load. According to the Stribeck model, all this experimental work is
within either boundary or mixed regimes of lubrication. Olsson et al (1997) reported that,
the total friction force at contact consists of two components; one is due to asperity contact
(equation 7.1) and the second is due to hydrodynamic viscous shear (Olsson et al., 1997):

F2(V)=Fc V

(7.2)

where:
F2 is the friction force due to hydrodynamic viscous shear (N);
Fc is a constant, dependent on lubricant viscosity, area of contact and minimum film
thickness of the lubricant in the hydrodynamic regime (N/m/s);
V is the sliding speed (m/s).
The negative slope (boundary and mixed regimes) of the Stribeck model is clearly seen
in the experimental results, shown as Figures 7.3 and 7.4.

7.3.2 Investigation of the friction coefficient and the effect of


contamination
The aim of these experiments was to investigate and analyse the effect of water
contamination on the relationship between friction and speed for NF22 (Railko) material
against an AISI 440C stainless steel disk.

Figures 7.5 to 7.8 present the graphs of the experimental results of the friction coefficient
for NF22 (Railko) against stainless steel under contaminated water-lubricated conditions,
at 8 N, 17.5 N, 27 N, 36.5 N, and 46 N loads, at 0.393 m/s, 0.767 m/s, 1.158 m/s, and
1.557 m/s sliding speeds, and at various levels of water contamination of 1%, 2%, 4%, and
6%, respectively.

92

CHAPTER 7 Experimental investigation of friction characteristics

Figure 7.5 Coefficient of friction of NF22 (Railko) against stainless steel for 1%
contaminated water lubrication

Figure 7.6 Coefficient of friction of NF22 (Railko) against stainless steel for 2%
contaminated water lubrication

93

CHAPTER 7 Experimental investigation of friction characteristics

Figure 7.7 Coefficient of friction of NF22 (Railko) against stainless steel for 4%
contaminated water lubrication

Figure 7.8 Coefficient of friction of NF22 (Railko) against stainless steel for 6%
contaminated water lubrication
These Figures show that the value of the coefficient of friction decreases with increasing
sliding speed and load, but increases with increasing water contamination.

94

CHAPTER 7 Experimental investigation of friction characteristics

According to Olsson et al. (1997), contamination is another factor that introduces


complications. The presence of small particles of foreign material between the sliding
surfaces gives rise to additional forces that strongly depend on the size and material
properties of the contaminants (Olsson et al., 1997).

When the results for contaminated water are compared with those for clean water
lubrication, the tribological behaviour can still be described by the Stribeck model. At
the same time, it can be seen that the friction coefficient increases with increasing
contamination due to the influence of the abrasive sand particles between interacting
bodies at all sliding speeds. Moreover, the maximum increase in the friction coefficient
was seen at the slowest sliding speed of 0.393 m/s due to asperity contact, as well as the
abrasive sand particles. For further analysis of the effect of water contamination on the
friction coefficient, the slowest sliding speed of 0.393 m/s which resulted in the largest
value of the friction coefficient was chosen. The results are shown in Figure 7.9.

Figure 7.9 Coefficient of friction vs. water contamination of NF22 (Railko) material
(sliding speed=0.393 m/s)
As mentioned above, friction in water-lubricated sliding bodies depends on friction due to
asperities contact, lubricant contamination, and hydrodynamic viscous shear friction.
95

CHAPTER 7 Experimental investigation of friction characteristics

Whether any nonlinearities due to the effect of water contamination on friction exist is not
yet known.

As seen on Figure 7.9, the friction force due to water contamination is a function of the
degree of contamination (g).

Thus, the total friction force under contaminated water lubrication can be described as
follows:

F=F1 (Pn)+F2 (V)+F3 (g)

(7.4)

where:
F is the total friction force (N);
F1(Pn) is the friction force due to asperities contact force (N);
F2(V) is the friction force due to hydrodynamic viscous shear force (N);
F3(g) is the friction force due to water contamination (N).
The complex effect of water contamination on friction has not been thoroughly
investigated, however to simplify further investigations of friction due to water
contamination, it is assumed that F3(g) has a linear dependence of friction coefficient
against the level of water contamination (g).

7.4 Conclusions
A new experimental approach using the POD experimental apparatus for the study of the
effect of water contamination on the friction coefficient has been conducted. The
experimental range of sliding speeds was identified as being within the boundary and
mixed (transition) regimes of the Stribeck curve.

The effect of water lubrication and water contamination on the pair of materials
comprising NF22 (Railko) composite and stainless steel was experimentally investigated
and the following conclusions can be drawn:

96

CHAPTER 7 Experimental investigation of friction characteristics

1.

The friction coefficient of NF22 (Railko) composite material decreases when the
applied load is increased due to the material properties (solid lubricating
component and plastic deformation).

2.

The friction studies of the composite material (NF22 (Railko)) against an AISI
440C stainless steel disk under various loads, sliding speeds, and water
contamination show an increase in the friction force with increasing water
contamination.

3.

It is not yet known if any nonlinearities from the effect of water contamination on
friction exist. As a first iteration, it can be assumed that friction force due to water
contamination increases linearly with an increase in water contamination.

4.

For the specific ranges of loads and speeds investigated in this study, the sliding
speed has a greater effect than the applied normal load on the friction coefficient
for this composite material.

The friction in water-lubricated bearings depends on a number of factors such as applied


load, sliding speed, bearing design, and environment conditions. Therefore, the value of
this POD test method is to predict the relative ranking of water-lubricated bearings
materials combinations rather than simulating in service water-lubricated bearings.

However, further extensive experimental and theoretical studies are required to properly
understand the effect of water contamination on friction and the physical implications of
variations in the material properties of composite materials used for engineering
applications.

A mathematical model of contact mechanics between the two friction surfaces is necessary
for the analysis and design of water-lubricated bearings materials under contaminated
water conditions. In this experimental work, a simple empirical linear assumption of
contact mechanics based on experimental results has been proposed (see Equation 7.4) for
water contamination at levels between 1% and 6%.

97

CHAPTER 7 Experimental investigation of friction characteristics

The next chapter uses the information from these experiments as a basis for the
investigation of the effect of water contamination on the mass loss and the specific wear
rate of materials used for water-lubricated bearings.

98

CHAPTER 8 Experimental investigation of wear

CHAPTER 8

EXPERIMENTAL INVESTIGATION OF WEAR

8.1 Introduction
In engineering applications of water-lubricated bearings, where the water contains different
types of contaminants including solid silica particles, the use of wear-resistant materials is
required. According to Prehn (Prehn et al., 2005), in addition to the contamination by silica
particles, the water used for lubrication may also be chemically aggressive. For these
reasons, the choice of materials is limited.

An experimental program was undertaken to examine the effect of contaminated water


lubrication regimes on the mass loss and the specific wear rate of materials used for waterlubricated bearings. This study investigated the sliding response of NF22 (Railko) material
against stainless steel. Experiments were conducted at room temperature for sliding speeds
in the range of 0.393 to 1.557 m/s and applied normal loads of 8 N to 46 N.

The three-body wear mechanism involved is identified as being a function of lubrication,


water contamination, and material properties. The relative wear resistance of samples is
compared using specific wear rate calculations as a function of mass loss, whilst results of
surface microscopy are also presented. The experimental study was conducted in
accordance with ASTM G 99-04 Standard test method for wear testing with a Pin-on Disk
apparatus (ASTM, 2004a).

8.2 Experimental study of wear: experimental plan and procedure


An experimental study of wear was conducted in order to examine the effect of varying
contaminated water lubrication regimes on the mass loss and specific wear rate of
materials for water-lubricated bearings.

The purposes of this experimental study were:

To investigate the effect of water lubrication and contamination on the wear rate of
materials used for water-lubricated bearings
99

CHAPTER 8 Experimental investigation of wear

To assess the material performance under contaminated and non-contaminated


water-lubricated conditions

To identify the significant wear mechanisms for contaminated water-lubricated


conditions.

The parameters used for the experimental study of wear, including specific test conditions
and experimental samples, are provided in Table 8.1.
Table 8.1 Parameters for wear test
Parameters

Experimental Rig

Applied normal force

8N
17.5 N
26 N
36.5 N
46 N

Lubrication

Water-lubricated

Water contamination

0%
1%
2%
4%
6%

Sliding speeds

0.393 m/s
0.767 m/s
1.158 m/s
1.557 m/s

Test duration

1 hour

Disk diameter

0.3 m

Pin diameter

0.01 m

Sand particles size range

53-106 m

Materials:

Disk

Pin

Stainless steel
NF22 (Railko)

This experimental work was conducted at room temperature. The POD test rig was run for
four hours before taking any measurements to ensure full contact between the pin and the
disk surface. For the first run of experiments, a constant sliding speed of 0.393 m/s and a
load of 8 N were used. Each run was repeated three times. After each run, the mass of the
100

CHAPTER 8 Experimental investigation of wear

pin was measured using digital scales and the mass loss recorded. Subsequently, the same
sliding speed was used; the load was increased in steps up to 46 N. The mass losses were
measured and recorded, and the specific wear rates calculated. This procedure was
repeated three times for all sliding speeds.

No measurable mass losses were recorded during wear tests under clean water lubrication.
The specific wear rate was therefore, zero. In the case of clean water lubrication, only
adhesive wear mechanisms are possible and adhesion is significantly reduced by
lubrication. This appears to be due to the specific material properties of the NF22 (Railko)
composite, which has a solid lubricant incorporated into its structure.

8.3 Experimental study of wear: results and discussions

8.3.1 Wear tests


Figures 8.1 to 8.4 present mass loss values for NF22 (Railko) under contaminated water
lubrication conditions at 8 N, 17.5 N, 27 N, 36.5 N, and 46 N loads, at 0.393 m/s,
0.767 m/s, 1.158 m/s, and 1.557 m/s sliding speeds, and at 1%, 2%, 4%, and 6% water
contamination, respectively.

101

CHAPTER 8 Experimental investigation of wear

Figure 8.1 Mass loss for NF22 (Railko) material at different load and speed values
under 1% water contamination

Figure 8.2 Mass loss for NF22 (Railko) material at different load and speed values
under 2% water contamination
102

CHAPTER 8 Experimental investigation of wear

Figure 8.3 Mass loss for NF22 (Railko) material at different load and speed values
under 4% water contamination

Figure 8.4 Mass loss for NF22 (Railko) material at different load and speed values
under 6% water contamination

103

CHAPTER 8 Experimental investigation of wear

From these graphs it is evident that the value of mass loss decreases with increasing sliding
speed and load, and increases with greater water contamination. Olsen et al (1997)
explained this effect as the adhesive wear mechanism is coupled with abrasive wear
mechanism under contaminated water lubrication. The presence of small particles of sand
between sliding surfaces gives rise to additional forces that strongly depend on the size and
material properties of the contaminants (Olsson et al., 1997).

Comparing the results with those of clean water lubrication, it is clear that the wear
mechanism can be described generally by adhesive-abrasive wear. This is where sand
particles are in contact with the surfaces of the pin and the disk, resulting in material being
displaced from the pin as well as the disk. At the same time, the mass loss increases with
increasing water contamination due to the role played by the abrasive sand particles
between contacting bodies. At the same time, it can be seen that the mass loss increases
with increasing water contamination due to the influence of abrasive sand particles
between interacting bodies at all sliding speeds in the range. Moreover, the maximum
increase in the mass loss was seen at the slowest sliding speed (boundary regime of
lubrication) of 0.393 m/s due to interactions between asperities as well as abrasive sand
particles. This speed was therefore, chosen for further analysis of the effect of
contamination on the mass loss. These results are shown in Figure 8.5.

Figure 8.5 Mass loss versus degree of water contamination for NF22 (Railko) material
at a sliding speed of 0.393 m/s
104

CHAPTER 8 Experimental investigation of wear

This graph suggests that, at low load, more sand particles are drawn into the contact area
which leads to greater mass loss. It is expected that with a change in particle size, the
character of this plot could be changed significantly. Figure 8.5 also shows that, the mass
loss due to water contamination is a nonlinear function of the degree of contamination. To
investigate the effect of water contamination as well as the type of contaminant, further
theoretical and experimental work is required. A theoretical three-body wear model is
necessary for determination of the wear mechanism under water contamination for further
stability analysis and design.

The task of obtaining a theoretical three-body wear model for these types of contact is
difficult due to the complexity of this mechanism with input variables of abrasive as well
as adhesive wear. It is therefore recommended for future work.

In the case of contaminated water lubrication, a combination of adhesive and severe


abrasive wear mechanisms occur. As shown in the experimental results, the mass loss
decreases with increased sliding speed and applied load, and increases with greater water
contamination. This can be explained by the boundary and mixed regimes of Stribeck
model of hydrodynamic and three-body wear mechanism (Torrance, 2005).

A micrograph of the pins surface before and after the full cycle of experiments is shown in
Figure 8.6.

a) before

b) after

Figure 8.6 Micrograph of pins worn surface before/after a full cycle of experiments,
at a magnification of X500

105

CHAPTER 8 Experimental investigation of wear

Despite the severe level of contamination, the images are similar and no visible structural
changes to the pins surface or embedded silica particles are evident. Torrance (Torrance,
2005) suggested that the process of wear depends on the mechanics of the contact between
the two surfaces entraining the abrasive particles, on particle size and percentage of
contamination, and often on chemical interactions between the wearing surface and the
surrounding media. It was reported in WRTSIL (2007) that it has generally been
assumed that the presence of organic fibres and solid lubricant in composite materials
makes it wear resistant.

Figure 8.7 presents the micrograph of the same worn surface of the pin. This micrograph
was obtained using an optical stereomicroscope. As seen from this Figure, the NF22
(Railko) material consists of two visible phases (matrix and organic fibres) creating good
wear resistance for this material.

Figure 8.7 Micrograph of pins worn surface after a full cycle of experiments, at a
magnification of X100

The micrographs of wear traces on the paired worn pins surface and stainless steel disk
are shown in Figures 8.8 and 8.9, respectively. Both samples were thoroughly cleaned,
using ethanol and compressed air, before microscopic examination.

106

CHAPTER 8 Experimental investigation of wear

As reported by Moore (1978), for abrasive particle contact with very low loads, the contact
will be predominantly elastic. Such contacts may result in material removal by surface
molecular mechanisms or from surface films or by Hertzian fracture of brittle materials
(Moore, 1978). As the load on an angular abrasive particle increases, contact on both
ductile and brittle materials will involve plastic deformation to a greater extent leading to
surface damage.

Figure 8.8 Micrograph of pins worn surface after a full cycle of experiments

Figure 8.9 Micrograph of stainless steel disks worn surface after a full cycle of
experiments, at a magnification of X100
107

CHAPTER 8 Experimental investigation of wear

It is clear that in spite of the good wear resistance of NF22 (Railko) material, the threebody abrasion wear mechanism could significantly damage both surfaces in a severely
contaminated water-lubricated environment although there are no visible structural
changes to the pin or sand particles embedded in the composite pin or steel disk.

8.3.2 Specific wear rate calculations


To analyse the volume of plastically-displaced material, a specific wear rate needs to be
calculated.

Unal et.al (2004) reported the following calculation procedure which was adopted for this
experimental study. The sliding wear data reported here should be the average of at least
three runs. The average mass loss was used to calculate the specific wear rate (Unal et al.,
2004):
K0=m/L*F*

(m2/N),

(8.1)

where:
m is the average mass loss (kg);
L is the sliding distance (m);
F is the applied load (N);
is the density of the material (kg/m3).

L=v*t

(m),

(8.2)

where:
v is the sliding speed (m/s);
t is the duration of the test (s).

Figures 8.10 to 8.13 illustrate the specific wear rate values calculated for each test sliding
speed and normal load under contaminated water-lubricated conditions, at 8 N, 17.5 N,
27 N, 36.5 N, and 46 N loads, at 0.393 m/s, 0.767 m/s, 1.158 m/s, and 1.557 m/s sliding
speeds, and at 1%, 2%, 4%, and 6% water contamination, respectively.

108

CHAPTER 8 Experimental investigation of wear

Figure 8.10 Specific wear rate for NF22 (Railko) material at different load and speed
values with 1% water contamination

Figure 8.11 Specific wear rate for NF22 (Railko) material at different load and speed
values with 2% water contamination

109

CHAPTER 8 Experimental investigation of wear

Figure 8.12 Specific wear rate for NF22 (Railko) material at different load and speed
values with 4% water contamination

Figure 8.13 Specific wear rate for NF22 (Railko) material at different load and speed
values with 6% water contamination

It is clear that the sliding speed variation has little influence on the specific wear rate of the
NF22 (Railko) material, but the value of the specific wear rate increases with increasing
110

CHAPTER 8 Experimental investigation of wear

water contamination. As mentioned above, the most significant variation of specific wear
rate versus water contamination and applied load was seen at the lowest sliding speed as
displayed in Figure 8.14.

Figure 8.14 Specific wear rate versus degree of water contamination of NF22 (Railko)
material for a sliding speed of 0.393 m/s

Unal et al. (2004) stated that the wear process involves fracture, tribochemical effects, and
plastic deformation. Variation to the applied normal load and contamination lead to
transitions between regions dominated by each of these processes and commonly give rise
to changes in the specific wear rate. Furthermore, this result is closely related to structural
characteristics and chemical effects occurring in the frictional processes, as well as to
transfer film formation on the counterfaces (Unal et al., 2004).

For the NF22 (Railko) material tested in this experimental investigation within the sliding
speed range from 0.393 m/s to 1.558 m/s, the speed and normal load had less influence on
the specific wear rate than changes to the level of water contamination. The main reason
for the specific wear rate increase is the complex three-body wear mechanism, including
abrasive and adhesive wear.

111

CHAPTER 8 Experimental investigation of wear

The main purpose of these wear tests was to analyse the performance of samples of NF22
(Railko) bearing material in highly abrasive operational conditions, when forced against an
AISI 440C stainless steel.

This resultant knowledge provides a base for further investigation into a three-body wear
mechanism under contaminated water conditions. This could lead to the development of a
more advanced three-body nonlinear wear model for water-lubricated bearings and other
engineering applications.

8.4 Conclusions
The effect of contaminated water lubrication on the wear mechanism of the pair of
materials comprising NF22 (Railko) composite and AISI 440C stainless steel was
experimentally investigated. The following conclusions can be drawn:

1.

Mass loss and the specific wear rate of NF22 (Railko) composite material decreases
as the applied normal load increases. This is due to plastic deformation of the
contacting surfaces.

2.

Wear studies of the composite material NF22 (Railko) against the AISI 440C
stainless steel disk under various normal loads, sliding speeds, and water
contamination show no significant mass loss under clean water lubrication due to
the material properties of the composite.

3.

The mass loss and specific wear rate increase nonlinearly with the increase of water
contamination due to the complex adhesive-abrasive wear mechanism.

4.

For the specific range of applied normal load, sliding speed, and level of water
contamination explored in this experimental study, water contamination has a
stronger effect on the mass loss and specific wear rate than the normal load and
sliding speed. This can be explained by the strong influence of abrasive nature of
the wear mechanism.

5.

The wear in water-lubricated bearings depends on a number of factors such as


applied load, sliding speed, bearing design, and environment conditions. Therefore,
112

CHAPTER 8 Experimental investigation of wear

the purpose of this test method is to predict the relative ranking of water-lubricated
bearings materials combinations, but not to simulate wear of in service waterlubricated bearings.

Thus, these experimental results represent a basis for further extensive experimental and
theoretical studies of the wear mechanism and the physical implications of a variation in
the material properties of composite materials used for water-lubricated engineering
applications.

The effect of water contamination and magnetic damping on the vibration and its
correlation to wear was experimentally investigated and results are presented and discussed
in the next chapter.

113

CHAPTER 9 Experimental investigation of vibrationwear relationship

CHAPTER 9 EXPERIMENTAL INVESTIGATION OF


VIBRATIONWEAR RELATIONSHIP
9.1 Introduction
Friction, vibration, and wear in machines are primarily due to the dynamic friction forces
acting in machinery. Friction-induced vibrations and wear on contact surfaces depend on
the combination of contact materials, load, sliding speed, lubrication, and contamination
(Krishna Kumar and Swarnamani, 1997).

Propeller shaft bearing systems on ships and submarines can make use of different types of
water-lubricated systems. A number of selection requirements have been developed and
provided in manufacturers catalogues and design manuals (THORDON, 2006,
WRTSIL, 6/09/2007, Litwin, 2009). These criteria include bearing pressure, sliding
speed, size, type of lubrication, surface finish, lubricant flow, fittings, and machine
tolerances. The essential factors to be considered when designing and selecting waterlubricated bearing materials are possible vibration, noise, power loss, and lifespan. The
manufacturers of most bearings provide detailed life-load analysis where bearing
calculations consider design requirements such as load, lubricant type, and speed of
rotation (sliding speed). According to Chowdhury and Helali (2007), the reduction of wear
and vibration depends on interfacial conditions such as applied load, geometry, relative
surface motion, sliding speed, surface finishes, type of materials, system rigidity,
lubrication, and lubricant contamination (Chowdhury and Helali, 2007). Nevertheless, the
relationships between these main factors (friction, wear, vibration, and contamination) in
water-lubricated bearings are unknown and have generated limited attention in the
literature.

In boundary and mixed regimes occurring in water-lubricated bearings, high three-body


contact pressures elastically and plastically deform the bearings contact surfaces. This
leads to surface fatigue. With improvements in bearing materials and surface preparations,
surface damage due to the vibrationwear relationship from water contamination has
become one of the main challenges associated with the performance of water-lubricated
bearings. Water contains different types of contaminants, which may include solid silica
particles and swear debris material particles as a consequence of the wear itself.
114

CHAPTER 9 Experimental investigation of vibrationwear relationship

According to Maru (2007), vibration measurement is commonly used for condition


monitoring of rotating machinery, being relatively easy to use. In this context, this chapter
presents an experimental investigation into the effect of water contamination on the
vibrationwear relationship for water-lubricated bearings materials using a modified POD
test rig. The contacting characteristics of sliding bodies were considered in order to
determine the existence of correlations between tribological aspects and vibration
phenomena (Maru et al., 2007b).

The ASTM G 99-04 Standard test Method for wear testing with a Pin-on Disk apparatus
was used to collect the wear data (ASTM, 2004a).

9.2 Operational conditions and experimental methods


An experimental program was conducted to examine the effect of varying contaminated
water lubrication regimes on the vibrationwear relationship of materials for waterlubricated bearings operating in boundary and mixed regimes. This study investigated the
lubricated sliding response of NF22 (Railko) material against stainless steel. Experiments
were conducted at room temperature for sliding speeds 0.393 m/s and 1.557 m/s under an
applied load of 8 N. To investigate the effect of damping on the relationship between
vibration and wear, a strong magnet was used as a magnetic field damper. Experiments
were repeated three times for each run for both damped and undamped conditions.

115

CHAPTER 9 Experimental investigation of vibrationwear relationship

Figure 9.1 POD test rig equipped with a magnet

Figure 9.1 shows the POD test rig equipped with a strong magnet mounted on the rig base.
The magnet is independent from the sliding load cell containing the pin. The magnet was
positioned within a millimetre of the load cell for the investigation of the vibrationwear
relationship under damped conditions.

An experimental study of the effect of water contamination and damping on the vibration
wear relationship was conducted to examine the effect of varying contaminated water
lubrication regimes and damped and undamped conditions on mass loss, specific wear rate,
and the frequency characteristics of the induced vibrations of water lubricated bearings
materials.

The purposes of this experimental study were to:

Investigate the effect of water lubrication and contamination on the relationship


between vibration and the wear rate of materials for water-lubricated bearings
under boundary and mixed regimes of lubrication

116

CHAPTER 9 Experimental investigation of vibrationwear relationship

Investigate the effect of damping of vibrations on the relationship between


vibration and the wear rate of materials for water-lubricated bearings materials

Assess the materials performance under water-lubricated conditions

Identify the effect of damping on the wear mechanism under water - lubricated
damped and undamped conditions.

The technical parameters adopted for the experimental study, including specific test
conditions and experimental samples, are shown in Table 9.1.

Table 9.1 Technical parameters adopted for the vibrationwear experiments

Parameters

Experimental Rig

Applied normal force

8N

Lubrication conditions

Water-lubricated damped
Water-lubricated undamped

Water contamination

0%
0.5%
1%
2%

Sliding speeds:
Boundary regime
Mixed regime

0.393 m/s
1.557 m/s

Calibration coefficient

9.0 N/volt

Test duration

2 hours

Disk diameter

0.3 m

Pin diameter

0.01 m

Sand particles size range

53-106 m

Materials:
Disk
Pin

AISI 440C Stainless steel


NF22 (Railko)

This experimental work was conducted at room temperature. The preparation procedure
included polishing the surfaces of the pin and the disk using various grades of wet and dry
silicon carbide paper down to 1200 grit.
117

CHAPTER 9 Experimental investigation of vibrationwear relationship

Two sets of tests were conducted under both: damped and undamped conditions. The
duration of each test was one hour for vibration-wear data acquisition. Prior to the
commencement of the tests, the POD test rig was run for an hour before taking any
measurements. This was done to ensure that there was full contact between the pin and the
disk surfaces. During all vibration and wear tests, the vibration parallel to the sliding
direction was measured and the friction-induced vibration voltage output signal from the
bending arm strain gauges was recorded using a USB-1408FS data acquisition device.

For the first run of experiments, a constant sliding speed of 0.393 m/s (boundary regime)
and a load of 8 N under clean water lubrication in undamped conditions were used. During
the test, a friction-induced vibration voltage output signal from the bending arm strain
gauges was recorded at a sampling rate of 0.01 sec/sample. Before and after each run, the
mass of the pin was measured and the mass loss calculated and recorded. At the same
sliding speed, clay free sand was introduced as the contaminant. The friction-induced
vibration voltage output signal and mass losses were measured and recorded, and the
friction force, Welch power spectrum densities, and specific wear rates were calculated and
analysed. This procedure was repeated for a sliding speed of 1.557 m/s (mixed regime).
The experimental runs were repeated three times at each condition and the average value
for the three runs was recorded for each data point.

To investigate the effect of damping on vibration and wear, the same set of experiments
was conducted using a damping magnet. Since it is generally known that the specific wear
rate depends on multiple variables such as load, sliding distance, sliding speed, property of
materials, and type of lubrication and contamination, the purpose of these two sets of
experiments was to collect data for further analysis of the effect of water lubrication, water
contamination, and damping on the vibrationwear relationship.

At sliding speeds of 0.393 m/s and 1.557 m/s the two different lubrication regimes
(boundary and mixed) were characterised by the behaviour of the friction force versus
time, the nature of pin surfaces as examined under the scanning electron microscope, and
the specific wear rate measured as mass loss per unit of sliding distance. The effect of
increasing water contamination and damping on the lubrication regimes was also
characterised.
118

CHAPTER 9 Experimental investigation of vibrationwear relationship

9.3 Experimental study of effect of water contamination on


vibrationwear relationship: results and discussions
9.3.1 Friction forcetime analysis
Figure 9.2 shows typical damped and undamped friction forcetime graphs for the NF22
(Railko) material in clean water and various levels of contaminated water lubrication. All
tests were conducted at a sliding speed of 0.393 m/s conditions, under an 8 N load and
under 0%, 0.5%, 1%, and 2% water contamination, respectively. The measurements were
obtained and recorded at the final stage of the two hours duration test.

119

CHAPTER 9 Experimental investigation of vibrationwear relationship

a) Contamination 0.0%, sliding speed 0.393 m/s

b) Contamination 0.5%, sliding speed 0.393 m/s


120

CHAPTER 9 Experimental investigation of vibrationwear relationship

c) Contamination 1%, sliding speed 0.393 m/s

d) Contamination 2%, sliding speed 0.393 m/s


Figure 9.2 Calculated friction force (N) versus time for undamped and damped
conditions under load 8 N and sliding speed 0.393 m/s
121

CHAPTER 9 Experimental investigation of vibrationwear relationship

Under clean water lubrication (see Figure 9.2 a), the process of friction can be described by
an adhesive mechanism of interaction under the boundary lubrication regime (Aronov et
al., 1983). As the water contamination is increased above a certain critical value, which in
this case is 0.5 %, a transition occurs from an adhesive to an adhesive-abrasive friction
process. This is characterised by the appearance of higher-amplitude unstable lowfrequency vibration shown in Figure 9.2 b), c), d). These unstable low-frequency
vibrations, due to the solid silica particles, are responsible for the formation of the threebody abrasive interaction coupled with an adhesive mechanism of friction. The typical
average friction force value does not change significantly and is within the range of 2-3 N.
At the same time, damping significantly decreases the amplitude of vibrations. The
maximum vibration suppression was observed with clean water lubrication and a 0.393 m/s
sliding speed. This was explained in the literature by disabling of adhesive wear
mechanism by damping of the amplitude of vibration (Maru et al., 2007a, Maru et al.,
2007b).
Figure 9.3 shows typical damped and undamped calculated friction forcetime graphs for
the NF22 (Railko) material in clean and various levels of contaminated water lubrication.
In this case the tests were conducted at a sliding speed of 1.557 m/s conditions, under an 8
N load and under 0%, 0.5%, 1%, and 2% water contamination, respectively. The
measurements were obtained at the final stage of the 60-minute duration test.

122

CHAPTER 9 Experimental investigation of vibrationwear relationship

a) Contamination 0.0%, sliding speed 1.557 m/s

b) Contamination 0.5%, sliding speed 1.557 m/s

123

CHAPTER 9 Experimental investigation of vibrationwear relationship

c) Contamination 1%, sliding speed 1.557 m/s

d) Contamination 2%, sliding speed 1.557 m/s


Figure 9.3 Calculated friction force (N) versus time for undamped and damped
conditions under load 8 N and sliding speed 1.557 m/s

124

CHAPTER 9 Experimental investigation of vibrationwear relationship

The friction can be described by the adhesive mechanism of interaction in the mixed
(transition) lubrication regime (Aronov et al., 1983), when clean water lubrication is used
(as shown in Figure 9.3 a). Comparing these results with those for the low sliding speed of
0.393 m/s results (shown in Figure 9.2 a), the reduction of vibration is due to a mixed
lubrication regime when a lubricant film is formed and resulting in reduced adhesive
interaction. When 0.5 % contamination was introduced, the complex three-body
mechanism of friction begins to play a significant role and an increased amplitude of
vibration was observed (Litwin, 2009, Meuter, 2006). A transition occurs from an adhesive
to an abrasive friction process, characterised by the appearance of high-amplitude unstable
low-frequency vibrations as seen in Figure 9.3 b), c), d) when the water contamination was
increased above a critical value of 1.0%. A similar effect was described my Maru et al.
(2007a) and (2007b). Similar to the low sliding speed, these unstable low-frequency
vibrations appear result of the solid silica particles producing for a three-body interaction
coupled with a mixed regime of lubrication. As the highest sliding speed is in the mixed
lubrication regime, the typical average friction force was lower compared to the low
sliding speed results and was within the range of 1-2 N. At the same time, damping
significantly decreased the amplitude of vibrations. The maximum vibration suppression
was observed for clean water-lubrication. This again can be explained by the disabling of
the adhesive wear mechanism by the damping of the amplitude of vibration as reported in
Maru et al. (2007a) and (2007b).

A significant reduction of vibration by damping was observed at 2% contamination and


1.557 m/s sliding speed. Greater amplitudes of vibrations with a normal load are due to
mixed lubrication at a high sliding speed and a higher level of water contamination when
solid silica particles are dragged between the interacting sliding surfaces of the pin and
disk. Vibration is reduced by damping and enables a stable gap between the two surfaces to
be formed during mixed lubrication (Chowdhury and Helali, 2007).

9.3.2 Power spectral density analysis


Welch power spectral density analysis (periodogram method) is used to estimate the power
of signal (vibration) at different frequencies. This method is based on converting the time
signal into frequency domain using Fourier transform.

125

CHAPTER 9 Experimental investigation of vibrationwear relationship

Welch power spectral density analyses were conducted to establish the conditions for lowfrequency vibration generation. The effects of undamped and damped conditions, water
lubrication and contamination, sliding speeds, and friction forcetime histories were
analysed.

Figure 9.4 presents the Welch power spectral density values data for NF22 (Railko) for
clean and contaminated water lubrication under an 8 N load, at a 0.393 m/s sliding speed,
and at 0%, 0.5%, 1%, and 2% water contamination, respectively, with and without
damping. The measurements were taken during the final stage of a 60-minute duration test.

126

CHAPTER 9 Experimental investigation of vibrationwear relationship

a) Contamination 0.0%, sliding speed 0.393 m/s

b) Contamination 0.5%, sliding speed 0.393 m/s

127

CHAPTER 9 Experimental investigation of vibrationwear relationship

c) Contamination 1%, sliding speed 0.393 m/s

d) Contamination 2%, sliding speed 0.393 m/s

Figure 9.4 Welch power spectral densities for damped and undamped conditions
under load 8 N and 0.393 m/s, sliding speed

Figure 9.4 presents the boundary regime of lubrication with different levels of
contamination. It is clearly seen that damping reduces the dominant resonance frequency
128

CHAPTER 9 Experimental investigation of vibrationwear relationship

of the system, typically in the 12-14 Hz frequency band. As the water contamination is
increased above a certain critical value, which in this case is 0.5%, the suppression of
resonance frequencies becomes more difficult. As mentioned above, transition occurs from
an adhesive to an adhesive-abrasive friction process, characterised by the appearance of
high-amplitude unstable low-frequency vibrations (Figure 9.4 b), c), d)). These unstable
low-frequency vibrations appear as soon as the solid silica particles become responsible for
the formation of the three-body interaction coupled with the adhesive friction mechanism.
As a result of the contamination, the intensity of vibration was increased by approximately
20 db/Hz. This type of vibration (low frequency and high intensity) at a low sliding speed
was generated due to the boundary regime of lubrication (Bhushan, 1980).

Figure 9.5 presents the Welch power spectral density values data for NF22 (Railko) with
clean and contaminated water lubrication under an 8 N load, at 1.557 m/s sliding speed and
at 0%, 0.5%, 1%, and 2% water contamination, respectively, with and without damping.

129

CHAPTER 9 Experimental investigation of vibrationwear relationship

a) Contamination 0.0%, sliding speed 1.557 m/s

b) Contamination 0.5%, sliding speed 1.557 m/s

130

CHAPTER 9 Experimental investigation of vibrationwear relationship

c) Contamination 1%, sliding speed 1.557 m/s

d) Contamination 2%, sliding speed 1.557 m/s


Figure 9.5 Welch power spectral densities for undamped and damped conditions
under load 8 N and 1.557 m/s, sliding speed

Figure 9.5 represents the Welch power spectral densities in a mixed regime of lubrication
with different levels of contamination. It is clearly seen that damping significantly reduces
the dominant resonance frequencies of the system as well as self-induced low-frequency
131

CHAPTER 9 Experimental investigation of vibrationwear relationship

vibrations, typically up to 15 Hz frequency band. At high sliding speeds, a few


power/frequency peaks were observed. These frequencies appear to be directly associated
with the sliding system (mixed lubrication) and excited by an adhesive-abrasive
mechanism of interaction coupled with the mixed regime of lubrication, which is similar to
that reported by Bhushan (Bhushan, 1980).

Overall, the level of power increases with increasing contamination due to the active
abrasive three-body interaction in undamped conditions. According to Torrance (2005) and
Younes (1993), the introduction of damping significantly decreases the intensity of
vibrations with a similar tendency of the power spectral densities to increase with
increasing water contamination (Torrance, 2005, Younes, 1993).

As seen from the results, low-frequency vibrations (up to 15 Hz) play the most significant
role in the vibrationwear relationship under these experimental conditions. No highfrequency vibrations were recorded during this set of experiments.

9.3.3 Vibrationwear analysis


For further vibrationwear relationship analysis, the root mean square (RMS) needs to be
calculated and analysed as it is one of the most important characteristics of vibration
analysis. The RMS parameter of the vibration signal was taken as a rough indicator of the
average level of vibration (Maru et al., 2007b, Maru et al., 2007a).

In the case of a set of n values

, the RMS value is given by equation

9.1:

(9.1)

where:
xn is a value of friction force recorded at a rate of 0.01 samples/sec;
n is the number of recorded samples.
According to previous results of calculated friction force time data and Welch power
spectral densities analysis, the low frequency (LF) band is much more sensitive to the
132

CHAPTER 9 Experimental investigation of vibrationwear relationship

operational conditions, being significantly affected by water contamination and damping


(Maru et al., 2005).

Table 9.2 presents the RMS calculated values for NF22 (Railko) in water-lubricated
conditions (clean and contaminated water lubrication), under an 8 N load, at 0.393 m/s and
1.557 m/s sliding speeds, and at 0%, 0.5%, 1%, and 2% water contamination, respectively,
with and without damping.

Table 9.2 Calculated RMS acceleration values at different sliding speeds, lubrication,
and contamination conditions, under an 8 N load

Calculated RMS acceleration values


Contamination,
%

Sliding speed,
m/s

Undamped

Damped

Difference
in RMS,
%

0.393

2.85443

1.562226

45.0

1.557

1.187191

0.748596

36.9

0.393

7.927545

4.666254

41.1

1.557

3.31157

1.445673

56.3

0.393

6.258963

3.179442

49.2

1.557

4.55773

2.067589

54.6

0.393

2.657814

2.168903

18.4

1.557

3.142426

1.927

38.7

0.5

Comparing the values of the RMS differences (RMS, %) between undamped and damped
conditions, it is seen that under a high level of water contamination (2%) the effect of
damping is much lower for both sliding speeds. At the same time, with an increasing
sliding speed, the RMS values reach a maximum at 0.5% contamination for a low sliding
speed and at 1% contamination at a high sliding speed. This is due to both adhesive and
abrasive wear mechanisms, which are in turn due to the transition from a boundary to a
mixed regime.

133

CHAPTER 9 Experimental investigation of vibrationwear relationship

Figure 9.6 illustrates the final experimental RMS values for NF22 (Railko) against the
stainless steel disk with various degrees of water contamination (0%, 0.5%, 1%, and 2%,
respectively), under undamped and damped conditions under a load of 8 N.

Figure 9.6 Calculated RMS values for NF22 (Railko) material at normal and damped
load (8 N)

These results show the differences in vibration levels for measured operational conditions,
obtained from tests performed under clean and contaminated lubrication. It can be seen in
Figure 9.6, that vibrations in the overall low-frequency bands are greatly affected by
applied operational conditions, especially by damping (which decreases the vibration for
all levels of contamination and at all sliding speeds), the level of contamination, and
sliding speed. For a low sliding speed, the damping is most effective and the highest value
(peak) of the RMS was recorded for 0.5% water contamination. With an increasing sliding
speed, the peak of vibration (RMS) decreased and also moved to the 1% water
contamination point. This is due to the transition from a boundary to a mixed regime (more
lubricant appears between interacting surfaces). It can be seen that vibrations in lowfrequency bands are greatly affected by operational conditions, especially by sliding speed
(a regime of lubrication), which is in strong agreement with results from the literature
(Maru et al., 2007a, Maru et al., 2007b, Maru et al., 2005). Figure 9.6 also shows that

134

CHAPTER 9 Experimental investigation of vibrationwear relationship

vibration is significantly affected by damping, especially at a low-speed range. This is due


to the lower frequency band of vibration generated.

Some of the main observations are highlighted below. As it was discussed by Maru et al.
(2007), the same trend is observed in all operational conditions in which the vibration level
increases with an increase of contaminant concentration and then decreases for higher
concentrations up to the limit value (Maru et al., 2007a). Another observation concerning
the graph (see Figure 8.6) is related to the comparison between levels of contamination. It
is seen that an increase in vibration is seen at low contamination levels. This shows a
significant change in the dynamic behaviour of the pin-disk system which is connected to
the dependency of the vibrationwear relationship on contamination. It is recommended
that these observations be investigated further.

For wear results, no measurable mass losses were recorded during tests under clean waterlubricated conditions. In the case of clean water lubrication, only the adhesive wear
mechanism is possible, which is significantly affected by the type of material and lubricant
used. This is due to the specific material properties of the NF22 (Railko) composite which
contains a solid lubricant in its structure.

Table 9.3 presents the average mass loss values for NF22 (Railko) under water-lubricated
conditions, at 0%, 0.5%, 1%, and 2% water contamination, under an 8 N load and at
sliding speeds of 0.393 m/s and 1.557 m/s, respectively, with and without damping.

135

CHAPTER 9 Experimental investigation of vibrationwear relationship

Table 9.3 Average mass loss (g) at different sliding speeds, lubrication, and
contamination conditions, under an 8 N load

Average mass loss, g


Lubrication

Contamination,
%

Sliding speed,
m/s

Undamped

Damped

Difference
in W, %

0.393

0.0000

0.0000

0.00

1.557

0.0005

0.0000

0.00

0.393

0.0043

0.0004

-3

1.557

0.0061

0.0105

+42

0.393

0.0064

0.0048

-33

1.557

0.0097

0.0216

+45

0.393

0.0118

0.0054

-55

1.557

0.0177

0.0334

+53

0.5
Water
1

Comparing the values of mass loss (W, %) between undamped and damped conditions, it
is seen that at a low sliding speed, the damping leads to a mass loss reduction due to the
boundary regime of lubrication. At a high sliding speed, the value of mass loss increases
due to a more aggressive abrasive wear mechanism coupled with the mixed regime of
lubrication. It is also seen from Table 9.3 that the value of mass loss increases accordingly
with increased water contamination.

Figure 9.7 presents the final experimental results of calculated specific wear rates of NF22
(Railko) against the stainless steel disk under water-lubricated conditions at various
degrees of water contamination (0%, 0.5%, 1%, and 2%), respectively.

136

CHAPTER 9 Experimental investigation of vibrationwear relationship

Figure 9.7 Specific wear rate for NF22 (Railko) material for undamped and damped
conditions

The results show two different vibrationwear relationship trends exist: one for low
(boundary lubrication), and one for high (mixed lubrication) sliding speeds for the
conditions used.

The adhesive wear mechanism plays the major role for the low speed range (boundary
regime) when zero or very little lubricant film is presented between the interacting
surfaces, even if it is coupled with abrasive wear. The reduction of vibration by damping
does not allow sand particles to move between the sliding surfaces and make a strong
three-body contact. This leads to a decrease of surface three-body interaction and, as a
consequence, the specific wear rate decreases.

At high speed (mixed lubrication), there is mostly an abrasive wear mechanism due to
lubricant being forced between the interacting bodies. This allows more sand particles to
be drawn into the contact area. Damping reduces vibration, but maintains a stable gap
between the pin and the disk which allows contaminated lubricant to be more aggressive
(Maru et al., 2007a).
137

CHAPTER 9 Experimental investigation of vibrationwear relationship

In spite of much lower mass loss and insignificant differences in surface damage between
undamped and damped conditions, the maximum increase in the specific wear rate was
obtained at the slowest sliding speed (0.393 m/s) in undamped conditions. The clearest
pictures of significant difference in the pins surface damage were found at the high-speed
range. The microscopy of the pin surface under undamped and damped conditions at a high
sliding speed (1.557 m/s) after a full cycle of experiments is shown in Figure 9.8.

a)

Pin surface before test

b) Undamped condition, contamination 0.5% c) Damped condition, contamination 0.5%

138

CHAPTER 9 Experimental investigation of vibrationwear relationship

d) Undamped condition, contamination 1% e) Damped condition, contamination 1%

f) Undamped condition, contamination 2% g) Damped condition, contamination 2%

Figure 9.8 Microscopy of the pins worn surface before and after a full cycle of
experiments at a high sliding speed of 1.557 m/s, at a magnification of X30

Regardless of vibration suppressions by damping and different levels of contamination, the


photos in Figure 9.8 show greater surface damage for damped conditions than for
undamped conditions. As discussed previously by Torrance (2005), the process of wear
depends on the mechanics of the contact between the two surfaces employing the abrasive
particles, on particle size and contamination, and often on chemical interactions between
the wearing surface and the surrounding media (Torrance, 2005). At a high sliding speed
range, when a film of lubricant is formed and damping suppresses vibration, more sand
particles move between the pin and the disk due to the mixed regime of lubrication. The
presence of an increased quantity of sand particles between the sliding surfaces gives rise
139

CHAPTER 9 Experimental investigation of vibrationwear relationship

to additional forces that strongly depend on the size and material properties of the
contaminants (Olsson et al., 1997).

9.4 Conclusions
In this chapter, water-lubricated bearings materials were experimentally investigated in
order to study the effect of water contamination on vibration and the correlation of
vibration to wear. The methods of vibration analysis (Welch power spectral analysis and
RMS versus specific wear rate) were effective in characterising the trends in the vibration
wear relationship due to water contamination.

The effect of water contamination on the vibrationwear relationship was recognisably


different in nature from that of the level of contamination. The level of vibration (RMS)
increased with the contamination level, and then stabilising at a specific limit. At the same
time, when water contamination was increased, the specific wear rate also increased.

Comparing the results for undamped and damped conditions, it can be seen that damping
can reduce vibration and wear due to adhesive wear at the lower range of sliding speeds,
which represents the boundary lubrication regime. At the same time, it is apparent that due
to the mixed lubrication regime, mass loss and specific wear rate increase by adding
damping at the higher range of speeds. The vibrations are suppressed by damping; sand
particles are free to move into the contact area between the pin and the disk due to the
influence of abrasive wear, sliding speed, and the formation of a lubrication film between
the interacting surfaces.

It is expected that, with a change of particle size, the character of the vibrationwear
relationship could be changed significantly due to the prevailing nature of wear
mechanisms. The results obtained can be used for further investigation into the effect of
damping, water contamination, and the type of contaminant on the vibrationwear
relationship. Further theoretical and experimental work and a theoretical three-body
vibrationwear model for simulation of the vibrationwear relationship under damping and
water contamination for stability analysis and design can be undertaken based on the
results presented here. The task of obtaining a theoretical three-body vibrationwear model
for this type of contact is difficult due to the complexity of this mechanism, which includes
140

CHAPTER 9 Experimental investigation of vibrationwear relationship

abrasive wear as well as adhesive wear. This task can be the centrepiece for a future
theoretical study.

As mentioned above, the specific wear rate in water-lubricated sliding bodies depends on
the vibrationwear relationship due to asperities contact, vibration, and abrasive
interaction between contacting bodies and the contaminant. As seen from the results, the
vibrationwear relationship due to undamped or damped conditions and water
contamination is a nonlinear function of the degree of contamination. The character of this
function is still not known, nor is it known how it can be numerically depicted. The present
methodology can be adopted for further experimental and theoretical investigations which
need to be undertaken to identify the complex vibrationwear relationship for different
types of materials, contaminants, and operational conditions.

It is well known that the vibrationwear process involves fracture, tribochemical effects,
and plastic deformation. Transitions between regions dominated by each of these
commonly give rise to changes in the vibrationwear relationship with undamped or
damped conditions and contamination (Song, 2008). Furthermore, these results are closely
related to structural characteristics and chemical effects occurring in frictional processes,
as well as transfer film formation on the counterface as described in the literature (Unal et
al., 2004, Maru et al., 2007b, Bryant and York, 2000, Ling Wu et al., 2010, Maru et al.,
2007a, Krishna Kumar and Swarnamani, 1997, Chowdhury and Helali, 2007). For the
NF22 (Railko) material tested in this experimental investigation and for sliding speeds of
0.393 m/s and 1.557 m/s, the sliding speed and load values as well as contamination have
shown significant influence on the vibrationwear relationship. They represent a complex
function of all of these variables. Friction-induced vibrations have been found to be a
function of sliding speed, operational (undamped or damped) conditions, and load. They
are produced by a complex mechanism of adhesive-abrasive wear at the bearing interfaces.
According to Krishna Kumar and Swarnamani (1997), this is due to the failure of
lubrication and other wear modes that can be considered to be dependent on the nature of
the surfaces and their physical properties (Krishna Kumar and Swarnamani, 1997). Those
physical properties can include surface roughness, hardness, and elastic modulus.

The main purpose of this vibrationwear experimental study was to analyse the influence
of vibration damping and water contamination on the vibrationwear relationship for NF22
141

CHAPTER 9 Experimental investigation of vibrationwear relationship

(Railko) bearing material in low and highly-abrasive conditions against a stainless steel
counterface material.

The effect of damping and water contamination on the vibrationwear relationship of a


pair of materials comprising an NF22 (Railko) composite and AISI 440C stainless steel
was experimentally investigated and shows very interesting results. In general terms, the
data obtained can be used to explained this relationship in terms of the complex three-body
adhesive-abrasive mechanism of the interaction of interacting materials, as a function of
load, sliding speed, sliding distance, and lubrication (Krishna Kumar and Swarnamani,
1997).

The following conclusions can be drawn from these experimental results:

1.

Specific wear rate of NF22 (Railko) composite material decreases when damping is
applied at a low sliding speed during a boundary lubrication regime.

2.

Specific wear rate of NF22 (Railko) composite material increases when damping is
applied at a high sliding speed during a mixed lubrication regime.

3.

Vibrationwear experimental studies of the composite material (NF22 [Railko])


against an AISI 440C stainless steel disk under constant load and either undamped
and damped conditions, at low and at high sliding speeds, show no mass loss under
clean water lubrication.

4.

Vibration increases and then decreases and specific wear rate increases with greater
water contamination. The highest value of RMS for the low sliding speed was
recorded at 0.5% contamination and for the high sliding speed at 1% water
contamination.

5.

For the specific applied load of 8 N, at low and high sliding speeds, and with the
water contamination explored in this experimental study, damping had a stronger
effect on specific wear rate. It is a function of sliding speed and a result of the
change in lubrication regime between selected sliding speeds.

142

CHAPTER 9 Experimental investigation of vibrationwear relationship

6.

The vibration-wear relationship in water lubricated-bearings depends on multiple


factors that include applied load, sliding speed, bearing design, and environmental
conditions. Thus, the value of this POD test method is to predict the relative
ranking of materials pairs and not for modelling vibration-wear for in service waterlubricated bearings.

7.

Future experimental studies must include an examination of the effect of active


magnetic dampers to facilitate adequate water-lubricated bearings materials
performance.

Thus, the POD experimental methodology and the results presented can be used for further
extensive experimental and theoretical studies of the vibrationwear relationship of various
combinations of materials with different material properties used for water-lubricated
engineering applications.

143

CHAPTER 10 Conclusions

CHAPTER 10 CONCLUSIONS AND RECOMMENDATIONS


10.1 Summary
The main aim of these experimental investigations was to develop new methods for the
experimental investigations of the tribological characteristics of water-lubricated bearings
materials using Pin-on-Disk test rig.

A review of the literature showed that the effect of water contamination on waterlubricated bearing materials and systems performance is highly complex. It depends on
many factors such as the chemical reactivity and corrosion characteristics of the materials
to water and any dissolved salts it may contain and the abrasive nature of soil and sands
entrained in the water. Limited attention has been given in the literature in terms of
analysing and predicting the effect of water contamination on tribological characteristics of
materials for the design and modelling of water-lubricated bearings. Previous experimental
studies, conducted at the School of Mechanical Engineering, at the University of Adelaide,
were analysed, and used as a starting point for this research.

To measure the friction coefficient, vibration, and wear due to the effect of water
lubrication and water contamination, the operational conditions and test rig design
requirements were identified, a new experimental approach was developed, and a specific
Pin-on-Disk (POD) experimental test rig was designed and built. This experimental
apparatus enabled the measurement of the local friction coefficient and vibration as well as
wear investigation for the material combinations used in water-lubricated engineering
applications.

Under the initial experimental requirements, the pair of materials chosen was an NF22
(Railko) composite and stainless steel (AISI 440C).

In accordance with the initial requirements and operating conditions, the POD test rig was
built and ordinary tap water was introduced as a lubricant. To model contamination, fine
sweeping sand with a particle sizes range between 53-106 m was chosen as a contaminant
and added to the lubricant.

144

CHAPTER 10 Conclusions

Experimental investigations of friction and wear were conducted under water lubrication
with a contamination range of 0% to 6% by weight for an applied load range of 8 N to 46
N and a sliding speed range from 0.393 m/s to 1.557 m/s.

The vibrationwear relationship was experimentally investigated under clean and


contaminated water lubrication with a contaminant range of 0% to 2%. For an applied load
of 8 N, the sliding speeds were 0.393 m/s and 1.557 m/s and undamped and damped
conditions were applied.

An analysis of friction, vibration, and wear test experimental results was conducted and
showed the significant effect of water contamination and damping on the friction
coefficient, vibration, wear, and calculated specific wear rate. It was found that there was a
significant increase in the friction coefficient, vibration, and specific wear rate at the
slowest sliding speed of 0.393 m/s. This is due to the boundary regime of lubrication, the
adhesive-abrasive wear mechanism, and specific material properties of NF22 (Railko)
material. At the same time, wear can be decreased by the damping of vibration at slow
sliding speeds.
The microscopy analysis of the pins surface was undertaken before and after wear
experiments and confirmed that wear resistance can be significantly improved by adding
organic fibres and solid lubricant to composite bearing materials. Two components of
complex wear mechanisms (adhesive and abrasive) were considered and analysed. It was
shown that material mass loss and specific wear rate depend mostly on abrasive wear.

As mentioned above, boundary and mixed regimes are the reasons for many tribological
problems such as vibration, power loss, and excessive wear. The simulation of friction in
boundary and mixed regimes is difficult due to their strong and complex interdependency.
To acquire an adequate prediction of material performance, for water-lubricated
applications, friction, vibration, and wear, and theoretical models in the boundary and
mixed lubrication regimes are still required and could be the next stage of further waterlubricated bearing modelling development. As a result of the analysis presented here, it is
suggested that additional components responsible for water contamination and damping
should be added to existing theoretical friction models. The three-body wear model can be
modified, based on the results obtained, as a response to the abrasive nature of water
145

CHAPTER 10 Conclusions

contamination. In addition, this new POD experimental approach can be used during
development and selection of new materials for water-lubricated bearings.

146

CHAPTER 10 Conclusions

10.2 Conclusions
In this experimental study, the effect of water contamination on the friction behaviour of a
pair of materials comprising a composite and stainless steel was experimentally
investigated. The following conclusions can be drawn from the results regarding the
friction coefficient, vibration, and wear for water-lubricated bearings materials:

Limited attention has been given in the literature to the effect of water
contamination on friction, vibration, and wear in water-lubricated bearings
materials and systems.

The friction coefficient of NF22 (Railko) composite material against AISI 440C
stainless steel decreases, when the load and sliding speed are increased under clean
water lubrication. This is a result of boundary and mixed regimes of lubrication and
the elastic and plastic deformation between the surfaces.

No mass loss was observed under both dry friction and clean water lubrication for a
composite material (NF22 [Railko]) against an AISI 440C stainless steel disk under
constant load, undamped and damped conditions, and low and high sliding speeds.

Mass loss and specific wear rate increase nonlinearly with an increase in the water
contamination level. In this experimental work, a simple empirical linear model of
contact mechanics, based on experimental results, has been offered for the range of
1% to 6% levels of water contamination as a first iteration.

For the specific range of loads and sliding speeds investigated in this experimental
study, load has a stronger effect than sliding speed on the friction, vibration, and
wear behaviour of composite material.

Damage to the frictional surface is indicative of boundary and mixed lubrication


regimes, as well as three-body abrasive wear. It is caused by the breakdown of the
lubricant film and by the formation and shearing of the adhesive junctions, as well
as the significant influence of abrasive three-body contacts.

147

CHAPTER 10 Conclusions

Under a boundary lubrication regime, the specific wear rate of NF22 (Railko)
composite material decreases when damping is applied at a low sliding speed.

Under a mixed lubrication regime, the specific wear rate of NF22 (Railko)
composite material increases when damping is applied at a high sliding speed.

For the vibrationwear relationship, it was found that the greatest value of RMS for
a low sliding speed was recorded at 0.5% water contamination and for a 1.557 m/s
sliding speed at 1% water contamination on account of the transition from a
boundary to a mixed lubrication regime.

For the specific applied load of 8 N, the low and high sliding speeds, and the water
contamination explored in this experimental study, damping has a stronger effect on
the vibrationwear relationship which depends on sliding speed and, as a result, on
the lubrication regime.

The tribological characteristics of water-lubricated bearings materials depend on


factors such as applied load, sliding speed, bearing design, and environmental
conditions. Therefore, the value of these POD test methods is to predict the relative
ranking of water-lubricated bearings materials combinations during the design,
development, or selection process and not to simulate all in-service waterlubricated bearings.

However, the analysis, the experimental results, and the experimental approach
presented here can be used for further extensive experimental and theoretical
studies for design simulation, and for the further development of a theoretical
model for the adequate prediction of the physical implications of a variation in the
operational conditions, material properties, and technical requirements of material
combinations used for water-lubricated engineering applications.

148

CHAPTER 10 Conclusions

10.3 Recommendations for future work


A proper understanding of the tribological behaviour of water-lubricated contact surfaces
for the design and development of new water-lubricated bearings is important. It is well
known that the effect of water contamination on the performance of water-lubricated
bearings is a complex mechanism and that it is extremely difficult to model. The results
from this experimental investigation allow certain recommendations to be made that will
be helpful in establishing future theoretical and experimental work to develop further
design approaches and theoretical models for reliable and quiet water-lubricated bearings.

Further progress in the theoretical and experimental study of friction in water-lubricated


bearings materials and systems will require the following:

Experimental and theoretical investigation of the effect of water contamination on


friction-induced vibration and wear. A set of experiments needs to be conducted to
further investigate the effect of water contamination on friction-induced vibrations.
Sources and frequency characteristics of noise need to be identified and analysed.

Further developments of theoretical nonlinear models of friction, vibration, and


wear are required to predict the effect of contamination on material performance in
water-lubricated bearings.

A new theoretical three-body model of friction, vibration, and wear using contact
mechanics is necessary for the analysis and design of water-lubricated bearings
under water contamination. It is still not known if any nonlinearity is due to the
effect of water contamination.

Proper material selection for experimental work. For example, all materials need to
be either properly analysed and/or all material property specifications should be
available in detail or determined.

Operational conditions, including the water lubrication regimes and surface


roughness, should be fully identified.

149

References

REFERENCES
AIPHAISS. 2009. Tribometers, Standard Tribometers (TRB). Alpha Instrument Supplies
and Servises Sdn BhD (AIphaISS). Available:
http://www.alphaiss.com.my/surface.html.
ANDREAUS, U. & CASINI, P. 2001. Forced Motion of Friction Oscillators Limited by a
Rigid or Deformable Obstacle. Mechanics of Structures and Machines, 29,
177-198.
ARONOV, V., D'SOUZA, A. F., KALPAKJIAN, S. & SHAREEF, I. 1983. Experimental
Investigation of the Effect of System Rigidity on Wear and Friction-Induced
Vibrations. Journal of Lubrication Technology, 105, 206-211.
ASTM 2003. E 4 03: Standard Practices for Force Verification of Testing Machines.
ASTM International.
ASTM 2004a. G 99 04: Standard Test Method for Wear Testing with a Pin-on-Disk
Apparatus. ASTM International.
ASTM 2004b. G 115 04: Standard Guide for Measuring and Reporting Friction
Coefficients. ASTM International.
ASTM 2006. E 74 06: Standard Practice of Calibration of Force-Measuring Instruments
for Verifying the Force Indication of Testing Machines. ASTM International.
ASTM G 99 04. Standard Test Method for Wear Testing with a Pin-on-Disk Apparatus.
ASTM International.
BARWELL, F. T. 1984. Advances in friction and wear mechanisms. Tribology
International.
BAYER, R. G. 2002. Wear Analysis for Engineers, New York, HNB Publishing.
BHATTACHARJEE, R. 2013. Free Vibration of a Cantilever Beam (Continuous System)
BHUSHAN, B. 1980. Stick-Slip Induced Noise Generation in Water-Lubricated Compliant
Rubber Bearings. Journal of Lubrication Technology, 102, 201-211
BISWELL, T. 2007. Bearings for longer shaft life. WRTSIL Technical Journal.
BRYANT, M. D. & YORK, D. 2000. Measurements and Correlations of Slider Vibrations
and Wear. Journal of Tribology, 122, 374-380.
CHOWDHURY, M. A. & HELALI, M. M. 2007. The effect of frequency of vibration and
humidity on the wear rate. Wear, 262, 198-203.
CHOWDHURY, M. A. & HELALI, M. M. 2009. The wear behavior of mild steel under
vertical vibration. Industrial Lubrication and Tribology, 61.
150

References

COMMITTEE, B. E. S. 1994. Plane bearings. ESC Report. BSA Education Service


Committee.
COURTEL, M. R. & TICHVINSKY, L. M. 1964. A brief history of friction. Naval
Engineers Journal, 76, 451-460.
CSM-INSTRUMENTS 2013. CSM Trobometers, Nano and Micro range for Tribology
studies. In: SA, C. I. (ed.) Rue de la Gare 4 Galileo Center CH-2034 Peseux
(Switzerland): Rue de la Gare 4 Galileo Center CH-2034 Peseux (Switzerland).
CUMBERLIDGE, P. 2009. Water lubricated bearings and seals in abrasive water
conditions. Wartsila Technical Journal, 01, 54-57.
DAUTZENBERG, J. H. 1986. A Model for Vibrations During Dry Sliding Friction. CIRP
Annals - Manufacturing Technology.
DULIAS, U. 2002. Effect of surface roughness of self-mated aluminia on friction and wear
in isooctane-lubricated reciprocating sliding contact. Wear.
DWEIB, A. H. & D'SOUZA, A. F. 1990. Self-excited vibrations induced by dry friction,
part 1: experimental study, part 2: stability and limit cycles. Journal of Sound and
Vibration, 137, 163-190.
ENGINEERING-ABC.COM. History of Science Friction. Available:
http://www.tribology-abc.com/abc/history.htm.
EOV 2004. Friction - Indused Vibrations. Encyclopedia of Vibrations (EOV).
GERE, J. M. 2002. Mechanics of Materials, Nelson Thornes Ltd. .
GINZBURG, B. M., TOCHIL'NIKOV, D. G., BAKHAREVA, V. E., V., A. A. &
KIREENKO, O. F. 2006. Polymeric Materials for Water-Lubricated Plain
Bearings. Russian Journal of Applied Chemistry, 79, 695-706.
GODFREY, D. 1995. Friction oscillations with a pin-on-disc tribometer. Tribology
International 28, 119-126.
GOTO, H. & ASHIDA, M. 1984. Effect of ultrasonic vibrations on the wear characteristics
of a carbon steel: analisys of wear mechanism. Wear, 94, 13-27.
HARRISON, J. A. 2008. Friction between solids. Philosophical Transactions of the Royal
Society. A Mathematical Physical and Engineering Sciences.
HE, L. 2010. Dynamical Adaptive Backstepping-Sliding Mode Control of Pneumatic
Actuator. University of Manitoba.
HERSEY., M. D. 1934. Notes on the history of lubrication. PART II. Journal bearings
experiments. Journal of the American Society for Naval Engineers, 46, 369-385.
HIRANI, H., RAO, T. V. V. L. N., ATHRE, K. & BISWAS, S. 1997. Rapid performance
evaluation of journal bearings. Tribology International, 3, 825834.
151

References

HOLINSKI, R. 2001. Fundamentals of dry friction and some practical examples. Industrial
Lubrication and Tribology, 53, 61-65.
HORI, Y. 2006. "Friction, Wear, and Lubrication", Hydrodynamic Lubrication, Tokyo,
Springer-Vedrlag.
IBRAHIM, R. A. 1994a. Friction-induced vibration, chatter, squeal, and chaos. Part 1:
Mechanics of contact and friction. Applied Mechanics Reviews, 47, 209-226.
IBRAHIM, R. A. 1994b. Friction-induced vibration, chatter, squeal, and chaos. Part 2:
Dynamics and modeling. Applied Mechanics Reviews, 47, 227-240.
IVANOV, M. G. & IVANOV, D. M. 2012. Nanodiamond Nanoparticles as Additives to
Lubricants. Ultrananocrystalline Diamond.
JONSON, M. S. 2000. Vibration tests for bearing wear. ASHRAE Journal
KAARSTAD, E. 2009. A Study of Temperature Dependant Friction in Wellbore Fluids.
SPE/IADC Drilling Conference and Exibition. DC.
KASADRA, M. E. F., MENDOZA, H., KIRK, R. G. & WICK, A. 2004. Reduction of
subsynchronous vibrations in a single-disk rotor using an active magnetic damper.
Mechanical Research Communications, 31, 689-695.
KASARDA, M. E. F. 2004. Reduction of subsynchronous vibrations in a single-disk rotor
using an active magnetic damper. Mechanics Research Communications, 11.
KATO, K., IWABUCHI, A. & KAYABA, T. 1982. The effect of friction-induced
vibration on friction and wear. Wear, 80, 307-320.
KHARAGPUR, I. Brief overview of bearings. Module 14. Version 2 ME ed.
KINGSBURY, I. 1997. A General Guide to the Principles, Operation and Troubleshooting
of Hydrodynamic Bearings. Philadelphia.
KOTOUSOV, A. 2009. Aft Bearing Study Engineering Report (Final). Adelaide:
University of Adelaide.
KRIM, J. 2002. Surface science and the atomic-scale origins of friction: what once was old
is new again. Surface Science, 500, 741-758.
KRISHNA KUMAR, V. & SWARNAMANI, S. 1997. Vibration monitoring in sliding
wear of plasma sprayed ceramics. Wear, 210, 255-262.
LEDOCQ, J. H. M. 1973. A Test Rig for Journal Bearings results with polymer bushes
Wear, 133-142.
LIANG, H. 2005. "Friction", Tribology in Chemical-Mechanical Planarization.
LING WU, LING-FENG HE, JI-XIN CHEN, XIN-PO LU & ZHOU, Y.-C. 2010.
Reciprocating Friction and Wear Behaviour of Zr2[Al(Si)]4C5-SiC Composite
Against Si3N4 Ball. Journal of the American Ceramic Society, 93.
152

References

LITWIN, W. 2009. Water-lubricated bearings of ship propeller shafts - problems,


experimental tests and theoretical investigations. Polish Maritime Research, 16, 4255.
MARINEDIESELS. 2011. Operational Information. Hydrodynamic Lubrication.
marinediesels.co.uk. Available:
http://www.marinediesels.info/2_stroke_engine_parts/Other_info/hydrodynamic_lu
brication.htm.
MARKLUND, P. & LARSSON, R. 2008. Wet clutch friction characteristics obtained from
simplified pin on dick test. Tribology International, 824-830.
MARTENS, A. 1888-1889. Schmierluntersuchungen (Investigations on oils). In:
SPRINGER, V. V. J. (ed.) Mitteilungen aus den Kniglichen technischen
Versuchsanstalten zu Berlin. Berlin: Ergnzungsheft.
MARTINS, J. A. C. 1990. A Study of Static and Kinetic Friction. International Journal of
Engineering Science.
MARU, M. M., CASTILLO, R. S. & PADOVESE, L. R. 2007a. Study of solid
contamination in ball bearings through vibration and wear analysis. Tribology
Iternational, 40, 433-440.
MARU, M. M., SERATTO-CASTILLO, R. S. & PADOVESE, L. R. 2005. Detection of
solid contamination in rolling bearing operation through mechanical signature
analysis. 12 International congress on sound and vibration. Lisbon, Portugal.
MARU, M. M., SERRATO-CASTILLO, R. & PADOVESE, L. 2007b. Influence of oil
contamination on vibration and wear in ball and roller bearings. Industrial
Lubrication and Tribology, 59, 137-142.
MATERIALS TRIBOLOGY LABORATORY, U. O. D. 2008. Overview of Polymer
Nanocomposite Tribology. Available:
http://research.me.udel.edu/~dlburris/research.html.
MEUTER, P. 2006. Protecting pumps against abrasive wear. World Pamps, 2006, 18-20.
MOKHIAMER, U. M., CROSBY, W. A. & EL-GAMAL, H. A. 1999. A study of a journal
bearing lubricated by fluids with couple stress considering the elasticity of the liner.
Wear, 194201.
MOKHTAR, M. O. A., ALY, W. Y. & SHAWKI, G. S. A. 1984. Experimental study of
journal bearings with undulating journal surface. Tribology International, 17, 1923.
MOORE, M. A. 1978. Abrasive wear. Materials in Engineering applications, 1, 97-111.
153

References

MORIWAKI, T. & SHAMOTO, E. 1991. Ultrapresision diamond turning of stainless steel


by applying ultrasonic vibrations. Annals of the CIRP, 40, 559-582.
MOSLEH, M., LAUBE, S. J. P. & SUH, N. 2002. A mechanism of high friction in dry
sliding bearings. Wear, 252, 1-8.
MOTTERSHEAD, J. E., OUYANG, H. & CARTMELL, M. P. 1997. Parametric
resonances in an annular disc, with a rotating system of distributed mass and
elasticity; and the effects of friction and damping. Proceeding of the Royal Society
London, 453, 1-19.
NEVEU, C. D., SONDJAJA, R., STAUHR, T. & IROFF, N. J. 2012. Lubricant and Fuel
Additives Based on Polyalkylmethacrylates. Polymer Science.
NIKOLAKOPOULOS, P. G. & PAPADOPOULOS, C. A. 2008. A study of friction in
worn misaligned journal bearings under severe hydrodynamic lubrication.
Tribology International, 41, 461-472.
NOUIRA, H. 2008. Experimental characterization and modelling of microsliding on a
small cantilever quartz beam. Journal of Sound and Vibration.
OLSSON, H., ASTROM, K. J., CAUDAS DE WIT, C., GAFVERT, M. & LISCHINSKY,
P. 1997. Friction Models and Friction Compensation. Department of automatic
Control, Lund Institute of Technology, Lund University; Laboratoire
d'Automatique de Grenoble; Control Department EIS, ULA
PADMANABHAN, K. 1992. Effect of lubrication on slip damping. Wear, 154, 65-76.
PALACI, I. 2007. Atomic force microscopy studies of nanotribology and nanomechanics.
DE DOCTEUR S SCIENCES NO 3905 (2007), cole Polytechnique Federale De
Lausanne.
PAN, C. H. T., LUND, J. W., GU, A. & WALOWIT, J. A. 1971. Stick-slip excitation of
propeller-shaft torsional vibration as a sourse of bearing squeal Mechanical
Tribology Inc.
PILIPCHUK, V. N. 2002. Disc Brake Ring-Element Modelling Involving Friction-Induced
Vibration. Journal of Vibration and Control, 08/01.
PLATT, N., HAMMEL, S. M. & HEAGY, J. F. 1994. Effects of Additive Noise on OnOff Intermittency. Physical Review Letters, 72, 3498-3501.
PML. 2009. Tribology, Basic Technical Information, Prime Master LubTech (PLM Ltd.)
Limited. Available: http://primelubtech.com/Techs.aspx.
POLAN, N. W., HEINE, M. A., POPPLEWELL, J. M. & GAFFOGLIO, C. J. 1981.
Erossion-Corrosion resistance of Copper Alloy C72200 in Sea Water Containing
Suspended Sand. Desalination, 38, 223-231.
154

References

POPOV, V. L. 2010. Frictionally Induced Vibration. Contact Mechanics and Friction.


PREHN, R., HAUPERT, F. & FRIEDRICH, K. 2005. Sliding wear performance of
polymer composites under abrasive and water lubricated conditions for pump
applications. Wear, 259, 693-696.
QIANG, L. 2009. Friction in Highly Loaded Mixed Lubricated Point Contacts. Tribology
Transactions.
QIAO, S. L. & IBRAHIM, R. A. 1999. Stochastic Dynamics of Systems with FrictionInduced Vibration. Journal of Sound and Vibration, 223, 115-140.
RAC, A. & VENCL, A. 2005. Tribological and Design Parameters of Lubricated Sliding
Bearings. Tribology in Industry, 27, 12-16.
READ, L. J. & FLACK, R. D. 1987. Temperature, Pressure and film Thickness
Measurements for an Offset Half Bearings. Wear, 197 - 210.
RUINA, A. & PRATAP, R. 2002. Introduction to Statics and Dynamics. Oxford
University Press,.
RUINA, A. & PRATAP, R. (2002. Introduction to Statics and Dynamics. Oxford
University Press,.
RYADCHENKO, G. V. 2003. Tribogical Characteristics of Elasomers Modified by
Antifrictional Fibers( ,
). PhD,
Don State Technical University.
SCHALLAMACH, A. 1963. A Theory of Dynamic Rubber Friction. Wear, 6, 375-382.
SCHALLAMACH, A. 1971. How Does Rubber Slide. Wear, 17, 301-312.
SIMPSON, T. A. & IBRAHIM, R. A. 1996. Nonlinear Friction-Induced Vibration in
Water-Lubricated Bearings. Journal of Vibration and Control, 2, 87-113.
SMITH, W. V. 1973. Material Selection Criteria for water Lubrication. Wear, 25.
SOLOMONOV, Y. 2009. Theoretical and Experimental Study of Friction Phenomena in
Water Lubricated Bearings. Research Proposals. Adelaide: University of Adelaide.
SOLOMONOV, Y., BROWN, I., HARDING, S. & KOTOUSOV, A. 2010. Experimental
apparatus and preliminary investigations of friction and wear phenomena in waterlubricated bearings 6th Australasian Congress on Applied Mechanics, ACAM 6 1215 December 2010, Perth, Australia.
SONG, H. J. 2008. Tribological behaviour of polyurethane-based composite coating
reinforced with TiO2 nanotubes. Europian Polymer Journal, 04.

155

References

STLE. 2008. Society of Tribologists and Lubrication Engineers, Basics of Lubrication.


Society of Tribologists and Lubrication Engineers. Available:
http://www.stle.org/resources/lubelearn/lubrication/default.aspx.
STRIBECK, R. 1902a. Die wesentlichen Eigenschaften der Gleit- und Rollenlager
(Characteristics of Plain and Roller Bearings). Zeit, des VDI 46.
STRIBECK, R. 1902b. The key qualities of sliding and roller bearings. Zeitschrift des
Vereines Seutscher Ingeniere, 46, 1342-1348.
SUDA, Y. & KOMINE, H. 1996. Contact vibration with high damping alloy. Wear, 191,
72-77.
SUNDARARAJAN, C. 2009. Compendium of Formulas for the Structural Vibration
Frequency Analysis of Beams. PDHengineer.com. Available:
http://www.pdhengineer.com/courses/s/S-3001.pdf.
SUTHERLAND, J. W. 2002. Vibration Damping, The John W. Sutherland Research Page,
Michigan Technological University. Available:
http://www.mfg.mtu.edu/cyberman/machtool/machtool/vibration/damping.html.
TEIDELT, E. 2012. Influence of Ultrasonic Oscilations on Static and Sliding Friction.
Tribology Letters.
THORDON 2006. Engineering Manual Version E2006.1. In: INC., T. B. (ed.). Burlington:
THORDON.
THURSTON, R. H. 1879. Friction and lubrication - Determination of the laws and
coefficients of friction by new methods and with new apparatus. In: CO, T. A.
(ed.). Ludgate Hill, London.
THURSTON, R. H. 1894. A treatise on friction and lost work in machinery and millwork,
New York.
TORRANCE, A. A. 2005. Modelling abrasive wear. Wear, 258, 281-293.
TWORZYDLO, W. W., BECKER, E. B. & ODEN, J. T. 1994. Numerical modeling of
friction-induced vibrations and dynamic instabilities. Applied Mechanics Reviews,
47, 225-274.
TWORZYDLO, W. W., HAMZEH, O. N., ZATON, W. & JUDEK, T. J. 1999. Frictioninduced oscillations of a pin-on-disk slider: analytical and experimental studies.
Wear, 236, 9-23.
UNAL, H., MIMAROGLU, A., KADIOGLU, U. & E. EKIZ, E. 2004. Sliding friction and
wear behaviour of polytetrafluoroethylene and its composites under dry conditions.
Materials and Design, 239-245.
UOS 2008. Life Science Notes: PM6 Friction. Lecture Notes. The University of Sydney.
156

References

VERICHEV, N., VERICHEV, S. & EROFEYEV, V. 2010. Damping lateral vibrations in


rotary machinery using motor speed modulation. Journal of Sound and Vibration,
329, 13-20.
VERICHEV, N. N. 2010. Damping lateral vibrations in rotary machinery using motor
speed modulation. Journal of Sound and Vibration, 01.
WRTSIL 6/09/2007. Water Lubricated Sturntube Bearings. Design and Procedures
Manual, RSD21.
WEBER, H., HERBERG, J., PILZ, R. & KARL-MARX-STADT, T. H. 1984. Turning of
machinable glass ceramics with an ultrasonic vibrated tool. Annals of the CIRP,
33, 85-87.
WEICHSEL, D. 1994. Plane Bearings. The Educational Services Committee
WERKTUIGBOUW&TRIBOLOGIE. 2010. History of tribology. Available:
http://www.tribologie.nl/backgrounds/history/history.htm.
YAMAJO, S. & KIKKAWA, F. 2004. Development and Application of PTFE Compound
Bearings. In: SESSION, D. D. (ed.) Dynamic Positioning Conference Dynamic
Positioning Sosiety.
YOUNES, Y. K. 1993. On the dynamics and stability of a semi-dry journal bearing. Wear,
169, 215-220.
ZHURAVLEV, V. F. 2010. History of the Law of Dry friction. Doclady Physics, 55, 344345.
FEODOSIEV, V. I. 1974. Strength of Materials ( ), Moscow,
Science, (, ).

157

APPENDIX A: Pin-on-Disk assembly drawings

APPENDIX A: PIN-ON-DISK ASSEMBLY DRAWINGS

(This page intentionally left blank)

158

159

160

También podría gustarte