You are on page 1of 242

DOC TOR A L T H E S I S

Andrew Spencer A Simulation Tool for Optimising Combustion Engine Cylinder Liner Surface Texture
Department of Engineering Sciences and Mathematics
Division of Machine Elements

A Simulation Tool for


ISSN: 1402-1544
ISBN 978-91-7439-789-5 (print) Optimising Combustion Engine
ISBN 978-91-7439-790-1 (pdf)

Lule University of Technology 2013


Cylinder Liner Surface Texture

Andrew Spencer
A Simulation Tool for
Optimising Combustion Engine
Cylinder Liner Surface Texture

Andrew Spencer

Lule
a University of Technology
Department of Engineering Sciences and Mathematics
Division of Machine Elements
Cover figure: Scania R 730 6x4 Streamline,
Ljusdal, Sweden.

Title page figure: Scania 730 hp 16-litre V8,


Euro V and EEV engine.

A Simulation Tool for


Optimising Combustion Engine
Cylinder Liner Surface Texture
Copyright Andrew Spencer (2013). This document is freely available at

http://www.ltu.se

The document may be freely distributed in its original form including the
current authors name. None of the content may be changed or excluded
without the permission of the author.

Printed by Lule University of Technology, Graphic Production 2013

ISSN: 1402-1544
ISBN 978-91-7439-789-5 (print)
ISBN 978-91-7439-790-1 (pdf)

Lule 2013

www.ltu.se

This document was typeset in LATEX 2


.
Preface

This thesis is the culmination of four years of work at the Division of Ma-
chine Elements at Lule a University of Technology. My research would not
have been possible without funding from Stiftelsen for Strategisk Forskning
(SSF), ProViking, and Vinnova through the Swedish Research School in
Tribology.
During the last four years I have had two extremely fruitful collabora-
tions which have contributed greatly to this thesis. I would like to thank Illia
Dobryden from the Division of Physics at LTU for his collabration which led
to Paper C in this thesis. Also Emin Yusuf Avan, for his collaboration at
the Leonardo Centre for Tribology at the University of Sheffield, work that
led to the publication of Paper D and E in this thesis. I am also grateful to
Professor Rob Dwyer-Joyce for allowing me to spend time at the Leonardo
Centre and supporting my research while I was there.
I would especially like to express my graditude to my two supervisors at
LTU, Professor Roland Larsson and Associate Professor Andreas Almqvist
for their help, support, guidance and many valuable discussions over the
course of this research. I would also like to thank all of my friends at the
Division of Machine Elements for making Lule a such a warm and enjoyable
place to work.
At Scania I would like to thank Peter Eriksson for helpfully answering
all of my many questions about piston rings and cylinder liners. Dr. Hubert
Herbsts contribution has been invaluable, guiding me through my research
and helping me to overcome many of the challenges that I have encountered
along the way. I thank Peter Daelander for giving me the opportunity to
join his group at Scania and for the support and time I needed to complete
this thesis.
I would also like to thank my parents for their support and encourage-
ment as I moved far away from home to continue my studies.
Finally, my biggest thanks goes to Elise, for her never-ending under-
standing and patience, and for always being there when I need her.

Andrew Spencer, S
odert
alje, November 2013

i
Abstract

Fuel efficiency is one of the most important areas of automotive vehicle re-
search and development today, with rising fuel costs, energy security and
environmental concerns being at the forefront of customers and legislators
minds. Heavy Duty Diesel Engines (HDDE) are the primary source of me-
chanical power generation in todays trucks and buses and this is likely
to continue for the foreseeable future. In the 2011 European Commission
White Paper on transport, a reduction of at least 60% of greenhouse gas
emissions from transport by 2050, with respect to 1990 levels, was called
for. The report concludes that acting on vehicles efficiency through new
engines, materials and design will help in the reduction of oil dependence,
the competitiveness of Europes automotive industry as well as health ben-
efits, especially improved air quality in cities. Therefore, the efficiency and
frictional losses in a vehicles powertrain are areas of great interest.
This thesis focuses on the Piston Ring to Cylinder Liner (PRCL) contact
and the potential for improving its performance through the specification
of an optimised cylinder liner surface texture. The PRCL contact is one
of the biggest contributors to mechanical losses in a HDDE and so there is
potential for large performance gains to be achieved through optimisation
of this contact.
This research has led to the development of a simulation tool capable of
calculating the friction, lubrication regime, oil consumption risk and wear
that occurs in the full ring-pack of a HDDE. Furthermore, the tool allows for
the evaluation of the relative performance of different cylinder liner surface
topographies. A mixed lubrication model, incorporating flow factors calcu-
lated using the homogenization technique, has been implemented to allow
all regimes of lubrication to be considered. A mass-conserving cavitation
algorithm, formulated as a Linear Complimentarity Problem, enables lubri-
cant cavitation, fully-flooded or starved inlet conditions and the quantity of
lubricant deposited on the cylinder liner surface to be modelled.
The simulation tool is validated with both reciprocating bench tests
and full single cylinder fired engine tests. The reciprocating bench tests
measured both friction and film thickness and both showed good correlation

iii
with the predictions from the simulation tool. Simulations and experiments
were conducted on four different cylinder liner variants and both ranked the
frictional performance of the cylinder liner variants in the same order.
A parametric study of honing depth, spacing and angle was undertaken
using the developed simulation tool and the influence of these parameters
on lubricant film thickness, friction, wear and oil consumption was investi-
gated. The thesis concludes that a reduction in specific fuel consumption is
achieveable through the optimisation of cylinder liner texture and outlines
how this might be achieved.
Contents

Nomenclature 1

I The Thesis 5

1 Introduction 7
1.1 Frictional losses in on-road vehicles and the need for improve-
ment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 The Piston Ring-Cylinder Liner Contact . . . . . . . . . . . . 9
1.2.1 The Piston Ring Pack . . . . . . . . . . . . . . . . . . 9
1.2.2 The Cylinder Liner . . . . . . . . . . . . . . . . . . . . 10
1.2.3 The Tribology of the PRCL contact . . . . . . . . . . 11
1.3 Optimizing the PRCL contact . . . . . . . . . . . . . . . . . . 12
1.4 State-of-the-art in PRCL simulation . . . . . . . . . . . . . . 13
1.5 Objectives of this thesis . . . . . . . . . . . . . . . . . . . . . 15
1.5.1 Implementation of an all-regime model incorporating
the effect of surface texture . . . . . . . . . . . . . . . 16
1.5.2 Validation of the simulation tool . . . . . . . . . . . . 16
1.5.3 Working towards optimum texture . . . . . . . . . . 16

2 Methods and Tools 17


2.1 The treatment of texture in a simulation tool . . . . . . . . . 17
2.2 The Reynolds equation and Cavitation . . . . . . . . . . . . . 19
2.3 Flow Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.4 Full film viscous friction . . . . . . . . . . . . . . . . . . . . . 30
2.5 A Mixed Lubrication Model . . . . . . . . . . . . . . . . . . . 30
2.6 Surface topography measurement . . . . . . . . . . . . . . . . 33
2.6.1 Vertical Scanning Interferometry . . . . . . . . . . . . 34
2.6.2 Atomic Force Microscopy . . . . . . . . . . . . . . . . 34
2.6.3 Confocal Microscopy . . . . . . . . . . . . . . . . . . . 35
2.6.4 Comparison of the different techniques . . . . . . . . . 36

v
3 Development of a PRCL simulation tool 39
3.1 Global overview of the problem . . . . . . . . . . . . . . . . . 39
3.2 Boundary conditions on the piston ring . . . . . . . . . . . . 45
3.3 Surface texture and flow factors . . . . . . . . . . . . . . . . . 47
3.4 Friction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.5 Wear . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.6 General solution procedure . . . . . . . . . . . . . . . . . . . 50
3.7 Model input parameters and convergence . . . . . . . . . . . 51

4 Experimental validation of the model 53


4.1 Surfaces under investigation . . . . . . . . . . . . . . . . . . . 53
4.2 Single cylinder engine tests . . . . . . . . . . . . . . . . . . . 54
4.2.1 Measurement of FMEP in a fired engine . . . . . . . . 55
4.2.2 Reciprocating bench tests . . . . . . . . . . . . . . . . 58

5 Optimization of texture 65
5.1 Parametric study of texture parameters . . . . . . . . . . . . 65
5.2 Evaluation of results . . . . . . . . . . . . . . . . . . . . . . . 69
5.2.1 The effect of honing angle . . . . . . . . . . . . . . . . 82
5.2.2 The effect of honing depth and spacing . . . . . . . . 85
5.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

6 Conclusions 91

7 Future work 93

II Appended Papers 95

A 97
A.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
A.2 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
A.2.1 Surface texture . . . . . . . . . . . . . . . . . . . . . . 101
A.2.2 Cavitation algorithm . . . . . . . . . . . . . . . . . . . 104
A.2.3 Boundary conditions . . . . . . . . . . . . . . . . . . . 106
A.2.4 Film thickness . . . . . . . . . . . . . . . . . . . . . . 106
A.2.5 Force balance and time dependence . . . . . . . . . . . 107
A.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
A.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
A.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
A.6 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . 110
B 111
B.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
B.2 Model Development . . . . . . . . . . . . . . . . . . . . . . . 116
B.2.1 Geometry and global problem . . . . . . . . . . . . . . 116
B.2.2 Global surface texture . . . . . . . . . . . . . . . . . . 117
B.2.3 Reynolds equation . . . . . . . . . . . . . . . . . . . . 119
B.2.4 Flow factors . . . . . . . . . . . . . . . . . . . . . . . . 120
B.2.5 Model input parameters . . . . . . . . . . . . . . . . . 121
B.2.6 Force Balance and time dependence . . . . . . . . . . 123
B.2.7 Friction . . . . . . . . . . . . . . . . . . . . . . . . . . 124
B.2.8 Grid size and convergence . . . . . . . . . . . . . . . . 124
B.3 Texture Investigation . . . . . . . . . . . . . . . . . . . . . . . 125
B.4 Model Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
B.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
B.6 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . 129

C 131
C.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
C.2 Surfaces under investigation . . . . . . . . . . . . . . . . . . . 136
C.3 Measurement Techniques . . . . . . . . . . . . . . . . . . . . 137
C.3.1 VSI Measurements . . . . . . . . . . . . . . . . . . . . 137
C.3.2 AFM Measurements . . . . . . . . . . . . . . . . . . . 137
C.4 Tribological Surface Parameters . . . . . . . . . . . . . . . . . 138
C.4.1 Rk parameters . . . . . . . . . . . . . . . . . . . . . . 138
C.4.2 Flow Factors and Contact Stiffness . . . . . . . . . . . 139
C.5 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . 142
C.5.1 Rk Parameters . . . . . . . . . . . . . . . . . . . . . . 145
C.5.2 Flow Factors and Contact Stiffness . . . . . . . . . . . 146
C.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
C.7 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . 153

D 155
D.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
D.2 Test Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . 160
D.2.1 Piston Ring-Liner Simulator . . . . . . . . . . . . . . . 160
D.2.2 Ultrasonic Sensors . . . . . . . . . . . . . . . . . . . . 164
D.2.3 Instrumentation . . . . . . . . . . . . . . . . . . . . . 165
D.2.4 Lubricant . . . . . . . . . . . . . . . . . . . . . . . . . 165
D.3 Ultrasonic Oil Film Measurement . . . . . . . . . . . . . . . . 166
D.3.1 Background . . . . . . . . . . . . . . . . . . . . . . . . 166
D.3.2 Data Capturing and Analysis . . . . . . . . . . . . . . 167
D.4 Numerical Model . . . . . . . . . . . . . . . . . . . . . . . . . 170
D.5 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . 172
D.5.1 Measured Film Thickness . . . . . . . . . . . . . . . . 173
D.5.2 Comparisons between Experimental and Numerical
results . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
D.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
D.7 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . 181
D.8 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

E 185
E.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
E.2 Experimental setup . . . . . . . . . . . . . . . . . . . . . . . . 190
E.2.1 Test setup, specimens and lubricant . . . . . . . . . . 190
E.2.2 Test conditions . . . . . . . . . . . . . . . . . . . . . . 192
E.2.3 Ultrasonic film thickness measurement . . . . . . . . . 193
E.3 Numerical model . . . . . . . . . . . . . . . . . . . . . . . . . 195
E.3.1 Lubricant properties . . . . . . . . . . . . . . . . . . . 198
E.4 Surfaces under investigation . . . . . . . . . . . . . . . . . . . 198
E.4.1 Calculation of flow factors and asperity contact pres-
sures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
E.5 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . 202
E.5.1 Friction . . . . . . . . . . . . . . . . . . . . . . . . . . 202
E.5.2 Film thickness . . . . . . . . . . . . . . . . . . . . . . 205
E.5.3 Time resolved data analysis . . . . . . . . . . . . . . . 208
E.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
E.7 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . 213
E.8 Funding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
List of Appended Papers

Paper A
A numerical model to investigate the effect of honing angle on the
hydrodynamic lubrication between a combustion engine piston ring
and cylinder liner
A. Spencer, A. Almqvist, R. Larsson
(Proceedings of the Institution of Mechanical Engineers, Part J: Journal of
Engineering Tribology, July 2011, vol. 225 no. 7, pp. 683-689 )
All simulation development and the writing of Paper A was accomplished by
the author, with discussions and guidance from A. Almqvist and R. Larsson.

Paper B
A semi-deterministic texture-roughness model of the piston ring -
cylinder liner contact
A. Spencer, A. Almqvist, R. Larsson
(Proceedings of the Institution of Mechanical Engineers, Part J: Journal of
Engineering Tribology, June 2011, vol. 225 no. 6, pp. 325-333 )
All simulation development and the writing of Paper B was accomplished
by the author with discussions and guidance from A. Almqvist and R. Lars-
son. The flow factors were calculated using pre-existing code collaboratively
developed at the Division of Machine Elements.

Paper C
The influence of AFM and VSI techniques on the accurate calculation
of tribological surface roughness parameters
A. Spencer, I. Dobryden, N. Almqvist, A. Almqvist, R. Larsson
(Tribology International, January 2013, vol. 57, pp. 242-250 )

ix
The author conducted the VSI surface measurements while I. Dobryden
carried out the AFM measurements. The author calculated the flow factors
and roughness parameters and then analysis and conclusions were drawn
together by the author and I. Dobryden. The main writer of Paper C was
the author but with significiant input from I. Dobryden. N. Almqvist, A.
Almqvist and R. Larsson contributed with discussions and guidance.

Paper D

Experimental and numerical investigations of oil film formation and


friction in a piston ringliner contact
E. Y. Avan, A. Spencer, R. S. Dwyer-Joyce, A. Almqvist, R. Larsson
(Proceedings of the Institution of Mechanical Engineers, Part J: Journal of
Engineering Tribology, February 2013, vol. 227 no. 2, pp. 126-140 )

For the authors contribution, see Paper E.

Paper E

An experimental and numerical investigation of frictional losses and


film thickness for four cylinder liner variants for a heavy duty diesel
engine
A. Spencer, E. Y. Avan, A. Almqvist, R. S. Dwyer-Joyce, R. Larsson
(Proceedings of the Institution of Mechanical Engineers, Part J: Journal of
Engineering Tribology, December 2013, vol. 227 no. 12, pp. 1319-1333 )

Paper D and E were a close collaboration between E. Y. Avan and the


author. The experiments, run at the Leonardo Centre for Tribology, were
designed, developed and run by both the author and E.Y. Avan. Following
this, E.Y. Avan processed the ultrasonic data (his area of expertise) and the
author ran the simulations of the experiment. E.Y. Avan took the lead in
writing Paper D while the author took the lead in writing Paper E, but both
papers contain significant contributions from both authors. R. S. Dwyer-
Joyce provided guidance on the experiments and discussion of results while
A. Almqvist and R. Larsson contributed with discussions on the simulation
results.
Nomenclature

Honing angle ( )

h Average separation (m)


T
h Local scale separation between surfaces (m)

Compressibility factor (Pa)

Reynolds temperature-viscosity coefficient (-)

i Local scale solution variable

m
in Unit mass flow immediately prior to ring (kg/m s)

w Wear rate (nm/h)

Separation parameter (m)

/ Friction coefficient (-)

0 Dynamic viscosity at the reference temperature T0 (Pas)

12

6U

/ Dynamic viscosity (Pas)

Integration domain (m2 )

Angular frequency of ultrasonic wave (rad/s)

Crank angle ( )

Honing generation variable (-)

i Local scale solution variable

Lubricant density (kg/m3 )

1
c Lubricant cavitation density (kg/m3 )

Fluid film fraction (-)

Plint spindle angle ( )

A ASTM D341 lubricant constant

A Real area of contact (m2 )

A Wave amplitude (V)

a/k Conrod length (m)

A0 Poiseuille flow factor matrix (-)

ai Poiseuille flow factor (m3 )

B Bulk modulus (Pa)

B0 Couette flow factor matrix (-)

b12 Couette flow factor (m)

c Speed of sound in lubricant layer (m/s)

cii /dii Flow factors for homogenized viscous friction (-)

D Honing groove depth (m)

d Bore diameter (m)

f Frequency of ultrasound (Hz)

FT Ring tension (N)

Fbound /FCP Asperity contact load (N)

FHYD Hydrodynamic supported load (N)

FTENS Piston ring tension load (N)

fbd Boundary friction (N)

FGAS Gas pressure load (N)

fhyd Hydrodynamic friction (N)

ftot Total friction force (N)

g Switch function (-)


H Material hardness (Pa)

h Film thickness (m)

h0 Separation between piston ring and cylinder liner (m)

hr Local scale topography height (m)

hT Texture height (m)

hex Height of lubricant on liner at ring outlet (m)

hin Height of lubricant on liner at ring inlet (m)

hmin Minimum film thickness (m)

K Stiffness of the layer (N/m)

k Honing generation variable (-)

kw Archards wear constant (-)

l Ring width (m)

L/s Stroke length (m)

l1 , l2 Local scale length (m)

N Engine speed (rpm)

N Plint speed (Hz)

p Hydrodynamic pressure (Pa)

p0 Homogenized Reynolds film pressure (Pa)

pc Cavitation pressure (Pa)

ps Solution pressure (Pa)

Pb/BOT T OM Gas pressure below piston ring (Pa)

PCP Average asperity contact pressure (Pa)

PHY D Hydrodynamic film pressure (Pa)

Pin Gas pressure at contact inlet (Pa)

Pout Gas pressure at contact outlet (Pa)

Pt/T OP Gas pressure above piston ring (Pa)


R Piston ring radius (m)

R Reflection coefficient (-)

r Complimentarity variable (-)

Ra Average roughness parameter (m)

S Spacing of honing grooves (m)

T Engine torque (Nm)

t Time (s)

Tl Cylinder liner temperature under ring ( C)

U Entraining velocity (m/s)

Vd Engine displaced volume (m3 )

w Honing generation variable (-)

Wi Indicated work done (J)

x Space coordinate (m)

y Space coordinate (m)

y1 , y2 Local scale coordinate (m)

z Acoustic impedence (N/s/m3 )

2CR Lower Compression Ring

BDC Bottom Dead Centre

BMEP Brake Mean Effective Pressure (Pa)

CA Crank Angle

FMEP Friction Mean Effective Pressure (Pa)

IMEP Indicated Mean Effective Pressure (Pa)

OCR Oil Control Ring

TCR Upper Compression Ring

TDC Top Dead Centre


Part I

The Thesis

5
Chapter 1

Introduction

In this chapter the motivation for undertaking tribological research on the


piston ring to cylinder liner (PRCL) contact of an internal combustion (IC)
engine is explained. Following on from this, the features of the PRCL con-
tact are introduced, with a brief description of their purpose and typical con-
figuration. Next follows a discussion of the requirements for optimizing the
contact, and the tribology involved. The chapter concludes with a summary
of state-of-the-art in PRCL simulation and an outline of the objectives for
this thesis.

1.1 Frictional losses in on-road vehicles and the


need for improvement
Fuel efficiency is one of the most important areas of automotive vehicle
research and development today, with rising fuel costs, energy security and
environmental concerns being at the forefront of customers and legislators
minds.
Heavy Duty Diesel Engines (HDDE) are the primary source of mechani-
cal power generation in todays trucks and buses and this is likely to continue
for the foreseeable future. In 2010 the global demand for diesel fuel was 16
million barrels of oil per day, in 2040 this is predicted to be 26 million bar-
rels per day, an increase of 60% [1]. EU transport still depends on oil and
oil products for 96% of its energy needs [2]. Although larger trucks and an
improvement in truck efficiency will improve the overall efficiency of road
transport, growing worldwide GDP will increase the need for road trans-
portation and road congestion will increase fuel use further. In the 2011
European Commission White Paper on transport [2], a reduction of at least
60% of greenhouse gas emissions from transport by 2050, with respect to
1990 levels, was called for. The report concludes that acting on vehicles

7
8 CHAPTER 1. INTRODUCTION

efficiency through new engines, materials and design will help in the reduc-
tion of oil dependence, the competitiveness of Europes automotive industry
as well as health benefits, especially improved air quality in cities. How-
ever, it is predicted that a considerable extent of shipments over short and
medium distances, i.e. below 300 km, and that 50% of shipments over 300
km, will remain on trucks even in 2050. Also, there will be a greater push
in cities to encourage the use of public transport with large fleets of urban
buses. There is great interest to improve Brake Thermal Efficiency (BTE),
which consists of closed-loop efficiency, open-loop efficiency and mechanical
friction losses. Great efforts have been undertaken in all three domains, by
means of better combustion, reduction of losses through the cylinder wall
and cylinder head, reduction of pumping losses in air handling and finally
optimisation of friction in lubricated contacts. Therefore, the efficiency and
frictional losses in a vehicles powertrain, particularly those in trucks and
buses powered by a HDDE, are areas of great interest.
Figures for the percentage breakdown of thermodynamic and mechani-
cal losses in a vehicle have frequently been presented in literature over the
last few decades. However, there is often a huge variation in the numbers
given. This is not surprising, as the various studies focus on a wide range of
vehicles. Furthermore, depending on the drive cycle investigated, the break-
down of percentage losses can vary widely. Finally, the studies available in
literature have been published over many years and do not all represent the
same level of technological development, and none represent todays state-
of-the-art. Despite these limitations, this valuable literature can be used as
a useful guide to the losses that occur in todays vehicles and the potential
magnitude of future improvement.
Richardson [3] stated that only about 40% of the fuel energy consumed
in a HDDE is converted to mechanical power. When it comes to passenger
cars, for a medium size car over an urban cycle only 12% of the total power
from the fuel is converted to useful energy at the wheels [4]. The rest of
this fuel energy is lost to a combination of thermal and mechanical losses.
Although the thermal losses are the greatest component, the mechanical
friction losses are significant at between 4-15% of the total fuel energy [3],
however as stated by Heywood [5], this figure can vary from 10% at full load
to 100% at idle. Pinkus [6] put these mechanical losses for passenger cars
on the EPA cycle at 6%, while Andersson [4] put the total frictional losses
in a passenger car at 17%.
Of these mechanical losses, the piston ring pack is the greatest single
contributor, amounting to 1.1-6.8% of the total losses in a HDDE [3]. An-
dersson [4] suggested that in a passenger car the piston ring pack friction
amounted to 40-50% of the total mechanical losses during an urban cycle,
whereas Bolander et al. [7] put this figure at between 20-40%. Meanwhile,
1.2. THE PISTON RING-CYLINDER LINER CONTACT 9

Spearot [8] showed that the piston ring contribution to the entire friction
loss was 19% in a light duty vehicle. Even with the significiant variation in
the figures presented here, it is clear that the ring pack is one of the most
interesting areas for consideration when trying to reduce mechanical friction
in an IC engine.
It is proposed that a realistically achieveable goal for the optimisation of
the PRCL contact is a fuel efficiency improvement in the region of 1%. By
optimising cylinder liner surface texture alone, it is suggested that the po-
tential is somewhere in the region of 0.25-0.50%. Although this improvement
may sound small to the reader, it is significant to an engine manufacturer
where every fraction of a percent improvement is worked hard for. A multi-
tude of small, seemingly insignificant, improvements in fuel efficiency, when
added together, lead to significiant improvements in future generations of
engine.

1.2 The Piston Ring-Cylinder Liner Contact


In this section the components of the PRCL contact are introduced, their
purpose outlined, and a brief description of how they are manufactured and
the tribological regime they run under is presented.

1.2.1 The Piston Ring Pack


A typical IC engine has three piston rings which form the piston ring pack.
Larger diesel engines may have four rings. The three piston rings all have
different roles and their shape varies because of this. Fig. 1.1 illustrates a
typical ring-pack for a HDDE.
Upper Compression Ring: Seals the combustion chamber

Lower Compression Ring: Provides additional sealing, scrapes oil


downwards

Oil Control Ring: Acts as a major barrier between the well lubricated
piston skirt area and the starved lubricated region of the ring belt
Essentially the piston rings have two tasks; to stop gases leaving the
combustion chamber and prevent excess oil entering the combustion cham-
ber, they can be classified as a seal. Typically HDDE piston rings are
manufactured from cast iron and the upper compression ring can have an
electrochemically deposited hard chromium coating on its running surface
for improved wear resistance. The manufacturing process for piston rings is
typically casting, followed by the running face being machined, often with
double cam turning.
10 CHAPTER 1. INTRODUCTION

Figure 1.1: An illustration of the piston ring-pack in a HDDE.

1.2.2 The Cylinder Liner


In a HDDE, cylinder liners are manufactured from cast iron. Typically
a machining process known as honing is used to apply the desired finish
to the cylinder liner surface. The grooves that the honing process leaves
behind are believed to be important in controlling the amount of oil available
in the contact, by both retaining oil on the liner surface and improving
the distribution of oil. Another function of the honing texture is to allow
debris, either external debris carried in with the intake air or internal debris
generated from wear of components in the engine, to be channelled away
from the conjunction so as to cause only minimal damage and scratching to
the smooth plateaux which are important for fluid film generation. Usually
in modern engines three stage, or plateau honing, is used. The three stages
are listed below and illustrated in Fig. 1.2.

1. The liner is honed roughly to size

2. The peaks are removed using a smoother honing tool

3. The material that has become embedded in the crosshatch is removed


with a very fine honing tool, or brush.

The second stage, creating the plateaux, partly replaces the running in
process as the peaks that would have worn away during initial running of
the engine are removed.
1.2. THE PISTON RING-CYLINDER LINER CONTACT 11

Figure 1.2: The three stage plateau honing process.

1.2.3 The Tribology of the PRCL contact

The Stribeck curve, Fig. 1.3, is often used to illustrate the different lubri-
cation regimes and corresponding friction coefficient. For a typical HDDE
the entraining velocity between cylinder liner and piston ring pack varies
between zero at the reversal points and up to 16 m/s at mid-stroke. The
temperature of the cylinder liner, which strongly influences the tempera-
ture in the contact and therefore the lubricant viscosity, can vary between
approximately 90 C and 210 C along the axial length of the cylinder liner.
The loading on the contact, which is a combination of static ring tension and
gas pressure acting on the back side of the piston ring, can vary from over
200 bar after combustion to just a few bar on the exhaust or intake strokes.
It is important to realise, due to these hugely varying operating conditions,
that the lubrication regime of the PRCL contact will vary from boundary
and mixed lubrication at the top and bottom of the stroke to fully hydro-
dynamic during mid-stroke. Therefore, when considering the optimisation
of the contact, all lubrication regimes must be considered.
12 CHAPTER 1. INTRODUCTION

Figure 1.3: A typical Stribeck curve.

1.3 Optimizing the PRCL contact


The overall objective of this research is to develop a numerical simulation
tool of the PRCL contact, validated by experiment, that can be used to
assist in optimising the contact with a focus on cylinder liner surface topog-
raphy. However, it is important to define what is meant by optimising the
contact. Three operating parameters should be considered for optimisation;

Frictional power losses

Blowby and oil consumption

Wear

The power losses from the PRCL contact should be minimised in order
to increase the overall efficiency of the engine. It is important to realise
that minimising power loss is not the same thing as minimising friction.
The highest frictional forces, where the friction coefficient is highest, will
occur around Top Dead Centre (TDC) when boundary lubrication occurs.
However, at this point in the engine cycle the piston is travelling at low
speeds and is momentarily stationary. Therefore, the power loss, in watts
or friction mean effective pressure, is minimal. During the midstroke the
lubrication regime is fully hydrodynamic and the friction coefficient is much
lower, however the piston is travelling at many metres per second which
may lead to greater power losses.
The second parameter to minimise is blowby and oil consumption. Sim-
ply put, the combustion gases should stay in the combustion chamber and
1.4. STATE-OF-THE-ART IN PRCL SIMULATION 13

the engine oil should stay in the engine. Blowby is defined as the combus-
tion gases that flow from the combustion chamber past the piston rings and
into the crankcase, resulting in a loss of power and efficiency. Blowby is
a volumetric power loss meaning that minimised blowby is also minimised
power loss. Blowby can occur through three routes;

1. Through the ring gap

2. Past the face of the ring (i.e. between the piston ring and the liner)

3. Between the flank of the ring and the ring groove

The second and third of these leakage paths are through tribological con-
tacts. In particular, the second is the contact that is to be optimized in this
research and therefore the effect on blowby of any optimisation needs to be
considered. Similarly, the amount of oil that passes the piston rings should
be kept to a minimum. Too much oil entering the combustion chamber is
undesirable as it will lead to a substantial increase in exhaust gas hydrocar-
bon levels. Finally, wear should be kept to a minimum. This is primarily
an issue where asperity contact occurs near TDC.

1.4 State-of-the-art in PRCL simulation


In this section the current state-of-the-art in PRCL simulation will be in-
troduced and evalulated in relation to this work. This contact is one of
the most widely investigated in the field of tribology with many simulations
developed at a great number of institutions over the past fifty or so years.
This review of existing work is not intended to be exhaustive, but simply
aims to highlight work which is close to that undertaken in this thesis or
that the author thinks is particularly relevant.
The PRCL contact is extremely complex with a great many phenomenon
that can be modelled, covering several branches of science and engineering,
not simply tribology. The field of dynamics is important for considering the
motion of both piston and rings. The flow of combustion gases past the
ring and heat transfer through the piston, rings and cylinder liner can also
be modelled. Considering the tribology, lubricant transport through the
contact, asperity, hydrodynamic and elasto-hydrodynamic loads, lubricant
cavitation and surface topography have all previously been modelled in lit-
erature. The chemistry of the lubricant, additives and tribofilm can also be
considered and investigated through simulations. It is unrealistic for any
one simulation tool to model all of these effects. Not only would modelling
all these effects be many years work, but the simulation time would also
span several years. Therefore, when developing a simulation tool the end
14 CHAPTER 1. INTRODUCTION

purpose and objectives should always be kept in mind and only the most
relevant effects should be modelled. With the objective of developing an
optimisation process, a simulation tool with a short execution time is re-
quired so that it can be used to evaluate many dozens of different design
permutations.
Extensive simulation work of the tribological conjunction between a pis-
ton ring and cylinder liner has been published over many decades. Two
major commercial tools exist, Excite Piston and Rings [9], developed by
AVL, and RINGPAK [10], developed by Ricardo. In almost all existing re-
search the Reynolds equation is solved across the contact width with either
Reynolds boundary conditions [11, 12, 13] or a cavitation algorithm type
solution [14, 15, 16, 17, 18], such as that derived by Elrod [19] or similar.
PRCL models must also consider any asperity contact that may occur at
the ring reversal points. In virtually all existing research this is achieved
with the implementation of a Greenwood and Tripp contact model [20]. An
axi-symmetric assumption, that the ring is infinitely long around its circum-
ference and hence that the problem can be considered one dimensional, is
also widely used. However, a few models [21] incorporate a circumferentially
variable oil film that is inherent with a distorted bore and model how the
ring conforms to this deformation. In this work the ring will be considered
axi-symmetric. The motivation behind this assumption is thus: while the
author feels that ring conformability is a most important phenomenon to
consider when investigating large scale effects such as bore distortion, this
thesis aims to investigate cylinder liner texture, a feature on a much smaller
scale which can be observed with an axi-symmetric model. Therefore, with
the motivation of limiting the computational time of the developed simula-
tion, the axi-symmetric assumption is applied.
Many models do not consider the effect of surface topography in full film
lubrication by assuming that the surfaces are smooth [11, 14, 22], but it is
proposed that the cylinder liner honing must have some effect even when
asperity interactions do not take place. However, others [12, 13] incorporate
Patir and Cheng [23] flow factors to simulate the effect of surface topography
in the full film lubrication regime. As the simulation developed for this thesis
aims to investigate cylinder liner surface texture in all lubrication regimes,
a similar approach should be used.
Most PRCL simulations assume that the inlet to the piston ring is always
fully flooded, i.e. there is always enough lubricant present at the inlet of
the contact to fill the gap between cylinder liner and piston ring. However,
it is questionable whether there is always enough oil available to achieve
this, particularly during the midstroke of the cycle whereby high entraining
speeds and low supported load allow for a high predicted film thickness.
However, there are some simulations that run with a starved inlet condition,
1.5. OBJECTIVES OF THIS THESIS 15

such as those by Ma et al. [24] and Liu and Tian [25]. These simulations
track the oil that is available on the liner, so that the inlet condition has
the possibility to become starved. As oil consumption is a function of the
PRCL contact that should be considered in any texture optimisation, the
possibility of a starved inlet condition should be considered in this work.
While many of the aforementioned simulation tools have been used to
predict friction in an IC engine ring pack and compared with experimental
results [26, 16], very few have been used to investigate the effect of differing
surface topography on the lubrication of the PRCL contact. Michail and
Barber [27] implemented an analytical implementation of the honed surface.
Their study used Patir and Cheng flow factors and took the honed surface
to be a combination of cosine waves with the peaks removed to simulate
the plateaux. However, they had no model for asperity contact and their
solution to the Reynolds equation was not mass-conserving. Jocsak et al.
[28] also considered varying honing angles on the effect of lubrication and
friction in a PRCL contact. While their work was mass-conserving unlike
Michail and Barbers work, only one ring was simulated and no oil transport
model was implemented. Therefore the effect of varying surface topography
on oil consumption and full ring pack friction was not fully captured.
Therefore, it is suggested that there is a gap in existing research for
further study of the effect of surface topography on the friction and oil
consumption of a full ring pack, throughout all lubrication regimes, with
the aid of an experimentally validated model. Furthermore, an attempt
should be made to optimise the texture using the developed model. To be
state-of-the-art, a mass conserving cavitation algorithm incorporating an
oil transport model should be implemented that allows for the possibility
of a starved inlet condition and an accurate evaluation of oil consumption.
As the purpose of this tool is to aid in optimising texture, while modelling
these effects the simulation time should be kept to a minimum so that many
simulations can be run to investigate many different surface topographies.

1.5 Objectives of this thesis

In this thesis, a validated simulation tool of the PRCL contact has been
developed to aid in the optimizing of cylinder liner surface texture. The
work presented here can be broken down into three separate sub-objectives,
in which attempts are made to address several gaps in existing research.
16 CHAPTER 1. INTRODUCTION

1.5.1 Implementation of an all-regime model incorporating


the effect of surface texture
As previously stated in the introduction, the PRCL contact runs in all
lubrication regimes, from boundary to full-film. A tribological model will
be developed that is able to simulate all regimes. Furthermore, the model
should;

1. be capable of differentiating between different cylinder liner surface


topographies. This is of great importance as the objective of this
research is to optimise surface topography.

2. run in a reasonable time that enables the tool to be applicable in an


industry setting.

The PRCL contact will be modelled using a modified version of the all
regime model developed by Sahlin et al. [29, 30], whereby the homogeniza-
tion technique is used in place of Patir and Cheng flow factors. By utilizing
flow factors both the space and time discretization can be made coarser
enabling a faster solution time in the global, or macro-scale, piston ring
solution.

1.5.2 Validation of the simulation tool


Another objective is to validate, with experiments, the numerical model
developed in respect to the friction calculated. Although it is notoriously
difficult to measure PRCL contact friction in-situ, due to harsh environmen-
tal conditions and difficulties in accessing the contact, validation should be
attempted to increase confidence in the model. Therefore, both bench tests
and full engine tests have been run to aid in validation.

1.5.3 Working towards optimum texture


As the thesis title indicates, the goal of this work is a tool, or methodology,
for optimizing cylinder liner surface texture. The question of optimum
texture is not an easy one to answer. It is maybe too bold to claim, with
over fifty years of research already carried out by the tribology community
in this field, to fully optimise texture - particularly in an academic work
several steps removed from the industrial application. However, an initial
parametric study of cylinder liner surface texture is undertaken and and
an initial theory and suggestions for working towards the optimum will be
presented.
Chapter 2

Methods and Tools

This chapter outlines some of the tools and techniques that have been eval-
uated, tested, or used in developing the final simulation model, which is
presented in the following chapter. Firstly, there is a discussion on the def-
initions of texture and roughness in a lubricated contact and how they can
be treated in a tribological simulation. Secondly, there is a discussion of
various solutions to the Reynolds equation including a variety of cavitation
algorithms, an evaluation of their suitability for tackling the PRCL problem,
and a decision on the algorithm to be carried forward. Following on from
this there is a review of flow factors, the homogenization technique imple-
mented in this work, and how this can be incorporated into an all regime
mixed lubrication model. Finally, the chapter concludes with an explanation
of the surface measurement tehniques used throughout this thesis.

2.1 The treatment of texture in a simulation tool


Cylinder liner surface topography, whether honed, lased textured or pro-
duced with some other manufacturing technique, can be said to have both
texture and roughness. These two definitions can be considered different
length scales of the surface topography. In this work, the word texture
describes surface topography that is non-gaussian in nature and has some
structure to it. Moreover, this structure is desired and is thought to be ben-
eficial to the tribology of the contact, whether by reducing friction, wear,
or aiding lubrication. Roughness is defined as the surface topography that
has a smaller wavelength than texture and does not have any defined struc-
ture. It is, in essence, the surface topography that lies between the texture.
It may or may not be desirable to the tribology of the contact and exists
due to the fact that no surface can be perfectly smooth, and with regard
to machining smoother generally equates to costlier. Fig. 2.1 illustrates the

17
18 CHAPTER 2. METHODS AND TOOLS

distinction between texture and roughness for a honed cylinder liner surface.
Of course, greater wavelengths of surface topography exist above texture -

Figure 2.1: Texture and Roughness of a cylinder liner surface

any machined surface will no doubt exhibit waviness, out-of-roundness and


some global shape. However, in defining the scope of this work a line must
be drawn somewhere, and here only texture and roughness will be investi-
gated. The work of Mishra et al. [11], Liu and Tian [25], Hu et al. [12]
and Ma et al. [15] are good references for the investigation of surface topog-
raphy of a longer wavelength than texture in the PRCL contact. Smaller
wavelengths also exist, and the field of tribology allows for contacts to be
investigated on the molecular level. These small, sub-micron scales will also
be neglected in this work to keep the task at hand manageable with respect
to solution time, as per the objectives of this work. It is proposed that a
so called one-dimensional axi-symmetric model, where the piston ring is
considered infinitely long and any out-of-roundness neglected, is sufficient
for a simulation developed with the objective of optimising texture, a fea-
ture existing on a smaller wavelength. The assumption is made that the
optimum texture for a perfectly round bore will be similar to that for a
deformed bore.
Now that texture and roughness have been defined, the discussion should
move on to how to include them in a tribological simulation. It could
be argued that ideally, when a texture and roughness have been defined
in a lubricated contact, such as in Fig. 2.2, the whole surface (or more
correctly the gap between surface and counter-surface) should be meshed,
with for example finite differences or finite elements, and then some form
of the Reynolds or Navier-Stokes equation solved to fully understand the
lubricant film in the conjunction. However, there are some drawbacks in
doing this. Firstly, is Fig. 2.2 representative of the whole surface? Should
the measurement be bigger? Or taken at another location on the surface?
Furthermore, how big does the mesh need to be to resolve all features? 500
by 500 nodes? 1000 by 1000? Will this create an excessively long simulation
time when a full, time-dependent, system is considered? Even further, what
if the measurement in Fig. 2.2 doesnt even exist? What if, as is often the
case with engineering simulation, we want to investigate a surface or contact
2.2. THE REYNOLDS EQUATION AND CAVITATION 19

Figure 2.2: Illustration of texture and roughness on an optical measurement


of a cylinder liner surface.

that hasnt been manufactured yet? What then? In this chapter the aim is
to address these issues.

2.2 The Reynolds equation and Cavitation


Traditionally for thin film flow in a lubricated tribological contact the Reynolds
equation is solved [31] to calculate the pressure in a lubricant film and the
corresponding load carrying capacity. With the assumptions of constant
lubricant density and viscosity, and entraining motion of one surface only
in the x direction, the Reynolds equation can be written in the form:
   
3 p 3 p h h
h + h = 6U + 12 (2.1)
x x y y x t

where x and y are space coordinates, h is the separation between the sur-
faces, p is the pressure in the lubricant film, is the dynamic viscosity of
the lubricant, t is time and U is the entraining velocity in the x direction.
The terms on the left hand side of the equation represent the Poiseuille, or
pressure-driven flow. The first term on the right hand side represents the
Couette, or shear-driven flow, and the final term represents the squeeze, or
time dependent flow. In Fig. 2.4 the Reynolds equation, Eq. (2.1), is solved
for the problem illustrated in Fig. 2.3. This geometry could represent a
typical PRCL contact, however a similar geometry can also be observed for
a journal bearing, but with a large radius for the lower surface as opposed
to the flat surface of the cylinder liner. In the leading, left-hand portion of
the ring the converging profile generates a positive pressure in the lubricant
film. In the trailing portion of the ring, as the two surfaces diverge away
from each other the solution of the Reynolds equation (Eq. (2.1)) predicts
a negative pressure of the same magnitude as the positive pressure in the
20 CHAPTER 2. METHODS AND TOOLS

Figure 2.3: Example piston ring geometry.

inlet (due to the symmetric nature of the contact). However, this negative
pressure is physically unrealistic. Normal lubricants cannot stand large and
continuous negative pressures without rupturing. If the pressure profile in
Fig. 2.4 is integrated to calculate the load carrying capacity of the lubricant
film, the solution will be zero. The Reynolds equation will never predict any
load carrying capacity for the PRCL contact in steady state sliding with a
symmetric ring profile. In reality, under these conditions the lubricant will
not be able to sustain such a high negative pressure and will break down
and cavitate. When this cavitation occurs the lubricant forms a mixture of
lubricant and gas, with the gas present as either bubbles or striations. The
pressure in this cavitated region is defined as the cavitation pressure, which
is typically somewhere in the region of atmospheric pressure. This phe-
nomenon has been observed experimentally in several pieces of literature,
such as the work of Dellis and Arcoumanis [32].
Therefore, the Reynolds equation in its original form, Eq. (2.1), is not

50
Lubricant Pressure (MPa)

25

25

50
0 0.2 0.4 0.6 0.8 1
Ring width (mm)

Figure 2.4: Solution of the Reynolds equation for a converging-diverging


wedge.
2.2. THE REYNOLDS EQUATION AND CAVITATION 21

suitable for modelling the piston ring contact and must be modified in some
way. One oft-used solution to cavitation modelling in the PRCL contact is
simply to set the negative pressures calculated in the diverging portion of the
ring to zero. This is known as the half-Sommerfeld solution. However, while
this clearly eliminates the issue of zero calculated load carrying capacity, it
is a non-physical solution and the mass continuity of lubricant flow through
the contact is lost. Over the last thirty years a multitude of mass-conserving
cavitation algorithms have been developed, of which the work of Elrod [19],
Vijayaraghavan and Keith [33], Ausas et al. [34] and Giancopini et al. [35]
are just a few. All of these works modify the Reynolds equation in such a way
that cavitation is modelled while mass-continuity is preserved throughout
the contact. Throughout the course of this research work all four of these
algorithms have been implemented. With [33] featuring in Paper A, [34] in
Paper B and [35] in Paper D and E. More detail can be found in the original
articles referenced above and in the papers appended to this thesis.
The Elrod [19] algorithm has been widely used and referenced by re-
searchers over the past thirty years. Vijayaraghavan and Keith [33] build on
the work of Elrod and presented the derivation of the differencing scheme, as
opposed to Elrod who resorted to considerable experimentation. However,
both the Elrod and Vijayaraghavan schemes can run into stability problems
around the cavitation boundary where it is difficult to find a converged
solution. An illustration of this issue is presented in [36]. An algorithm
presented by Ausas [34] was found to be more stable and implemented in
Paper B, however the real breakthrough in this work came when the au-
thor implemented the Giacopini et al. [35] algorithm. This was found to
be incredibly stable and elegant to implement, and although not quite as
fast (in this authors implementation) as the Elrod or Vijayaraghavan ap-
proaches, was chosen as the cavitation algorithm of choice, to be used for
the remainder of this work.
The Giacopini et al. algorithm [35] is used for the majority of appended
papers and therefore a brief description of it will be given now. Stating the
Reynolds equation in one-dimension for simplicity:
 
3 p h h
h = 6U + 12 (2.2)
x x x t

where the density, , is now variable. First complimentarity conditions must


be introduced. Specifying the pressure as:

p = ps + pc (2.3)

where pc is the cavitation pressure and ps is a so-called solution pressure.


In the full film regime the solution pressure, ps , will always be greater than
22 CHAPTER 2. METHODS AND TOOLS

0, while in the cavitated zone the solution pressure is assumed to equal 0,


so that when Eq. (2.3) is applied the lubricant pressure, p, equates to the
cavitation pressure, pc . Therefore,

ps 0 (2.4)

Introducing a new variable, r,



r =1 (2.5)
c
where c is the reference lubricant density in the full film region and is
the density of the lubricant and gas mixture, which in the cavitated region
will be less than c . Therefore, it can be stated that

r0 (2.6)

and:
ps r = 0 (2.7)
These conditions mean that ps and r can be defined as linear complimentar-
ity variables. Now Eq. (2.2) must be modified to introduce the new variable
r into the equation. Rearranging Eq. (2.5), substituting into Eq. (2.2) and
cancelling out density, gives:
   
ps (ps ) h (rh) h (rh)
h3 rh3 = 6U 6U +12 12
x x x x x x t t
(2.8)
In the cavitated region the pressure, ps = 0 and also ps /x = 0, and in
the active region r = 0, therefore:
ps r
=0 (2.9)
x
giving a final equation of:
 
3 ps h (rh) h (rh)
h = 6U 6U + 12 12 (2.10)
x x x x t t
with the complimentarity conditions:

ps 0
r0 (2.11)
ps r = 0

ps can be solved for and then Eq. (2.3) can be used to calculate the lubri-
cant pressure p. The advantage with the LCP formulation is that standard
2.3. FLOW FACTORS 23

techniques can be used to solve the problem numerically, e.g. Lemkes piv-
oting algorithm. See e.g. the book by Cottle et al. [37]. This alleviates
the problematics associated with discrete formulations that change at the
boundaries between the cavitated and the full film zones. Moreover, this
solution technique finds the solution in a finite number of steps, hence issues
related to iterative processes are avoided.
In Fig. 2.5 the previous problem, illustrated in Fig. 2.3, is solved as it
was in Fig. 2.4. The dashed line, the half-Sommerfeld solution, is equal to
the solution in Fig. 2.4 with the negative pressures set to zero. The solid
line, LCP, is the solution to the Giacopini et al. [35] cavitation algorithm,
Eq. (2.10). The figure illustrates that the predicted hydrodynamic pres-
sure, and therefore load carrying capacity, is greater with the LCP solution.
Furthermore, the mass flow through the contact is constant with the LCP
solution, as it should be, the solution is mass conserving. With the half-
Sommerfeld solution the mass flow is not conserved in the cavitating region
where the pressure is set to zero.

2.3 Flow Factors


As previously stated the objective of this research is to model the effect
of texture on the oil transport of a piston ring-pack. At this point, it is
important to distinguish between stationary and non-stationary texture.
In this work the solution domain is the piston ring contact width. If
surface texture was applied to the piston ring surface, then it could be
described as stationary, as the texture does not move relative to the solution
domain. However, if the texture is applied to the liner surface, as it is in
the case with the honing grooves that this research is investigating, then
the texture must be described as non-stationary, as it moves relative to the
solution domain. This is illustrated in Fig. 2.6.
Stationary and non-stationary texture will influence the lubrication of
the PRCL differently, particularly when it comes to oil transport through
the contact. Imagine that the lubricant oil film becomes very small. If
transverse texture is applied to the ring (stationary), then very little lubri-
cant will pass through the contact as it becomes trapped in the texture,
having the velocity profile of the ring (U = 0 with the moving coordinate
system attached to the ring). However, if the transverse texture is applied
to the liner (non-stationary), then the texture will, in effect, drag lubricant
though the contact with a velocity U and the oil transport will be greater.
Therefore it is very important that in this work, with the objective of
investigating oil transport through the piston stroke, that the texture is
treated as non-stationary. However, non-stationary texture presents certain
24 CHAPTER 2. METHODS AND TOOLS

60

Lubricant Pressure (MPa)


LCP
Half-Sommerfeld
40

20

0
0 0.2 0.4 0.6 0.8 1
Ring width (mm)
(a)
30
LCP
25
Half-Sommerfeld
Mass flow (-)

20

15

10

5
0 0.2 0.4 0.6 0.8 1
Ring width (mm)
(b)

Figure 2.5: Comparison of hydrodynamic pressure (a) and mass flow (b) for
the Half Sommerfeld and LCP solution.

challenges in simulation. If the surface texture is included in the solution


domain, then a very fine mesh in the space domain is required to be able
to resolve the texture across the piston ring width. This in itself may not
be a major issue, however as the texture is non-stationary sufficiently small
time steps must be taken to accurately resolve (in the time domain) the
texture moving through our solution domain, the ring width. In the engine
simulated in this work the stroke is 160 mm and the honing texture has
a wavelength of approximately 0.3 mm. If it is assumed that 10 steps in
2.3. FLOW FACTORS 25

Figure 2.6: An illustration of the difference between stationary and non-


stationary texture, with stationary texture illustrated on the left and non-
stationary on the right. The green grid illustrates the necessary space do-
main mesh and the red grid represents the time domain mesh.

time are necessary per texture wavelength, then over 20,000 time steps are
required to complete the 4 stroke cycle, where at each time step a solution
must be found on what is already a very fine mesh, for three piston rings.
The result is a simulation time that is far too long to allow the resulting
tool to be useful for investigating the PRCL contact in an industry envi-
ronment, where a trade off between accuracy and computational cost must
always be sought. Therefore, flow factors are calculated using a homogeniza-
tion technique to incorporate the effect of non-stationary surface texture in
the PRCL contact. These flow factors allow a smooth solution domain to be
solved and the effect of non-stationary texture incorporated in the solution
by the addition of flow factors in the Reynolds equation. These flow factors
are pre-calculated by analysis of a representative section of cylinder liner
texture. This solution method allows for a much coarser mesh in the space
and time domains, and hence a much faster solution time.
At this point the scales global and local will be introduced. The contact
solution domain, incorporating the contact geometry, that the Reynolds
equation Eq. (2.2) or LCP cavitation equation Eq. (2.10) are solved on is
referred to as the global scale. The representative area of shorter wavelength
surface topography, used to calculate the flow factors, is referred to as the
local scale. Fig. 2.7 illustrates these global and local scales.
The flow factors implemented in this thesis are based on pre-existing
work in the Division of Machine Elements at Lule a University of Technol-
ogy, in particular that of Almqvist [38]. Almqvist and co-authors developed
the particular homogenization theory of the Reynolds equation that is im-
plemented here. Therefore, in this thesis there will not be a huge amount of
detail and repetition of the theory and derivations involved, but instead a
focus on the utilisation and details surrounding the implementation of this
pre-existing work.
26 CHAPTER 2. METHODS AND TOOLS

Figure 2.7: An illustration of the separation of the problem into Global and
Local scales and the coupling between them with flow factors.

Homogenization is a mathematical averaging technique. When applied


to the Reynolds equation, small wavelength features of the film thickness, i.e.
surface roughness, are assumed to be periodic in nature. By assuming that
the wavelength of these features (roughness) tends to zero, a homogenized
form of the Reynolds equation, Eq. (2.12), can be formulated whereby the
problem can be considered smooth across the solution domain, on the global
scale, with the film thickness replaced with so called flow factors:


p



p

b12
h
a11 + a22 = 6U + 12 (2.12)
x x y y x t
These flow factors, a function of film thickness, are calculated by solving
cell problems, the local scale, which consist of one periodic window of the
small wavelength features, or roughness, on solution domain Y . The cell
problems which must be solved for a non-stationary problem are;
h3
Y h3 Y 1 =

y1
h3
Y h3 Y 2 = (2.13)

y2
h
Y h3 Y 3 =

y1
which are solved with periodic boundary conditions where 1 , 2 and 3
are local scale solution variables. In the local solution domain, Y , the
coordinate system (y1 , y2 ) is used. The film thickness h is defined as a
range of separations between the cylinder liner surface topography and a
smooth piston ring. Mathematically;
T
h = hr + h (2.14)
Where hr is the measured surface topography and h T ranges from between
a minimum separation and the biggest separation we might see in the piston
ring contact.
2.3. FLOW FACTORS 27

In this section, with the objective of introducing flow factors, only flow
factors in the full film regime will be considered. Therefore, when surface
separation is mentioned it refers to h T , the separation between our sur-
face topography hr and the smooth counter surface. However, in section
2.5, when moving on to discuss a full mixed lubrication model where con-
tact occurs between the surfaces, care must be taken with the definition of
separation and the definition of h T will be revisited.
Once these local cell problems have been solved for a range of h T , the
local scale variables can be integrated over the cell domain Y to give the
flow factors. For the pressure flow factors, aii ;
 
1 1
Z

a11 (hT ) = 3
h 1+ dy
l1 l2 Y y1
T ) = 1 2
Z
a12 (h h3 dy
l1 l2 Y y1
(2.15)
1 3 1
Z

a21 (hT ) = h dy
l1 l2 Y y2
 
T ) = 1 2
Z
a22 (h h3 1 + dy
l1 l2 Y y2
where l1 and l2 are the lengths of our cell problem in the y1 and y2 direc-
tions respectively. a11 and a22 feature in the modified Reynolds equation,
Eq. (2.12). a12 and a21 are the cross flow terms, which can also be incorpo-
rated into the Reynolds equation. However, if the cell problem is periodic
then the cross flow terms equal zero and can be neglected. Due to the way
flow factors are implemented in this thesis a12 and a21 are both excluded
from the Reynolds equation. The shear flow factor, b12 , is defined as;
Z  
T ) = 1
b12 (h h h3
3
dy. (2.16)
l1 l2 Y y1
Finally, the average film height in the squeeze term can be calculated from;
1
Z

h(hT ) = hdy. (2.17)
l1 l2 Y

where h is the average separation, which in the full film regime will equal
T . All of the flow factors for the homogenized Reynolds equation have
h
now been defined. One limitation of the flow factor method is that the
scales must be separable. If the wavelength of the homogenized surface to-
pography is not of a separable-enough scale from the contact width then
a significant error could result from including this topography within ho-
mogenized flow factors. There is no definitive rule as to what constitutes
separable scales. Almqvist and Dasht [39] compared deterministic solutions
28 CHAPTER 2. METHODS AND TOOLS

with varying roughess wavelengths with the homogenized solution. They


found that with a sinusoidal roughness profile the error in supported load
with a fixed separation fell from 7% with a roughness wavelength of 1/24 of
the contact width to 0.5% with a wavelength of 1/28 of the contact width.
Another factor to consider is that any cavitation that might occur due to
features of the surface topography will not be considered if those features
are homogenized; in the model developed in this thesis only cavitation on
the global scale will be simulated. In Fig. 2.8 a11 , a22 , b12 , c11 and d11 flow

2
Normalised Flow Factor

a11
1.5 a22
b12

0.5

0
101 100 101 102

Separation, hT (m)
(a) Reynolds Flow Factors
0
Normalised Friction Factor

0.2

0.4 c11
d11
0.6

0.8

1
101 100 101 102
T (m)
Separation, h
(b) Friction Flow Factors

Figure 2.8: Reynolds and friction flow factors calculated for a typical cylin-
der liner surface against a smooth surface for a range of separations.
2.3. FLOW FACTORS 29

factors are presented for a typical cylinder liner surface moving against a
smooth stationary surface. The c11 and d11 flow factors are introduced in
the next section. In the Reynolds equation (Eq. (2.2)) a11 and b12 replace
the film thickness in the Poiseulle and Couette terms respectively. In the
plots a11 is normalised by h 3 and b12 is normalised by h
T . The flow factors
T
show that at large separations the contact behaves in the same way as the
smooth case, the normalised flow factor is equal to 1. However, as the film
T , is reduced, the a11 flow factor reduces with respect to h
thickness, or h T ,
i.e. the normalised flow factor becomes less than 1. This shows that the
Poiseulle term with homogenized roughness reduces with respect to the case
with a smooth reference surface. The b12 flow factor does the opposite and
the normalised value increases as the separation reduces. This is because
the topography is moving and in effect drags lubricant through the contact
under small separations, leading to a higher Couette flow than with the
smooth case. If the topography was stationary then the opposite would
be true and b12 would reduce as the separation reduced due to lubricant
becoming trapped in the topography.
The author thinks that it is important to note at this point that the cal-
culation of the flow factors for a real surface topography is a non-trivial task.
A large mesh and careful integration of the resulting local scale variables
is required to produce accurate and converged results. For the solutions
presented in this thesis equation 2.13-2.17 were solved with a finite element
solution in Comsol Multiphysics 4.3b.
As was discussed in the preceeding paragraphs, in the PRCL contact it is
desirable to solve the Reynolds equation with the LCP cavitation algorithm
incorporated, see Eq. (2.10). We can substitute our flow factors into the
pressure and shear flow terms in Eq. (2.10):

r

h
 
p b12 (rb12 ) h
a11 = 6U 6U + 12 12 (2.18)
x x x x t t

It is not strictly correct to do this, as the flow factors were derived from
homogenizing the Reynolds equation and not the LCP cavitation problem.
Therefore, mathematically the flow factors should only be used in the ho-
mogenized Reynolds equation, Eq. (2.12). However, from an engineering
perspective it is suggested that it is acceptable to incorporate flow factors
into the cavitation algorithm solution, as this is extremely beneficial in de-
riving a solution that both incorporates cavitation and gives a reasonably
fast solution time. This approach has already been used with Patir and
Cheng flow factors and a switch-type cavitation algorithm by, amongst oth-
ers, Jocsak et al. [28].
30 CHAPTER 2. METHODS AND TOOLS

2.4 Full film viscous friction


One of the most desirable outputs from the final simulation is the friction
in the PRCL contact. Viscous full film friction is a key part of this and for
a smooth case can be calculated per unit length as;
Z  
U h P
F = dx (2.19)
Y h 2 x

When incorporating the effect of rough surfaces in our contact, additional


flow factors c11 , d11 and h1 must be calculated:

1 h 3
Z

c11 (hT ) = dy
l1 l2 Y 2 y1
T ) = 1 h 1
Z
d11 (h dy (2.20)
l1 l2 Y 2 y1

1 1
Z
dy
(hT ) =
h l1 l2 Y h

which can be incorporated in the friction equation to give;


 
Z 
  
1 h ps
fhyd = U (1 r) + 6c11 + d11 dx (2.21)
Y h 2 x1

The calculation of boundary and mixed lubrication friction is discussed in


the next section.

2.5 A Mixed Lubrication Model


As discussed in the introduction to this work, in order to model the PRCL
contact, the capability to model all regimes of lubrication is required. In
the previous sections a method for simulating full film lubrication has been
introduced. Now, a method for simulating asperity contact, as well as how
it is incorporated into the full film model to derive a model for mixed lubri-
cation, will be outlined. The technique described here is based on the work
of Sahlin et al. [29, 30].
In virtually all models of the PRCL contact a Greenwood and Tripp con-
tact model [20] is used for asperity contact ([13, 16, 17, 12, 28, 25, 15, 18, 27,
11, 40]). Although widely used the Greenwood and Tripp method has some
disadvantages. The model does not consider the real surface, but instead
takes into account statistical properties of the asperities on the surface.
These may not accurately describe the stiffness of the surface, furthermore
they can be difficult to quantify for a real surface. Another disadvantage
2.5. A MIXED LUBRICATION MODEL 31

with the Greenwood and Tripp approach is that it is assumed that all asper-
ities are hemispheres with the same radius and there is no influence on the
deformation of one asperity due to contact pressure applied at a neighbour-
ing asperity. Therefore, a full contact mechanics model is implemented in
this work. The model, described in detail by Sahlin et al. [29], is an FFT-
accelerated, boussinesq-type elasto-plastic contact mechanics model. The
model deforms the local solution topography (Fig. 2.7) against a smooth
counter-surface in a frictionless elastic perfectly plastic displacement due
to surface asperity contact. The method is based on that by Stanley and
Kato [41] and later adapted by Almqvist et al. [42] to incorporate plastic
deformation. The output from the contact mechanics simulation is an av-
erage of asperity contact pressure as a function of separation as well as the
full elasto-plastically deformed surface for a range of separations. Fig. 2.9
Mean Asperity Contact Pressure, PCP (MPa)

350

300

250

200

150

100

50

0
101 100 101 102
T (m)
Separation, h

Figure 2.9: Mean Asperity Contact Pressure as a function of separation,


T .
h

shows average asperity contact pressure as a function of h T for a range of


separations for a typical cylinder liner surface coming into contact with a
perfectly smooth section of piston ring. The elastic modulus for the cylin-
der liner is 117.5 GPa and 200 GPa for the piston ring. Poissons ratio was
taken as 0.3 for both surfaces. Plastic deformation was assumed to occur
above a pressure of 2.65 GPa.
The solution process for the mixed lubrication model is outlined in
32 CHAPTER 2. METHODS AND TOOLS

Fig. 2.10. The model takes real measured surface topography as input and
calculates mean asperity contact pressure and flow factors for a range of
separations. This model is simply a pre-processing stage for a global tribo-
logical model. The application of this model to a global problem, that of
the PRCL contact, is described in the next chapter. In step 4) in Fig. 2.10,

Figure 2.10: Solution Process for the mixed lubrication model.

Solution of local scale PDEs, where the local scale variables 1 , 2 and 3
are solved for, for a range of separations, our separation h T must be defined.
This separation h T has already been introduced for the full film regime in
section 2.3 as the distance between the mean line through the rough surface
and the smooth counter surface. However, in the mixed lubrication regime,
where a deformed surface from the contact mechanics model is used, h T
must be defined. The definition of hT is given in Fig. 2.11. In the mixed lu-

brication regime hT is defined as the distance between the mean line through
the deformed surface and the smooth counter surface. In all cases, a small
value of 1011 m is added to h T . This is to prevent film thicknesses of ex-
actly 0 in the mixed lubrication regime for numerical reasons when solving
Eq. (2.13).
The calculation of viscous friction has already been outlined in section
2.6. SURFACE TOPOGRAPHY MEASUREMENT 33

T in the full film and mixed lubrication regimes.


Figure 2.11: Definition of h

2.4. Boundary friction is calculated by taking the average asperity pressure


and multiplying by a friction coefficient found from experiment:
Z
fbd = PCP dA, (2.22)

where is the measured friction coefficient and PCP is the average asperity
contact pressure, taken from Fig. 2.9 for a given separation. The friction
coefficient used in the simulations presented in Chapter 5 was 0.09.

2.6 Surface topography measurement


A stated in the previous section, a real 3D surface measurement can be used
as input into the mixed lubrication model (for more detail, refer to Chapter 3
and Paper D and E). Throughout this work, three techniques have been used
to measure the real surface topography of a cylinder liner surface, vertical
34 CHAPTER 2. METHODS AND TOOLS

scanning interferometry, atomic force microscopy and confocal microscopy.


These technqiues will be outlined in brief now.

2.6.1 Vertical Scanning Interferometry


The key components of a Vertical Scanning Interferometer (VSI) are shown
in Fig. 2.12. Light is split and directed at both the sample to be measured

Figure 2.12: Schematic of the working principles of a Vertical Scanning


Interferometer.

and a reference surface. The light is reflected and recombined. As the


light recombines, interference fringes are formed, which are projected onto
a CCD. The interference signal will have a maximum modulation when the
optical path lengths between the sample and reference surfaces are identical.
The sample is scanned in the z-direction, and therefore the height of each
point on the surface (corresponding to a pixel on the CCD) can be ascer-
tained by the z-height when the modulation is greatest. The VSI technique
was used to measure the surfaces in Paper A, B, C, D and E.

2.6.2 Atomic Force Microscopy


The cylinder liner surface was also measured by an Atomic Force Microscope
(AFM). AFMs can be operated in several different modes, however here
only contact mode will be discussed as this is the technique that was used
for the measurements contained within this thesis (Paper C). Fig. 2.13 shows
a schematic of the key components of an AFM. In contact mode, the force
2.6. SURFACE TOPOGRAPHY MEASUREMENT 35

Figure 2.13: Schematic of the working principles of an Atomic Force Micro-


scope in contact mode.

between the tip and the surface is kept constant during the scanning, by
maintaining a constant deflection. Either the sample or the probe can be
scanned in the x and y directions. Due to the relatively large and heavy (for
an AFM) sample size (a 10 x 10 x 8 mm section of cylinder liner weighing 5.8
g) the probe rather than the sample was scanned. A piezo-electric element
moves the probe in the z-direction to maintain a constant force and this
allows the z-height to be found. The force applied to the probe is found by
monitoring its deflection. A laser is pointed at the back-side of the probe
and any movement will cause the reflected laser light to move location on
the detector.

2.6.3 Confocal Microscopy


The final measuring technique used during this research is Confocal Mi-
croscopy. Although no confocal microscopy data is presented in this thesis,
the technique has been used throughout this work to investigate cylinder
liner surfaces. A schematic of a Confocal Microscope is shown in Fig. 2.14.
The Confocal Microscope only images one discrete point on the surface at
any given time. A laser beam is scanned across the surface with two rotat-
ing, scanning mirrors and the reflected beam is received by a detector. The
sample is scanned in the z-direction and as each point comes into focus the
intensity of the reflected light increases. When the intensity is at a peak for
a given point, its height on the surface can be ascertained.
36 CHAPTER 2. METHODS AND TOOLS

Figure 2.14: Schematic of the working principles of a Confocal Microscope.

2.6.4 Comparison of the different techniques


The AFM measurements were used as a benchmark for the other techniques
and were of extremely high quality (more detail in Paper C). However, the
AFM measurements took a long time to perform compared to the other
techniques and were susecptible to contamination and environmental fac-
tors, such as vibration. Just as importantly, it should also be noted that
a high level of expertise is required to perform AFM measurements well,
to the extent that they were beyond the expertise of this author and were
performed by Illia Dobryden in the Division of Physics at Lule a University
of Technology. The authors opinion is that they are extremely valuable in a
scientific environment for benchmarking purposes, but not for frequent mea-
surements of automotive components by engineers in an industrial setting.
The vertical scanning interferometry measurements were fast to perform,
were within the skill range of the author and provided a high enough level
of accuracy to be used in the later simulations (see Paper C for a comparison
of measurements from the AFM and VSI techniques). Therefore, this was
the preferred technique for measuring the surfaces. Confocal microscopy
also provided high quality measurements, but with the caveat that a high
enough magnification must be used. This is due to the increasing numerical
aperture and depth of field of lower magnification objective lenses, which
2.6. SURFACE TOPOGRAPHY MEASUREMENT 37

causes a reduction in the vertical resolution of the confocal microscope. The


reason for confocal microscopy ever being favoured over vertical scanning
interferometry was purely a practical machine specific reason, with the con-
focal microscope that the author had access to having an excellent stage
that allowed the surface to be scanned and large measurements stitched
together.
38 CHAPTER 2. METHODS AND TOOLS
Chapter 3

Development of a PRCL
simulation tool

This chapter describes the full model of the PRCL contact that has been
developed using the methods and tools outlined in the previous chapter. In
the first section, a global overview of the problem is given, outlining the
environmental conditions that the PRCL is exposed to, the geometry and
variables of the contact. Following on from this is a section on the method
for calculating the boundary conditions that act on the ring. The third sec-
tion describes how texture is treated in the model with the mixed lubrication
model and flow factors described in the previous chapter. The penultimate
section deals with friction and wear in the contact and the final section
outlines the general solution procedure for the full ring-pack model.

3.1 Global overview of the problem


When building a model of the PRCL contact it is first beneficial to define the
entraining velocity, pressure and temperature that the contact is exposed
to.

Entrainment velocity

The three piston rings that together form the piston ring-pack reciprocate
along the axial length of the cylinder liner with the piston. In this work,
due to the lack of a ring dynamics model, it is assumed that the rings travel
at exactly the same velocity as the piston and do not move relative to the
piston in the axial direction, i.e. no ring flutter occurs. The velocity of the

39
40 CHAPTER 3. DEVELOPMENT OF A PRCL SIMULATION TOOL

piston can be described with Eq. (3.1).



LN sin cos
U= 1 +  (3.1)

1/2
60 2k 2

L sin2

where L is the stroke length, k is the conrod length, N is the engine speed
and is the crank angle. The engine studied in this thesis has a stroke (L)
of 0.16 m, a conrod length (k) of 0.255 m and an engine speed (N ) ranging
from between 600 rpm at idle and 2400 rpm. In a truck the engine spends
most of its time at 1200 rpm, which corresponds to highway cruising speed,
and so it is this speed which will be used throughout this thesis. Fig. 3.1
illustrates the piston velocity through one revolution of the engine at idle,
1200 rpm cruising speed, and 2400 rpm.

600 RPM
20
Piston Velocity (m/s)

1200 RPM
2400 RPM

20

0 90 180 270 360



Crank Angle ( )

Figure 3.1: Piston velocity for engine under investigation at 600, 1200 and
2400 rpm.

Gas Pressure
It is necessary to know the gas pressures that the piston rings are exposed
to for two reasons. Firstly, in the case of the upper compression ring (and
in some cases the lower compression ring), the majority of the load that
the contact must support originates from the gas pressure acting on the
back face of the piston ring, forcing it against the cylinder liner surface.
In the model developed in this thesis it is assumed that the largest of the
two pressures acting on top and bottom flank of the ring forces the ring
against the opposite side of the piston groove and acts on the back face
3.1. GLOBAL OVERVIEW OF THE PROBLEM 41

Figure 3.2: Illustration of the pressure acting on the back face of the ring.

of the piston ring. This is illustrated in Fig. 3.2. The pressure above the
upper compression ring, the top land pressure, is subject to the combustion
chamber pressure which can reach over 200 bar during combustion. Across
the first ring land, between the upper and lower compression rings, the
pressure is less due to the sealing action of the upper compression ring.
Even so, the first ring land pressure can reach similar values to the top land
pressure due to leakage past the upper compression ring at some points in
the engine cycle. The oil control ring provides very limited sealing capacity
and the pressure across the second ring land, between the lower compression
ring and oil control ring, is very low. The crankcase pressure, due to its large
volume and the crankcase breather, is always atmospheric.
The pressure data used in the simulations in this thesis is an input
parameter and is provided by Scania. The combustion chamber pressure is
measured by placing a pressure transducer in the cylinder head and then
the inter-ring pressures are calculated by means of an analytical orifice and
volume type model, using the commercial software AVL Excite Piston and
Rings [9]. Fig. 3.3 illustrates typical input data for top, first and second
ring land pressures resolved against crank angle.

Temperature
Another important parameter is the temperature of the lubricant in the
contact. The temperature of the lubricant will change the viscosity and this
is a key parameter when calculating the film thickness and viscous friction
in the contact. In this work it is assumed that the lubricant in the contact
is at the same temperature as the cylinder liner surface. In reality it is likely
42 CHAPTER 3. DEVELOPMENT OF A PRCL SIMULATION TOOL

250
Top Land
200 First Ring Land
Pressure (bar)
Second Ring Land
150

100

50

0
0 90 180 270 360 450 540 630 720

Crank Angle ( )

Figure 3.3: Piston land pressures as a function of crank angle.

that the temperature of the lubricant will be higher than that of the cylinder
liner surface due to shear heating of the lubricant as it passes through PRCL
contact, and also due to heat transfer through the ring pack from the hotter
piston. However, in the absence of a measured contact temperature or ther-
mal model, such as that developed by Morris et al. [43], the cylinder liner
surface temperature will be used as the lubricant temperature in this work.
It is suggested that any lack of accuracy caused by neglecting these effects
is most probably smaller than the inaccuracy of the initial estimation of
cylinder liner surface temperature anyway. The cylinder liner surface tem-
perature is calculated from a 2D heat transfer model, illustrated in Fig. 3.4.
The model calculates the heat transfer through the liner from the combus-
tion chamber to the water cooling jacket. This calculation was performed
by the author as part of his work at Scania, using a method developed in-
ternally at Scania by Peter Daelander. The results from the 2D model were
validated by comparing with experimental data where thermocouples were
placed in the cylinder liner wall of a fired engine. The calculation technique
is not part of this thesis and Fig. 3.4 is only included to show the input
into the model and the temperature variation along the liner length. For
confidentiality reasons the scale has had to be witheld from the figure. The

Figure 3.4: 2D temperature field of the cylinder liner.


3.1. GLOBAL OVERVIEW OF THE PROBLEM 43

readers attention should be drawn to the higher temperatures around the


top section of the liner, closest to the centre of combustion - this will reduce
the lubricant dynamic viscosity at the location where the load acting on the
top compression ring from the combustion chamber gas pressure is highest.

Ring geometry
The geometry of the piston ring running surface is another critical input as
this global geometry will influence how lubricant is entrained into the inlet
region and the onset of cavitation in the outlet region. The running face
of each of the three piston rings was measured with a confocal microscope
and the filtered and levelled profiles are illustrated in Fig. 3.5. These are
worn profiles, i.e. they are measured from piston rings that have been
run-in in an engine test. The motivation for this is that the wear rate of
new piston rings in an engine is initially at a high rate. It is difficult to
accurately simulate this initial running in and therefore to be able to run a
simulation with geometries that are representative of the rings for most of
their life, the measured run-in profiles are used. The profiles are all plotted

50
TCR
40 2CR
Ring Height (m)

OCR
30

20

10

0
1 0.5 0 0.5 1 1.5 2
Ring Width (mm)

Figure 3.5: Ring Profiles of the three rings forming the ring-pack. In the
figure the right hand side faces towards the combustion chamber and the
left hand side faces towards the crankcase.

on the same figure so that comparisons can easily be drawn between their
different profiles. The oil control ring in the engine simulated is twin-land,
meaning that there are two rails, or running profiles, separated by a gap
(as illustrated in Fig. 1.1) of 2.8 mm. These lands are assumed to have the
same profile and therefore only one is illustrated in Fig. 3.5.
44 CHAPTER 3. DEVELOPMENT OF A PRCL SIMULATION TOOL

Global solution
The global solution domain for each of the piston rings including all the
associated nomenclature is illustrated in Fig. 3.6. In Fig. 3.6, U is the

Figure 3.6: Schematic of lubrication in the PRCL contact and corresponding


nomenclature.

entraining velocity of the piston (see Fig. 3.2), hin is the height of lubricant
available on the liner at the ring inlet, hex is the height of lubricant left
behind on the liner at the ring outlet, Pin is the pressure in the lubricant
at the ring inlet and Pout is the pressure in the lubricant at the ring outlet.
The calculation method for obtaining hin and hout is discussed in detail
in section 3.2. The pressure of the lubricant at the inlet and outlet of the
contact, Pin and Pout , is assumed to equal that of the gas pressure acting on
either side of the ring, which is an input parameter to the model as explained
previously (see Fig. 3.2). When Eq. (2.18) is solved across the width of the
ring the cavitation pressure in the lubricant, pc in Eq. (2.3), is set to a
linear gradient between Pin and Pout across the ring width. This means
that it is assumed that the lubricant will cavitate at the local atmospheric
pressure, which is interpolated across the ring width if a pressure gradient
exists between the upper and lower sides. As was previously mentioned in
section 2.3, it should be remembered that the coordinate system follows the
piston so that even though in reality it is the piston that is moving, in this
study it is the cylinder liner that moves through the solution domain.

Other parameters
Other parameters that are required as input into the model are the flow
factors which are necessary for the solution to Eq. (2.18) and in the viscous
friction calculation, Eq. (2.21). These flow factors must describe the surface
3.2. BOUNDARY CONDITIONS ON THE PISTON RING 45

texture that we aim to optimise in this work and the process implemented is
described in section 3.3. Finally, the lubricant viscosity must be calculated
as a function of temperature. This is achieved with the implementation of
the Vogel equation;
B
= Ae (Tl +C) (3.2)
where is the lubricant dynamic viscosity, Tl is the cylinder liner tempera-
ture underneath the piston ring and A, B and C are constants in the Vogel
equation, taken as 0.04899 P a s, 1174.2 C and 123.9 C respectively. The
constants were provided by Scania.

3.2 Boundary conditions on the piston ring


In the simulations presented in this chapter and the results presented in the
next chapter the problem is considered axi-symmetric and therefore one-
dimensional; the piston rings are modelled as infinitely wide and the effect
of the ring gap is assumed to be negligible. However, in Paper A and B the
cylinder liner texture is treated deterministically with periodic boundary
conditions and a 2D problem is solved. In Paper D and E simulations are
performed on a section of piston ring and the infinitely wide assumption is
no longer appropriate and so a finite width 2D solution domain is considered.
The LCP formulation of the Reynolds equation incorporating flow factors,
Eq. (2.18), is solved across the width of the piston ring, Fig. 3.6. We must
specify boundary conditions when we solve Eq. (2.18). As specified in the
previous section, pc is set to equal Pin at the ring inlet and Pout at the
ring outlet. Therefore, in Eq. (2.18) we specify ps to equal zero at the inlet
and outlet. However, we must also specify a value of r, which due to our
complimentarity conditions must be greater than 0. The value of r0 (the
0 subscript refers to location x0 , at the inlet to the ring) is a function of
hin , the lubricant on the liner at the inlet to the contact, left behind by the
previous passage of a piston ring. Immediately prior to the ring, in front of
the solution domain, the mass flow per unit width, m in , can be described
as;
m
in = U hin (3.3)
where is the density of the lubricant. In this work the density is kept
constant with temperature and a value of 900 kg/m3 is used. Immediately
on entering the solution domain, at x0 , the mass flow can be derived from
Eq. (2.18) as;
b120 (1 r0 )U a110 ps
m0= (3.4)
2 12 x
It should be observed here that if r > 0, i.e. the inlet is starved, then ps
must equal 0 due to the complimentarity requirements and therefore the
46 CHAPTER 3. DEVELOPMENT OF A PRCL SIMULATION TOOL

second term in the equation can be neglected. If there is more lubricant


approaching the contact at xin than is needed to fully-flood the contact
inlet (when r = 0) i.e;
b120 U
U hin > (3.5)
2
then the contact begins to scrape oil into a zero-space-dimensional reservoir
in-front of the ring, holding a lubricant volume ms , at a mass flow rate per
unit width of m s . The phrase zero-space-dimensional simply means that
the reservoir doesnt exist anywhere in the solution domain, it is a store of
the oil that is being scraped outside of the solution domain in front of the
piston ring. This mass of lubricant is stored until the ring inlet is no longer
fully-flooded, (i.e r > 0), at which point it becomes a mass flow back into
the contact. This analysis of the inlet boundary condition is best explained
with a flowchart, as shown in Fig. 3.7. This flowchart simplifies the process
somewhat by assuming that the Poiseuille flow is insignificant in comparison
to the Couette flow. In practice, when the inlet is fully flooded the Poiseuille
term must be considered too. At the contact exit, xL , the solution is rather

Figure 3.7: Flow chart describing the inlet solution procedure


3.3. SURFACE TEXTURE AND FLOW FACTORS 47

simpler, rL is set to 1 at xL as this value does not affect the upwind solution.
At the contact exit it is assumed that all of the lubricant is deposited on
the liner surface, at a height given by:

b120
hex = (3.6)
2
By now the reader has most likely realised that in the method presented
here there is a velocity profile discontinuity over the boundaries at x0 and
xL . While mass is conserved both through and either side of the solution
domain, at the x0 and xL boundaries the velocity profile switches from
having a constant profile outside of the solution domain to the linear profile
of Couette flow inside it. This anomaly is an inherent feature in using
this cavitation aglorithm across the entire solution domain. In effect, as
soon lubricant enters the contact it is assumed to bridge the gap between
piston ring and liner, forming striations. Clearly, this is a false assumption,
however the anomaly has no effect on the reformation point, cavitation
point, pressure build-up or load carrying capacity of the lubricant film, and
therefore the authors propose that while the velocity profile in the inlet
region is incorrect, it is an acceptable assumption to make as it simplifies
the solution by allowing for the application of Eq. (2.18) across the entire
solution domain. Having said this, this assumption of Couette flow in the
inlet would cause an over-estimation of the viscous friction in the contact.
Therefore, a correction for this is applied when viscous friction is calculated,
this is discussed further in section 3.4.

3.3 Surface texture and flow factors


To be able to effectively investigate and optimise surface texture many dif-
ferent surfaces should be evaluated. However, real measured surface texture
is only available for current, in production components. Therefore, in this
section the generation of artificial surface texture is introduced. This allows
any combination of honing groove depth, spacing or angle to be investi-
gated easily while controlling all other parameters. Only traditional, cross-
hatched, honed surface texture is investigated in this paper. However, there
is no reason why this method cannot be applied to laser textured dimpled
such as that presented by Rahnejat et al. [44] or Etsion [45].
Fig. 3.8 illustrates the three steps for generating honing texture. In
the first step, Fig. 3.8a, a single honing groove is generated using Eq. (3.7)
(taken from [46]).
2
h (x, y) = D10w(x+y) cos [2 (x + y)] . (3.7)
48 CHAPTER 3. DEVELOPMENT OF A PRCL SIMULATION TOOL

(a) A single honing groove

(b) Mirrored groove to form texture

(c) Final texture with plateau roughness applied

Figure 3.8: Steps for generation of surface texture.

where D is the honing groove depth (see Fig. 3.8c), w is a variable used to
define the spacing S of the honing grooves and is related to the desired
3.4. FRICTION 49

angle of the honing groove, according to:

1
= (3.8)
tan

where is the desired honing angle (see Fig. 3.8c, where the honing angle
is marked relative to the entraining direction). In the simulations presented
in this thesis the width of the honing groove at plateau level was kept at
a constant 25 m, which after investigation of real surfaces (Fig. 2.2) was
found to be representative of the majority of honing grooves. For a desired
honing spacing, S, the value of w in Eq. (3.7) was iterated for to keep the
width of the groove constant at 25 m. In the next step, Fig. 3.8b, this single
groove is mirrored in the x and y directions to give one artificial X, or
periodic texture pattern. Only one, periodic, X-like groove is necessary:
when flow factors are calculated (using the method described in Chapter
2) periodic boundary conditions are applied on this X pattern. In the
final step (Fig. 3.8c), real, measured isotropic plateau roughness (the small
wavelength roughness between the honing grooves) is applied to the surface
to give the final result.
In order to evaluate a range of different surface textures and work to-
wards an optimum, flow factors were calculated and then simulations run
for a range of different values of S, D and . These flow factors and the
results from the simulations are presented in Chapter 5.

3.4 Friction

The calculation of viscous and boundary friction has already been intro-
duced in Chapter 2, whereby Eq. (2.21) and Eq. (2.22) can be solved and
summed to calculate the friction in the contact. However, as outlined ear-
lier a modification must be made to the solution in the inlet region. If
the contact inlet is starved then with the equations presented thus far it
is assumed that before reformation occurs (the point at which a positive
pressure ps begins to form in the lubricant film) and the lubrication regime
becomes full film, striations exist bridging the gap between piston ring and
cylinder liner. This assumption of striations will add slightly to the viscous
friction in the contact. In reality, it is proposed that the lubricant is sitting
on the cylinder liner surface and does not make contact with the ring and
contribute to friction until the reformation point is reached, when r = 0.
Therefore, if the contact is starved any viscous friction contribution before
the node where r first equals zero is neglected.
50 CHAPTER 3. DEVELOPMENT OF A PRCL SIMULATION TOOL

3.5 Wear
The wear rate in the contact is calculated from equation Eq. (3.9), a simple
Archard type wear equation;


kw PCP
w = U (3.9)
H
where w is the wear rate, kw is a wear constant taken as 1.5e-10 and H is
the material hardness, taken as 1200 MPa.

3.6 General solution procedure


The simulation is run as a time dependent problem for a full ring-pack with
three rings. A general overview of the problem is shown in Fig. 3.9. As

Figure 3.9: Overview of the full problem

Fig. 3.9 illustrates, below the oil control ring there is always assumed to be
enough oil on the liner surface, due to splashing and spray of lubricant in
the crankcase and from piston cooling, to fully flood both lands of the oil
control ring. Furthermore, the second inter-ring land, between the lower
compression ring and oil control ring, is always assumed to contain enough
oil to fully flood the lower side of the lower compression ring and upper side
of the oil control ring. This assumption was made due to the work of Vokac
and Tian [47], who investigated this experimentally.
3.7. MODEL INPUT PARAMETERS AND CONVERGENCE 51

However, as the ring-pack slides downwards the amount of oil left behind
by the lower and upper compression rings, hex in Fig. 3.6, is deposited and
recorded on the cylinder liner surface at each specific location. Hence, at
any specific location on the cylinder liner surface, the upper compression
ring only has oil available at its inlet, hin (see Fig. 3.6), that was left behind,
hex , at the same location by the lower compression ring. On the upstroke
the upper compression ring only has the lubricant available (hin ) that it left
behind (hex ) at the same location on the previous downstroke.
For each time step and for each ring, first the degree of filling is calcu-
lated using the procedure described in Fig. 3.7. Following this, the force
balance, Eq. (3.10), can be solved for each ring using the bisection root
finding method to locate the separation between piston ring and cylinder
liner that balances the equation:

FHYD (h) + FCP (h) = FTENS + FGAS (3.10)

where FHYD is the hydrodynamically supported load, FCP is the asperity


contact load, FTENS is the ring tension and FGAS is the gas pressure load.
The hydrodynamically supported load, FHYD , comes from integrating the
hydrodynamic pressure found by solving Eq. (2.18). The asperity contact
load, FCP , is found from integrating the asperity contact pressure, PCP ,
found from the contact mechanics simulation (see Fig. 2.9). FT EN S is the
piston ring tension, calculated from Eq. (3.11) per unit length:

2FT
FT EN S = (3.11)
B
where RT is the ring tension specified by the piston ring manufacturer.
FGAS is the integral over the back face of the ring of either the gas pressure
above or below the ring, whichever is greatest (see Fig. 3.2 and Fig. 3.3).

3.7 Model input parameters and convergence


Table 3.1 gives the input parameters used when the simulation tool is imple-
mented in Chapter 5. The space and time domain mesh sizes were decided
upon after a convergence study was run and the minimum oil film thick-
ness, boundary and viscous friction were compared as the mesh size was
increased. When increasing the mesh resolution by 100 nodes had less than
a 1% effect on the results, the values were considered appropriate to use
in the simulations. The simulation was run through three complete engine
cycles which was enough to ensure that convergence was reached for the
lubricant left behind on the liner surface, a long duration time dependent
effect. Initially, a starting value of 1 m was placed along the entire length
52 CHAPTER 3. DEVELOPMENT OF A PRCL SIMULATION TOOL

Table 3.1: Input parameters used in the model.

Parameter Value
Engine speed 1200 RPM
Engine stroke 0.160 m
Conrod length 0.255 m
Bore diameter 0.130 m
Boundary friction coefficient 0.09
Lubricant density 900 kg/m3
TCR ring tension 37.3 N
2CR ring tension 25.6 N
OCR ring tension 58.9 N
TCR mesh size 300 x 1
2CR mesh size 300 x 1
OCR mesh size 100 x 1
Time steps per engine cycle 720
Mixed lubrication model mesh size 512 x 512
Piston ring elastic modulus 200 GPa
Cylinder liner elastic modulus 117.5 GPa
Piston ring Poissons ratio 0.3
Cylinder liner Poissons ratio 0.3

of the cylinder liner surface. For most of the stroke, this starting value had
no effect on the final solution. However, it did have a small effect on the
quantity of lubricant deposited by the TCR at TDC. This lubricant is swept
up to the top of the cylinder liner and then stays there, so a large initial
starting value of lubricant on the liner surface will increase the height of
lubricant deposited here. Later, in Chapter 5 when the results are analysed
this issue will be discussed again.
Chapter 4

Experimental validation of
the model

In this chapter the work that has been undertaken to validate the simulation
described in the previous section will be presented. Four cylinder liner vari-
ants were investigated in this validation stage and these will be introduced
in the first section of the chapter. The validation consisted of two parts.
The first step of validation was to run full fired single cylinder engine tests
and calculate the Friction Mean Effective Pressure (FMEP). Secondly, re-
ciprocating bench tests were run to measure the friction between a sample
of top compression ring and four cylinder liner variants. Furthermore, an
ultrasonic technique was used to measure the lubricant film thickness in the
PRCL contact during these tests.

4.1 Surfaces under investigation


Four cylinder liner variants have been tested in this work, illustrated in
Fig. 4.1. The surface images presented in this section were acquired with a
Vertical Scanning Interferometer (VSI), one of the measurement techniques
used throughout this work and introduced in Section 2.6.1. The first surface,
denoted as STD55, is a standard production grey cast iron cylinder liner
with a honing angle of 55. The second surface, STD35 is also a grey cast
iron cylinder liner, but with a honing angle of 35 and a slightly rougher
surface finish. The third surface, denoted ANS is a grey cast iron cylinder

liner treated with ANS Triboconditioning . This process is incorporated
into the honing process with the use of a special honing tool and fluid
[48, 49]. A layer of Tungsten Disulphide (WS2 ) is deposited on the surface
and the topography is also modified to give a more run-in like surface finish.
The final surface, denoted PL, is a grey cast iron liner that is plasma

53
54 CHAPTER 4. EXPERIMENTAL VALIDATION OF THE MODEL

(a) STD55. (b) STD35.

(c) ANS. (d) PL.

Figure 4.1: Surface images.

Table 4.1: Roughness parameters for each cylinder liner variant. All rough-
ness parameters are given in microns (m) unless otherwise stated.

Sample Sa Sq Spk Sk Svk M r1 (%) M r2 (%)


STD55 0.454 0.739 0.436 1.040 1.999 15.81 86.21
STD35 0.542 0.782 0.461 1.096 1.867 8.51 82.31
ANS 0.327 0.512 0.408 0.569 1.228 8.22 80.07
PL 0.249 0.554 0.255 0.391 1.196 9.01 83.91

sprayed with a stainless steel and ceramic composite before being honed
with a very small grit size tool which leaves an approximately 100m thick
coating on the surface. This surface is relatively porus with deep holes
in the coating rather than the deep honing grooves of the STD55, STD35
and ANS surfaces, giving it notably different texture. The Sk functional
roughness parameters, based on the Abbott curve (Fig. 4.2), are presented
in Table 4.1.

4.2 Single cylinder engine tests


Engine tests were run at Scanias technical centre in S
odert
alje, Sweden.
During the tests Friction Mean Effective Pressure (FMEP) was calculated
4.2. SINGLE CYLINDER ENGINE TESTS 55

Figure 4.2: Illustration of the Abbott Curve that the Rk functional param-
eters are based on. The terminology Sk (as opposed to Rk) refers to the
same parameters but applied to a surface measurement rather than a profile
measurement.

at various steady state operating conditions, from motored to full load, at


three different speeds. In this section a brief introduction to the theory will
be given, followed by presentation and discussion of the results.

4.2.1 Measurement of FMEP in a fired engine


The term Mean Effective Pressure (MEP) is defined as the work done per
engine cycle divided by the total cylinder volume displaced per cycle. The
advantage of using MEP as a measure of engine performance is that it is
independent of engine displacement or number of cylinders. Therefore, the
results from these single cylinder engine tests can be considered represen-
tative of larger multi-cylinder production engines. Good MEP values have
been established for a range of engine types, therefore it is an effective way
of comparing different engines. Three different forms of MEP are referred
to in this thesis:

1. Indicated Mean Effective Pressure (IMEP) - this is the MEP present


in the combustion chamber; the work done by the combustion gases
on the piston(s).
56 CHAPTER 4. EXPERIMENTAL VALIDATION OF THE MODEL

2. Brake Mean Effective Pressure (BMEP) - this is the MEP as measured


at the flywheel, the work done by the engine.

3. Friction Mean Effective Pressure (FMEP) - this is defined as the differ-


ence between IMEP and BMEP, the total of any losses in the engine.

FMEP, the engine losses, can be calculated by measuring IMEP and


BMEP and subtracting one from the other. IMEP is measured by placing a
pressure transducer in the cylinder head to record the in-cylinder pressure.
By integrating the pressure curve on the resulting P-V diagram, Fig. 4.3,
the work done over the cycle is found. The area between the expansion and

Figure 4.3: Pressure-Volume Diagram for a 4-stroke engine cycle

compression curves (the High Pressure Cycle) gives the work done by the
combustion gases on the piston. The area between the exhaust and intake
curves (the Low Pressure Cycle) must be subtracted from the High Pressure
Cycle, this is the work that the piston must do to expel the exhaust gases
and draw in new air on the intake stroke. This work done, Wi , is divided
by the displaced volume, Vd , to give IMEP:

Wi
IM EP = (4.1)
Vd

BMEP is captured by measuring the torque produced by the engine and


applying the equation;
4T
BM EP = (4.2)
Vd
4.2. SINGLE CYLINDER ENGINE TESTS 57

where T is the torque produced by the engine. As should be apparent, one


limitation of this method is that only the total losses present in the engine
are measured, PRCL friction cannot be differentiated from the total losses.
However, in theory the change in FMEP between tests on different cylinder
liner variants can be observed, and the cylinder liners compared in relation
to one another.
The results for the engine tests are shown in Fig. 4.4. The data has been
non-dimensionalised for confidentiality reasons. It is difficult to draw any

1900 RPM
1500 RPM
1200 RPM
FMEP (-)

STD55
STD35
ANS

50 0 50 100 150 200 250 300 350 400 450


Torque (Nm)

Figure 4.4: FMEP for three engine speeds across a torque sweep for three
cylinder liner variants

firm conclusions from this data as no one cylinder liner clearly outperforms
the other two under all operating conditions, although overall the STD55
liner exhibits the lowest frictional losses. When comparing the experimen-
tally measured difference in FMEP between cylinder liner variants with
the predicted ring-pack losses from simulation, quite often the difference
in FMEP between cylinder liners is greater than the total predicted FMEP
ring-pack losses from simulation. It is also clear that ring pack FMEP is only
a proportion of the total FMEP losses. It is suggested that the likely error
in these experiments is of a similar magnitude to the ring pack FMEP that
these experiments attempt to observe. There are several possible sources
for these errors. This technique is an indirect measure of friction. Two large
numbers are measured, IMEP and BMEP, which are then subtracted from
each other to give FMEP, a smaller number. Therefore, any error in the
measurement of IMEP or BMEP will lead to a greater error, expressed as
58 CHAPTER 4. EXPERIMENTAL VALIDATION OF THE MODEL

a percentage, in the calculation of FMEP. When calculating IMEP, there


are possible sources of error when measuring the P-V diagram. Not only in
the measurement of the in-cylinder pressure, which could be significant, but
also the exact synchronisation of piston TDC, which can be an imprecise
measurement also.
The inclusion of these results in this thesis aims to illustrate the difficul-
ties in evaluating frictional losses from individual components in an engine
when conducting full engine tests. Other, more advanced techniques are
available, such as the measurement of instantaneous IMEP [50, 51, 40] or
the direct measurement of friction with a floating liner [52, 53, 54]. However,
these are not simple tests to perform and require significant modification
to a standard engine. Therefore, after these tests were completed it was
decided to run simplified bench tests to directly measure PRCL friction for
the purpose of validating the model.

4.2.2 Reciprocating bench tests


Collaborative research work was undertaken with E.Y. Avan at the Leonardo
Centre for Tribology at the University of Sheffield. The work is presented
in detail in Paper D and Paper E of this thesis and will therefore not be
discussed in great detail here, rather a summary of the experiments and
results will be given.
The motivation for undertaking this work in collaboration with the
Leonardo Centre for Tribology was in their expertise in the application
of ultrasound to measure oil film thickness in lubricated contacts [55]. The
author of this thesis does not pretend to be an expert in ultrasonic film
thickness measurement and credit for this element of the work must go to
E.Y. Avan [56]. The author of this thesis ran the simulations in Paper D
and Paper E, worked with E.Y. Avan to develop, manufacture and run the
tests at the University of Sheffield and became familiar with the practicali-
ties of ultrasonic film thickness measurement. However, E.Y. Avan is solely
responsible for the processing and evaluation of the ultrasound data.
A Plint TE-77 high frequency reciprocating tribometer, schematically
shown in Fig. 4.5, was used in this work. The machine configuration consists
of sections of piston ring and liner where the ring specimen reciprocates over
the stationary liner sample. The normal load is applied via a spring-balance
through a lever and stirrup mechanism. The normal force is transmitted
directly onto the ring section by means of a needle roller cam follower on a
carrier head and a running plate on the loading stirrup. The whole assembly
is mounted on flexible supports providing for free movement in horizontal
directions, and connected to a stiff force transducer (Kistler type 9203) that
measures tangential force in both directions.
4.2. SINGLE CYLINDER ENGINE TESTS 59

Figure 4.5: Plint TE-77 simulated piston ring-liner contact schematic.

A section of run-in cylinder liner and upper compression ring from the
same heavy duty diesel engine were used to create the contact. The four
cylinder liner surfaces previously introduced in section 4.1 were used in these
tests. Each cylinder liner sample measured 50 mm in length and 20 mm in
width and was cut from a complete cylinder liner with a bore diameter of
130 mm. The piston ring was sectioned into a length of 45 mm. The width
of the piston ring is 3 mm and it has an asymmetric barrel shaped face, as
illustrated previously in Chapter 3.
A lubricant bath was modified to hold the liner specimens. Six grub
screws were used to secure the liner specimen in place; this allows for align-
ment of the liner in both the axial and lateral directions. To retain the
ring section, a special ring holder attached to the carried head was manu-
factured from an original production piston. A ring section was clamped to
the ring holder using two slotted plates either side of the ring holder and a
grub screw in the centre. This clamping system bent the ring section and
allowed it to conform over the liner section. The conformability between
the ring and liner was checked using pressure paper and a good conformal
contact was obtained, as described in detail in Paper D.
For each of the four liner surfaces, tests were run for a range of speeds
and loads. The load was varied from 40 N to 200 N in steps of 20 N and
60 CHAPTER 4. EXPERIMENTAL VALIDATION OF THE MODEL

the speed was varied from 2.5 Hz to 17.5 Hz in steps of 2.5 Hz, giving 63
test points in total (7 speeds and 9 loads).
At this point it is of interest to make a comparison with the real engine
operating conditions. The maximum piston speed at 1200 RPM is 10.5 m/s.
At a peak combustion pressure of 200 bar, if this pressure is assumed to act
entirely on the back of the top compression ring, the load on a section of
ring of the size used in these tests (20 mm x 3mm) would be 1200 N. This
is the maximum speed and load that the top compression ring is subjected
to in the engine. Therefore, the speed in these tests is representative of the
reversal points, but not mid-stroke. Also, the load is representative of most
of the engine cycle, but not around the region of peak combustion pressure.
It would of course be of interest to run at higher speeds and loads, but the
conditions used here were the upper limits of the capabilities of the test
rig used in this study. This highlights the disadvantages of bench tests in
investigating the PRCL contact, however the disadvantages with full engine
tests have been discussed in the previous section - both methods have their
own strengths and weaknesses.
The stroke of the machine was set to 15 mm (the maximum value). The
liner specimen was fully immersed in pure base oil without an additive pack-
age. The oil bath temperature was logged at a stable 22 C throughout the
tests. It is acknowledged that the lubricant temperature is unrepresentative
of real engine running conditions. However, the primary goal of this work
is to accurately and repeatably investigate different liner surface topogra-
phies rather than recreate the exact engine operating conditions. It was
decided that running at a lower temperature, giving a higher viscosity and
therefore higher film thickness would go some way towards compensating
for the lower entraining speeds in the test rig compared to the real engine
operating conditions. In addition, the large quantity of lubricant in the oil
bath ensures that the inlet is always fully flooded allowing for good, accu-
rate comparisons with the numerical model. It also serves in maintaining a
stable temperature of the liner surface during the short tests.
An ultrasound film thickness measurement technique was applied to the
contact which is described in great detail in Paper D and E. The contact
was modelled numerically using the simulation tool previously described in
Chapter 3, but adapted to model the test rig setup illustrated in Fig. 4.5.
The oil transport model was not implemented due to the previously stated
assumption that the ring is always fully flooded in these tests.
In Fig. 4.6, the friction coefficient averaged through the reciprocating
cycle is presented for each of the four surfaces. On each plot both the
experimental values and the simulated values are given. Overall there is
very good correlation between the experimental and simulated values. As
would be expected, the friction coefficient is highest, for all the surfaces,
4.2. SINGLE CYLINDER ENGINE TESTS 61
Avg. Friction Coefficient

Avg. Friction Coefficient


0.11 0.11
0.1 0.1
0.09 0.09

Experiment

Experiment
Simulation

Simulation
0.08 0.08
0.07 0.07
0.06 0.06
0.05 0.05
0.04 0.04
0.03 0.03
0.02 0.02
0.01 0.01
2.5 60 40 2.5 60 40
5.0 100 80 5.0 100 80
Sp7.5 120 ) Sp7.5 120 )
eed10.0
(Hz12.5
140 d (N eed10.0
(Hz12.5
140 d (N
) 15.0
17.5 200
180
160
Loa ) 15.0
17.5 200
180
160
Loa

(a) STD55. (b) STD35.


Avg. Friction Coefficient

Avg. Friction Coefficient

0.11 0.11
0.1 0.1
0.09 0.09
Experiment

Experiment
Simulation

Simulation

0.08 0.08
0.07 0.07
0.06 0.06
0.05 0.05
0.04 0.04
0.03 0.03
0.02 0.02
0.01 0.01
2.5 60 40 2.5 60 40
5.0 100 80 5.0 100 80
Sp7.5 120 ) Sp7.5 120 )
eed10.0
(Hz12.5
140 d (N eed10.0
(Hz12.5
140 d (N
) 15.0
17.5 200
180
160
Loa ) 15.0
17.5 200
180
160
Loa

(c) ANS. (d) PL.

Figure 4.6: Average friction coefficient maps.

at 200 N and 2.5 Hz, i.e. the highest load and lowest speed. The physical
explanation for this is that the entraining speed is insufficient to generate an
oil film to support such a high load. Looking at the rest of the friction map,
as the speed increases, or the load reduces, the average friction coefficient
drops in value. This indicates moving further to the right on the Stribeck
curve as lower load, or higher speed, allows for a thicker oil film to be
generated and the majority of the lubrication becomes either mixed or full
film rather than boundary. The lowest average friction coefficient for all the
surfaces occurs at the lightest load and highest speed; 40 N and 17.5 Hz.
Here the thickest film is observed and the asperity friction contributes the
least to the total friction. Overall, for all the surfaces the simulation slightly
overestimates the friction coefficient. There are a number of reasons for this,
e.g., the viscosity or boundary friction coefficient used in the simulation
could be slightly higher than what it actually was during the experiment.
It is reasonable to assume that the shear heating of the lubricant during the
test could raise the oil temperature and lower its viscosity, thereby reducing
the viscous friction compared to that predicted by the numerical simulation.
Further analysis of this data, including time-resolved friction through the
stroke, can be found in Paper D.
The numerically simulated and ultrasonically measured film thickness
is compared next. Fig. 4.7 presents the numerically and experimentally
62 CHAPTER 4. EXPERIMENTAL VALIDATION OF THE MODEL

5.0 Hz 2.5 Hz 5.0 Hz 2.5 Hz


3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz 3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz
3.0 3.0

2.6 2.6
MOFT (m)

MOFT (m)
2.2 2.2

1.8 1.8
1.4 1.4
1.0 1.0
0.6 0.6
0.2 0.2

40 40
60 60
80 80
Lo

Lo
100 100
120 120
a

a
140 140
d(

d(
160 160
180 TB T 180 S1 S5 S1
N)

N)
TB TB TB S1 S5 S1 S5
200 TB TB 200 S1 S5 S1 S5 S1 S5
B S5
Position Sensors

(a) STD55 Simulation. (b) STD55 Experiment.

5.0 Hz 2.5 Hz 5.0 Hz 2.5 Hz


3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz 3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz
3.0 3.0

2.6 2.6
MOFT (m)

MOFT (m)

2.2 2.2

1.8 1.8
1.4 1.4
1.0 1.0
0.6 0.6
0.2 0.2

40 40
60 60
80 80
Lo

Lo

100 100
120 120
a

140 140
d(

d(

160 160
180 TB T 180 S1 S5 S1
N)

N)

TB TB TB S1 S5 S1 S5
200 TB TB 200 S1 S5 S1 S5 S1 S5
B S5
Position Sensors

(c) STD35 Simulation. (d) STD35 Experiment.

5.0 Hz 2.5 Hz 5.0 Hz 2.5 Hz


3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz 3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz
3.0 3.0

2.6 2.6
MOFT (m)

MOFT (m)

2.2 2.2

1.8 1.8
1.4 1.4
1.0 1.0
0.6 0.6
0.2 0.2

40 40
60 60
80 80
Lo

Lo

100 100
120 120
a

140 140
d(

d(

160 160
180 TB T 180 S1 S5 S1
N)

N)

TB TB TB S1 S5 S1 S5
200 TB TB 200 S1 S5 S1 S5 S1 S5
B S5
Position Sensors

(e) ANS Simulation. (f) ANS Experiment.

Figure 4.7: Film thickness comparison of three liner samples (for sensor
positions refer to Fig. E.1). T refers to the reversal point closest to the
load cell and B refers to the reversal point closest to the reciprocating arm.
4.2. SINGLE CYLINDER ENGINE TESTS 63

obtained film thickness data, for the STD55, STD35 and ANS liner. There
is no data for the PL liner. This is because the ultrasound technique was
not able to measure the film thickness through the plasma coating due to
attenuation of the ultrasound signal in the coating.
In each sub-figure, seven surface plots are presented, one for each of the
sliding speeds, with position on the x and load on the y axes. Only one
sliding direction can be shown, in all cases this is from the left to the right
of the image in Fig. 4.5. The stroke in the other direction, from right to
left, was not considered due to potential anomalies in the ultrasound mea-
surement, caused by cavitation of the lubricant. This effect is discussed in
detail in Paper D. The simulated film thickness maps contain 50 steps for
each stroke. On the z axis the Minimum Oil Film Thickness (MOFT) is
given. At a given sensor position this is the smallest film thickness that oc-
curs between the cylinder liner surface and the piston ring as the ring passes
over the sensor. A deconvolution operation is performed to get the true
minimum rather than the average over the sensor width. This technique
is described in Paper D. In the simulations, the definition of MOFT is the
distance between the mean plane of the surface roughness and the (smooth)
ring profile, as explained in Chapter 3.
All plots show the same trend in film thickness, where the MOFT in-
creases towards the mid-stroke - as the entraining speed increases, and then
reduces - as the ring comes to a halt at the end of the stroke. The film thick-
ness maps, from both simulation and experiment, are all skewed so that the
maximum MOFT occurs just past mid-stroke, where the entraining speed is
greatest. This is due to the time dependent nature of the oil film build-up,
and time dependence is also considered when solving the modified Reynolds
equation during the numerical simulation.
At low speeds the simulated and experimental film thicknesses, for all
three liner surfaces, compare well. According to the ultrasound measure-
ments, the ANS coated liner surface seems to generate a lower film thickness
than the other two liners and this is also what the numerical simulation pre-
dicts. There is a greater difference at higher speeds where measured film
thickness is significantly lower than the value computed by the model. Due
to the fact that the experimentally measured and numerically simulated
friction compare very well at this point, see Fig. 4.6, it is suggested that
this difference occurs because the ultrasonic film thickness measurement is
underestimating the film thickness. At high speeds more cavitation occurs
in the trailing portion of the contact, as shown by Dellis and Arcoumanis
[32], and also visible in the results of the numerical model introduced in
Chapter 3. Unfortunately, the effect of cavitation on the speed of sound
and density in a medium confined between two surfaces in relative motion
has, to the authors knowledge, not yet been researched. However, these
64 CHAPTER 4. EXPERIMENTAL VALIDATION OF THE MODEL

properties are needed to translate the measured reflection coefficient into


film thickness and the uncertainty in these values could lead to the errors
in film thickness measurement at high speed.
Chapter 5

Optimization of texture

In this chapter results from the PRCL simulation model described in Chap-
ter 3 will be presented and discussed. The parametric study undertaken will
be introduced along with an illustration of some of the calculated flow fac-
tors for these cases. In the second section, the results from the simulations
will be presented. Firstly one case will be presented in detail with a presen-
tation of the calculated film thicknesses, friction, oil consumption risk and
wear. Following on from this there is discussion on the effect of the three
varied parameters - honing angle, depth and spacing. Finally, the chapter
concludes with a discussion on the findings of the parametric study and the
potential reduction in frictional losses by modifying cylinder liner surface
texture.

5.1 Parametric study of texture parameters


In section 3.3 a method was introduced for generating artificial surface tex-
ture with varying depth (D), spacing (S) and angle (). The texture opti-
misation investigation presented in this chapter is conducted in two stages,
as presented in Fig. 5.1. In the first stage the honing depth (D) and spac-
ing (S) is kept constant at 1 m and 200 m respectively - these values
are broadly representative of the current honing structure as illustrated in
Fig. 2.2. The honing angle is varied between 30 and 150 - the current
liner is honed at an angle of 50 . Fig. 5.2 illustrates the definition of hon-
ing angle used in this thesis. Changing the honing angle of a real cylinder
liner between these limits is entirely possible with the current manufac-
turing process. Simply, the ratio of rotational speed to axial speed of the
honing machine is altered to change the honing angle. Fig. 5.3 illustrates
the mean asperity contact pressure at the extremes of honing angle, 30 and
150 . It is clear from Fig. 5.3 that the stiffness of the surface is identical for

65
66 CHAPTER 5. OPTIMIZATION OF TEXTURE

Figure 5.1: Test cases for texture optimisation.

both cases. This is entirely as expected as 150 honing angle is exactly the
same as 30 , just rotated by 90 . This validates a small part of the texture
generation model, the result is as expected. Further, Fig. 5.4 illustrates
the Reynolds flow factors for the same two surfaces, 30 and 150 honing
angle. The Reynolds flow factors also illustrate this rotation of the surface
by 90 . In the case of the surface with 30 honing angle, the Poiseuille flow
(a11 and a22 ) is restricted more in the entraining direction (a11 ) as the tex-
ture, or honing grooves, cross the surface in a more transverse, rather than
longitudinal, manner. This is illustrated with the a11 flow factor reducing
towards zeros earlier, or at a greater film thickness, than the flow factor
describing the change in longitudinal flow, a22 . It is clear in the figure that
the Poiseuille flow factors for the 150 honing angle surface are the exact
opposite, meaning that the flow is less restricted in the entraining direction
due to the 90 rotation of the surface, giving the texture a more longitudinal

Figure 5.2: Illustration showing the definition of honing angle.


5.1. PARAMETRIC STUDY OF TEXTURE PARAMETERS 67
Mean Asperity Contact Pressure, PCP (MPa)

200

150

= 30
= 150
100

50

0
101 100 101 102
T (m)
Separation, h

Figure 5.3: Comparison of mean asperity contact pressure with 30 and


150 honing angle.

1.5
Normalised Flow Factor

1
a11 - 30
a22 - 30
0.5 b12 - 30
a11 - 150
a22 - 150
0
b12 - 150

0.5
101 100 101 102

Separation, hT (m)

Figure 5.4: Comparison of Reynolds equation flow factors with 30 and 150
honing angle.
68 CHAPTER 5. OPTIMIZATION OF TEXTURE

lay. Examining the b12 flow factor, describing the effect of the honing on
Couette flow, in both cases as the film thickness reduces the flow increases
with respect to a smooth surface. As discussed previously in Chapter 2, this
is due to the non-stationary texture dragging lubricant through the contact
with it. The flow factors show that the 30 honing angle surface, with the
more transverse roughness, is more effective at dragging lubricant through
the contact.
In the second stage of texture optimisation the honing angle is kept
constant at = 50 (representative of the current honed surface, Fig. 2.2)
and both the honing depth (D) and spacing (S) are varied according to the
test matrix in Fig. 5.1. The asperity contact pressure for a honing depth
(D) of 0.5 m and 3.0 m is shown in Fig. 5.5 with a constant honing
spacing (S) of 200 m. An important feature of Fig. 5.5 should be drawn
Mean Asperity Contact Pressure, PCP (MPa)

200

150

D = 0.5 m
D = 3.0 m
100

50

0
102 101 100 101 102
T (m)
Separation, h

Figure 5.5: Comparison of mean asperity contact pressure with 0.5 m and
3.0 m honing depth.

to the readers attention. The gradient of the two stiffness curves are the
same (the log xaxis creates a false impression of a shallower gradient of
the 0.5 m curve). However, the value of the average separation, h T , where
the surfaces come into contact, is very different. Increasing the depth of the
honing grooves, while everything else remains constant, has only a very small
effect on the surface stiffness, but it clearly has a huge effect on the average
separation. This is caused by the mean line (refer back to Fig. 2.11) through
5.2. EVALUATION OF RESULTS 69

the surface moving further away from the flat plateau surface as the honing
grooves deepen. The corresponding flow factors for honing depths of 0.5
m and 3.0 m are shown in Fig. 5.6. The flow factors behave in a similar

1.5 a11 - 0.5 m


Normalised Flow Factor

a22 - 0.5 m
b12 - 0.5 m
1 a11 - 3.0 m
a22 - 3.0 m
b12 - 3.0 m
0.5

0.5
102 101 100 101 102
T (m)
Separation, h

Figure 5.6: Comparison of Reynolds equation flow factors with 0.5 m and
3.0 m honing depth.

manner. While they do not have exactly the same shape, the dominant
effect visible on the graph is this offset in the horizontal direction caused by
the shift in mean line through the cylinder liner surface. Therefore, caution
must be taken when evaluating the average separation of the surfaces - even
though the average separation could appear very different the lubrication
regime, amount of asperity contact and friction could be exactly the same.
The difference in average separation is just an artifact of the shift in the
mean line of the cylinder liner surface. This is not to say that a change
in honing depth will have no effect on the lubrication regime of the PRCL
contact, simply that the results must be evaluated with this in mind.

5.2 Evaluation of results


In this section one case will be evaluated in detail to present typical results of
film thickness, oil filling, friction, wear and oil consumption risk. Then in the
following sub-sections the effect of the three parameters, honing angle (),
honing depth (D) and honing spacing (S) are evaluated. In the following
figures and discussion, the standard case, = 50 , D = 1 m and S = 200
70 CHAPTER 5. OPTIMIZATION OF TEXTURE

m will be evaluated, these parameters are reasonably typical of the honing


structure of a current production cylinder liner (as illustrated in Fig. 2.2).
The ring nomenclature first introduced in Fig. 1.1 will be used throughout
the remainder of this thesis. To recap:

Upper Compression Ring - TCR

Lower Compression Ring - 2CR

Oil Control Ring - OCR

Film Thickness
Fig. 5.7 shows the minimum average film thickness for the three piston rings.
At firing TDC the TCR has the lowest oil film thickness due to the high gas

6
Minimum average film thickness (m)

TCR
2CR
5
OCR

0
0 90 180 270 360 450 540 630 720

Crank Angle ( )

Figure 5.7: Minimum average oil film thickness vs. crank angle for D = 1.0
m, S = 200 m and = 50 .

loading on the back face of the ring in comparison to the other two rings (see
Fig. 3.3). As the rings move away from TDC the film thickness for all three
rings increases due to higher entraining speed (Fig. 3.1), lower load (Fig. 3.3)
and lower temperature (Fig. 3.4), leading to higher viscosity (Eq. (3.2)). The
TCR always has a lower oil film thickness than the 2CR. This is for two
reasons - one, it has a higher loading for most of the first 180 crank angle
(CA). Secondly, due to the oil availability model implemented in this work,
on the downstrokes it only has the lubricant available to it that the 2CR left
5.2. EVALUATION OF RESULTS 71

trailing behind it. Therefore, the film thickness of the TCR cannot exceed
for any significiant duration of time that of the 2CR. However, it is possible
for the 2CR to have a significantly larger film thickness than the OCR due
to the assumption that the second ring land always contains enough oil to
fully flood the 2CR. The OCR has a greater film thickness than the TCR
around the reversal points due to lighter gas loading on the ring. However,
at mid-stroke it doesnt generate as much lift due to a small converging gap
(Fig. 3.5) and relatively high ring tension. The saddle shape that can be
observed in the mid-stroke region for both the TCR and 2CR is caused by
the temperature profile of the cylinder liner surface. The cylinder liner is
cooler towards the bottom of the stroke due to the geometry of the cooling
jacket around the cylinder liner, this in turn increases the viscosity of the
lubricant which leads to a larger minimum oil film thickness just after mid-
stroke, rather than at mid-stroke, the point of highest entraining velocity.
The minimum film thickness at the BDC reversal point for all three rings
occurs slightly after 180 CA when the ring-pack is stationary. This is due
to the time dependent nature of the problem and the oil squeeze effect.
After BDC reversal at 180 CA the ring-pack begins the exhaust stroke,
travelling in the upwards direction. The TCR, although less heavily loaded
than it was on the expansion stroke, has a slightly lower oil film thickness.
This is because now that the ring is travelling in the upwards direction it
only has the lubricant to run on that it left behind on the previous expansion
stroke (0-180 CA), and therefore cannot for any length of time exceed the
film thickness that it left behind on the downstroke. The 2CR is in a similar
situation and cannot exceed the film thickness that the TCR left behind as
they both travel upwards. At around 210 CA the film thickness of the
scraper ring fluctuates, dropping and then increasing again. This is because
at this crank angle it is running over the location on the cylinder liner
where the TCR reversed. As the TCR stops and reverses, due to a low or
zero entraining velocity, the lubricant film under the ring collapses and the
ring experiences a low minimum oil film thickness, scraping the lubricant
on the liner surface ahead of it instead. Therefore, as it reverses away on
the upstroke it leaves behind it an oil film on the liner surface that has a
high point (the lubricant the ring was scraping in front of it), followed by
a low point (the minimum oil film thickness). As the 2CR runs over this
varying oil film on the liner surface it creates the fluctuation in minimum oil
film thickness that can be seen in Fig. 5.7. The OCR has a similar oil film
thickness on the expansion stroke as it is always running in a fully flooded
condition with a symmetric ring profile and little gas loading, therefore the
oil film thickness does not vary significantly between strokes.
For a period of time around non-firing TDC reversal (360 CA) the film
thickness stays reasonably constant for both the TCR and 2CR. This is
72 CHAPTER 5. OPTIMIZATION OF TEXTURE

due to the lubrication regime being mixed, heading towards boundary, due
to a severe lack of lubricant availability. Most of the load is supported by
asperity interaction and so sliding speed has little effect on film thickness.
Soon after reversal heading into the intake stroke (360-540 CA) the film
thickness increases rapidly for both the 2CR and TCR. The film thickness
is greater than on the previous down stroke (expansion, 0-180 CA) due
to there being less loading on the piston rings. Again, as was the case for
the expansion stroke, the TCR has a slightly lower oil film thickness than
the 2CR due to the oil availability model limiting the amount of lubricant
available on the liner. This limits the lubricant available to the TCR to
that which has passed by the 2CR.
On the following upstroke (compression, 540-720 CA) the TCR has a
marginally lower oil film thickness than the previous downstroke and as
before the 2CR runs on the lubricant that the TCR leaves behind it. The
same oscillation as on the exhaust stroke is visible at around 570 CA as
the 2CR runs over the reversal point of the OCR.

Frictional power losses


In Fig. 5.8 the friction mean effective pressure (FMEP) is calculated for
each ring and plotted against crank angle. Friction mean effective pressure

TCR
80 2CR
OCR

60
FMEP (kPa)

40

20

0
0 90 180 270 360 450 540 630 720

Crank Angle ( )

Figure 5.8: Friction Mean Effective Pressure (FMEP) vs crank angle for D
= 1.0 m, S = 200 m and = 50 .
5.2. EVALUATION OF RESULTS 73

is calculated from Eq. (5.1),


U (fhyd + fbd )
F M EP = 2 . (5.1)
d2 LN
Plotting the results as FMEP gives a more representative idea of where the
greatest power losses occur rather simply the greatest frictional values. For
example, a high friction force at TDC is then multiplied by a sliding speed
of 0 m/s to give a power loss of zero. It is irrelevant what the friction force
is if there is no entraining motion, no power can be lost. Fig. 5.8 illustrates
that the greatest power losses occur from the TCR from just after firing
TDC to around 60 CA, not at TDC due to the lack of entraining motion.
The reader may be more familiar with viewing plots where the highest
friction occurs at the reversal points, such as Fig. E.12 in Paper E. It must
be remembered that these plots show friction force, rather than friction
power, as presented here. The high losses here occur due to the extremely
high gas loading (Fig. 3.3) that demand a very small minimum average oil
film thickness (Fig. 5.7) to support the load, with a significant quantity of
asperity contact. Furthermore, the TCR has very little lubricant available
to lubricate itself - the only lubricant present on the liner above 2CR reversal
is the lubricant that the TCR sweeps up itself.
An interesting phenomenon that is clearly visible in these plots is the
twin peaks in the TCR FMEP after firing-TDC reversal. There is a clear
explanation for this. On the compression stroke (540-720 CA) the 2CR
pushes a wedge of oil in front of itself. As it reverses it leaves this peak
of oil on the liner surface. This is clearly visible in a few pages time in the
middle-left sub-plot of Fig. 5.10. As the TCR reverses it initially runs in
a highly starved mixed lubrication condition. However, it soon runs over
this lubricant peak left at the 2CR reversal point and temporarily has a
surplus of lubricant to run in a more full film lubrication regime, with a
corresponding reduction in FMEP. Soon it has run over and passed this
deposited lubricant peak and the film thickness reduces again leading to
the second peak in FMEP. As the TCR moves further down the liner the
quantity of lubricant left by the 2CR increases and the TCR enters the
hydrodynamic lubrication regime.
At 360 CA non-firing TDC the TCR and 2CR experience considerably
less frictional power losses than at 0 CA firing TDC. This is due to signifi-
cantly less gas pressure acting on the rings. In the mid-stroke region of the
intake and compression strokes the OCR gives the greatest contribution to
friction. This is due to the relatively low oil film thickness (see Fig. 5.7)
leading to high viscous shear of the lubricant. The mean loss for each ring
is given in Table 5.1. The TCR contributes the greatest frictional power
loss however the OCR is close behind with the 2CR contributing less than
74 CHAPTER 5. OPTIMIZATION OF TEXTURE

half of the losses of the TCR. The test condition is 1200 RPM and 50%
Table 5.1: Mean FMEP values for each piston ring.

TCR 5.77 kPa


2CR 2.45 kPa
OCR 4.28 kPa

load. If the load on the engine was reduced then correspondingly the gas
pressure on the rings would also reduce and the TCR friction, especially
around TDC, would be lower. This would lead to the OCR contributing
a greater percentage to the overall friction as it is not affected, or affected
to a very small extent, by gas loading. Conversely, at full load the TCR
contribution to friction will be even greater.

Wear
The wear rate as a function of crank angle, calculated using Eq. (3.9),
is illustrated in Fig. 5.9. The wear rate is shown for each of the piston

8
TCR
2CR
OCR
6
Wear rate (nm/h)

0
0 90 180 270 360 450 540 630 720

Crank Angle ( )

Figure 5.9: Wear rate vs crank angle for D = 1.0 m, S = 200 m and
= 50 .

rings, rather than for the cylinder liner surface. The cylinder liner surface
is made of a much softer material than the piston rings (cast iron, rather
5.2. EVALUATION OF RESULTS 75

than chromium) and so experiences a total rate of wear much greater than
the piston rings. However, piston ring wear is concentrated over a much
smaller area than cylinder liner wear and therefore becomes much more of
a critical issue than cylinder liner wear. Furthermore, if the TCR loses
its asymmetric barrel-like profile then it can end up scraping oil upwards,
leading to increased oil consumption. Therefore, interest is very much in
evaluating piston ring, rather than cylinder liner, wear.
For most of the stroke the TCR is running in the full film lubrication
regime and therefore does not experience any wear. However, due to the
high load the TCR experiences wear at firing TDC. The SCR runs in full
film lubrication for all but a tiny portion of the stroke at TDC and therefore
experiences extremely small amounts of wear. The OCR, with a narrow
width and high ring tension, but not exposed to such high gas loading, runs
consistently in the mixed lubrication regime and undergoes moderate wear
throughout the stroke. Table 5.2 gives the average wear rate throughout the
stroke for each ring. The wear rates seem reasonable. Herbst [57] presented

Table 5.2: Mean wear rates for each piston ring.

Top Ring 0.20 nm/h


Scraper Ring 0.01 nm/h
Oil Control Ring 0.62 nm/h

experimental data of top ring wear in a heavy duty diesel engine. For
50% load and 1200 rpm, the same conditions investigated here, the wear
rate for the TCR was around 0.5 nm/h, the same order of magnitude as
calculated here. This is a crude comparison, however it is in the right area
and acceptable when the purpose of this work is to compare different surface
textures, rather than calculate absolute values for wear.

Oil filling
It is of particular interest to visualise the degree of oil filling throughout the
width of the contact at different locations throughout the engine cycle. This
gives a clear illustration of the quantity of lubricant available to the contact.
Across the following pages are some illustrations of oil filling throughout the
ring pack for each of the three piston rings. In Fig. 5.10 - Fig. 5.13 the sub-
plots on the left hand side show the ring profiles, average separation from
the cylinder liner surface (y = 0 on the plots) and the lubricant present in
the contact. If the contact is fully flooded (r = 0, and therefore allowing for
a hydrodynamic pressure to be generated) then the gap between the piston
76 CHAPTER 5. OPTIMIZATION OF TEXTURE

ring is entirely full of lubricant (shown in orange in these figures). However,


if the contact is starved or cavitating (r > 0) then the gap is only partially
filled with lubricant. In these plots the lubricant is visualised as sitting on
the cylinder liner surface, however in reality striations could form bridging
the gap between the piston ring and cylinder liner. The sub-plots on the
right hand side of the figure show the hydrodynamic and asperity contact
pressure across the ring width.
Fig. 5.10 illustrates the degree of lubricant filing for 5 CA, on the ex-
pansion stroke. The piston rings are moving to the left with a velocity of 0.7
m/s. Both the TCR and 2CR have deposited the lubricant that they were
previously scraping on to the cylinder liner surface. In the case of the 2CR,
this is the lubricant that the TCR will run over in a few moments time. The
TCR is running in a relatively starved condition with its inlet less than half
filled as it only has available to run on the lubricant that it previously left
behind on the compression stroke (540-720 ). Both inlets of the OCR and
the 2CR are fully flooded as per the boundary conditions in Fig. 3.9. The
pressure profiles show that the TCR is running in the mixed lubrication
regime, albeit with the majority of the load supported by hydrodynamic
pressure. The 2CR is running entirely in the full film regime while the OCR
has an almost equal contribution from asperity and hydrodynamic pressure.
Moving on to Fig. 5.11, 90 CA, the ring-pack is now at mid-stroke
travelling downwards, from right to left of the figure, at 10.1 m/s. Both the
TCR and SCR are in the fully hydrodynamic lubrication regime whilst the
OCR is still just in the mixed lubrication regime with a small amount of
asperity contact pressure.
In Fig. 5.12 the ring-pack is shown just after reversal at BDC; 185 CA,
now travelling at 0.7 m/s in the opposite direction, from left to right in the
figure. The TCR and SCR are still running in the fully hydrodynamic lu-
brication regime, while the OCR is running in the mixed lubrication regime.
with a relatively equal split in load carrying between hydrodynamic and as-
perity contact pressure, as was the case at 5 CA. Similar to the phenomenon
at TDC, the TCR leaves behind the lubricant that it was scraping on the
liner surface which the 2CR will shortly run over. The 2CR, having reversed
direction, no longer has a fully flooded inlet but instead initially runs on
the lubricant that it left behind on the upstroke before running on the lu-
bricant that the TCR leaves behind. The OCR is still fully flooded due to
the second ring land fully flooded assumption introduced earlier.
Finally, Fig. 5.13 shows the ring at mid-stroke again, 270 CA, like
Fig. 5.11 however now the ring-pack is travelling in the opposite direction
towards the right of the figure and the degree of oil filling differs accordingly
at a velocity of 10.1 m/s. As was the case at 90 CA the TCR and 2CR
are running in the full film regime and the OCR is in the mixed lubrication
5.2. EVALUATION OF RESULTS 77

U = 0.7 m/s
TCR Oil Filling TCR Pressure
10 300
PHYD
Ring height (m)

Pressure (MPa)
PCP
200
6

4
100
2

0 0
0 1 2 0 1 2
Ring width (mm) Ring width (mm)
2CR Oil Filling 2CR Pressure
15 10
Ring height (m)

8
Pressure (MPa)

10
6

4
5
2

0 0
0 1 2 0 1 2
Ring width (mm) Ring width (mm)
OCR Oil Filling OCR Pressure
4 15
Ring height (m)

Pressure (MPa)

3
10
2
5
1

0 0
0 1 2 3 0 1 2 3
Ring width (mm) Ring width (mm)

Figure 5.10: Oil filling at 5 CA.

regime.

Oil consumption risk


As mentioned in the introduction, the degree of oil consumption is an im-
portant parameter to evaluate when optimising the PRCL contact. There
are many factors and mechanisms that can lead to oil consumption in an
engine and it is out of the scope of this work to capture them all and come
78 CHAPTER 5. OPTIMIZATION OF TEXTURE

U = 10.1 m/s
TCR Oil Filling TCR Pressure
10 4
PHYD
Ring height (m)
8

Pressure (MPa)
3 P
CP
6
2
4
1
2

0 0
0 1 2 0 1 2
Ring width (mm) Ring width (mm)
2CR Oil Filling 2CR Pressure
15 10
Ring height (m)

Pressure (MPa)
10
6

4
5
2

0 0
0 1 2 0 1 2
Ring width (mm) Ring width (mm)
OCR Oil Filling OCR Pressure
4 15
Ring height (m)

Pressure (MPa)

3
10
2
5
1

0 0
0 1 2 3 0 1 2 3
Ring width (mm) Ring width (mm)

Figure 5.11: Oil filling at 90 CA.

up with an absolute value for oil consumption. However, an attempt will


be made to assess how changing surface texture might at least influence oil
consumption. Initially, the standard case, = 50 , D = 1 m and S =
200 m will be evaluated. Fig. 5.14 shows the quantity of lubricant left
behind on the liner surface on the two downstrokes, intake and expansion,
by the TCR. In the figure the piston ring-pack is moving from right to left.
The step-like profile of the deposited lubricant on the cylinder liner surface
is just an artifact of the time steps (refer back to Table 3.1, 720 steps per
5.2. EVALUATION OF RESULTS 79

U = 0.7 m/s
TCR Oil Filling TCR Pressure
10 4
PHYD
Ring height (m)

Pressure (MPa)
3 PCP
6
2
4
1
2

0 0
0 1 2 0 1 2
Ring width (mm) Ring width (mm)
2CR Oil Filling 2CR Pressure
15 10
Ring height (m)

8
Pressure (MPa)

10
6

4
5
2

0 0
0 1 2 0 1 2
Ring width (mm) Ring width (mm)
OCR Oil Filling OCR Pressure
4 15
Ring height (m)

Pressure (MPa)

3
10
2
5
1

0 0
0 1 2 3 0 1 2 3
Ring width (mm) Ring width (mm)

Figure 5.12: Oil filling at 185 CA.

engine cycle or one step per degree crank angle). The mean height of lubri-
cant at the exit of the PRCL contact, hex , is deposited on the cylinder liner
surface at the axial positions that the piston ring has slid over since the
previous time step. The quantity of lubricant displayed in Fig. 5.14 is that
which is left above the plateau surface, so any lubricant below this height
(effectively residing in the honing grooves) is not shown. This explains why
there is an extremely small quantity of lubricant on the liner towards the
80 CHAPTER 5. OPTIMIZATION OF TEXTURE

U = 10.1 m/s
TCR Oil Filling TCR Pressure
10 4
PHYD
Ring height (m)
8

Pressure (MPa)
3 PCP
6
2
4
1
2

0 0
0 1 2 0 1 2
Ring width (mm) Ring width (mm)
2CR Oil Filling 2CR Pressure
15 10
Ring height (m)

Pressure (MPa)
10
6

4
5
2

0 0
0 1 2 0 1 2
Ring width (mm) Ring width (mm)
OCR Oil Filling OCR Pressure
4 15
Ring height (m)

Pressure (MPa)

3
10
2
5
1

0 0
0 1 2 3 0 1 2 3
Ring width (mm) Ring width (mm)

Figure 5.13: Oil filling at 270 CA.

top of the stroke (the right hand side of the figure). The TCR is running in
the boundary or mixed lubrication regime, with asperity contact, scraping
lubricant down the liner (to the left). The lubricant in the honing grooves
is not plotted because it is the authors opinion that it does not contribute
significantly to the oil consumption risk. For instance, if there is oil in a 1
m deep honing groove, or three times the quantity of lubricant in a 3 m
deep honing groove, it does not necessarily mean that there is three times
5.2. EVALUATION OF RESULTS 81

the oil consumption risk. The lubricant in the bottom of the 3 m honing
groove could be sufficiently sheltered from evaporation risk that it should
not be considered for the purposes of oil consumption risk analysis.
An interesting and important feature of Fig. 5.14 is the large quantity
of oil (over 2 m) at TCR TDC reversal on both the firing and non-firing
reversals. This is caused by the lubricant scraped up on the compression
and exhaust strokes being dumped at TDC, as illustrated in Fig. 5.10. This
creates a high risk of oil consumption as it is a significant height of lubricant
above the plateau surface in close proximity to where combustion occurs.
As discussed in section 3.7, the quantity of lubricant deposited here, at TDC
TCR reversal, is somewhat sensitive to the starting value of lubricant de-
posited on the cylinder liner surface. However, when evaluating the relative
differences in oil consumption risk between different surface textures, it is
suggested that this source of error is acceptable.
As the ring moves towards mid-stroke and enters the hydrodynamic
lubrication regime then more lubricant is left behind on the surface. There
is more lubricant deposited on the intake stroke than the expansion stroke
due to less loading on the piston ring from lower gas pressure, leading to a
greater minimum oil film thickness and greater lubricant deposited behind
the ring on the liner surface. This lubricant is at risk from evaporating
Average film thickness left on liner surface (m)

2.5
Intake Stroke
Expansion Stroke
2

1.5

0.5

0
100 120 140 160 180 200 220 240
Axial position on cylinder liner (mm)

Figure 5.14: Height of lubricant left behind on liner vs. axial position on
liner for D = 1.0 m, S = 200 m and = 50 .
82 CHAPTER 5. OPTIMIZATION OF TEXTURE

off the surface and leading to increased oil consumption. By evaluating the
quantity of lubricant left behind on the liner as the piston moves downwards
an indication of the risk of oil consumption can be obtained.
There are other mechanisms with which lubricant can be transported
into the combustion chamber and lead to oil consumption. If the ring moves
in the piston groove then lubricant can travel around the back side of the
piston ring. In particular, when the pressure below the ring is greater than
that above it (see Fig. 3.3) then blow-back, the flow of gases back in to the
combustion chamber, can occur. This blow-back can transport oil upwards
past the rings into the combustion chamber and will not be captured in this
analysis. Furthermore, this oil transported behind the piston ring can aid
in the lubrication of the PRCL contact. It is the authors opinion that in
reality the top ring is not as starved as is calculated in this model due to
additional lubricant provided through this route.

5.2.1 The effect of honing angle


In this section the effect of honing angle will be presented and discussed.
The TCR is chosen for further analysis of oil film thickness for two reasons
- it is the greatest contributor to friction and it has rather an equal mix of
lubrication regimes, from boundary to full film. The minimum average oil
film thickness for the TCR for D = 1 m, S = 200 m and three variations
in , 30 , 90 and 130 , is presented in Fig. 5.15. The figure is designed
to illustrate that there is very little difference in minimum average film
thickness as the honing angle is varied. However, on closer inspection, for
most of the cycle the shallower, or smaller, the honing angle, the greater the
minimum average oil film thickness. At mid-stroke the values of minimum
oil film thickness for all honing angles are incredibly similar, however in
Fig. 5.16, a close-up of BDC, where minimum oil film thickness is lower, the
difference becomes more apparent. At TDC where the lubrication regime is
predominantly boundary then any effect of honing angle disappears. This is
because a significant proportion of the load is carried by asperity contact and
the stiffness curves are extremely similar for all honing angles, as illustrated
in Fig. 5.3. It is also clear from Fig. 5.4 that at large separations the
surfaces behave the same. The greatest effect is seen at small separations,
but before asperity contact occurs. Only three honing angles are shown for
clarity, 30 , 90 and 130 , however the other curves slot in place in order of
honing angle. Table 5.3 gives the FMEP for each ring for 30-130 . The
table shows that the friction is reduced with a reduction in honing angle,
caused by the increase in minimum average oil film thickness illustrated in
Fig. 5.16.
The shallower honing angles give a more transverse texture for the sliding
5.2. EVALUATION OF RESULTS 83

5
Minimum average film thickness (m)

= 30
= 90
4 = 130

0 90 180 270 360 450 540 630 720



Crank Angle ( )

Figure 5.15: Comparison of TCR minimum average oil thickness with vary-
ing for D = 1.0 m and S = 200 m.

1.56
Minimum average film thickness (m)

= 30
1.54 = 90
= 130
1.52

1.5

1.48

1.46

1.44

1.42
181 182 183 184 185 186

Crank Angle ( )

Figure 5.16: Close-up of TCR BDC minimum average oil thickness with
= 30 , 90 and 130 for D = 1.0 m and S = 200 m.

piston ring. This, in turn, drags more lubricant into the contact and when
the minimum oil film thickness becomes small and the contact is starved, al-
84 CHAPTER 5. OPTIMIZATION OF TEXTURE

Table 5.3: Mean FMEP values for each piston ring and honing angle, all
values in kPa.

1CR 2CR OCR Total


30 5.77 2.44 4.08 12.29
50 5.77 2.45 4.28 12.50
70 5.96 2.47 4.74 13.17
90 6.47 2.54 6.89 15.90
110 6.96 2.65 7.10 16.71
130 7.86 2.60 7.80 18.26

lows the oil film thickness to be ever so slightly larger which delays the onset
of asperity contact and mixed lubrication with the increased friction that
this brings. With larger honing angles and more longitudinal roughness, the
lubricant can be pushed out of the contact in the entraining direction and
the contact is not as effective at generating hydrodynamic pressure, leading
to a lower oil film thickness and higher FMEP. Correspondingly, the wear
rate is also lower for the lower honing angles as illustrated in Table 5.4. The

Table 5.4: Mean wear rates for each piston ring and honing angle, all values
in nm/h.

1CR 2CR OCR Total


30 0.192 0.000 1.070 1.262
50 0.193 0.000 1.232 1.425
70 0.214 0.002 1.647 1.863
90 0.270 0.007 3.665 3.942
110 0.312 0.018 4.285 4.715
130 0.387 0.015 4.464 4.866

wear rate is as would be expected, with a lower wear rate for the smaller
honing angles which also have the lowest FMEP. This supports the theory
that the lower friction is due to the postponing of asperity contact at the
reversal points, rather than a reduction in full film viscous friction.
The final performance indicator that should be evaluated is oil consump-
tion. Therefore, the oil left behind on the intake and expansion stroke is
plotted for = 30 , 90 and 130 in Fig. 5.17. Towards the top of the stroke,
when the TCR is running in the mixed or boundary lubrication regime, the
lubricant deposited on the liner surface is extremely similar. However, as lift
5.2. EVALUATION OF RESULTS 85

Int. Stroke - = 30 Int. Stroke - = 90


Average film thickness left behind on liner surface (m)

Int. Stroke - = 130 Exp. Stroke - = 30


Exp. Stroke - = 90 Exp. Stroke - = 130
2

1.5

0.5

0
100 120 140 160 180 200 220 240
Axial position on cylinder liner (mm)

Figure 5.17: Height of lubricant left behind on liner vs. axial position on
liner for = 30 , 90 and 130 .

off occurs the shallower honing angle leaves slightly more lubricant behind,
due to the improved hydrodynamics discussed previously. Therefore, with
shallower honing angles the risk of oil consumption might increase slightly,
but this would probably be insignificant due to the increase in deposited
lubricant taking place a distance from TDC, far away from the location of
combustion and where the cylinder liner surface is hottest.

5.2.2 The effect of honing depth and spacing


Unlike with honing angle, the minimum average oil film thickness is not
plotted here due to the irrelevance of comparing minimum average separa-
tion for surfaces that have varying topography. This issue was discussed
previously in section 5.1. Instead, the total FMEP will be compared for
varying S and D. Fig. 5.18 shows a power loss map for this varying spac-
ing and honing. The figure shows that while 0.5 m and 1.0 m honing
depth give similar results of FMEP, increasing the honing depth further
to 2-3 m leads to a significant increase in friction. This is caused by a
large increase in boundary friction after TDC for the TCR, as illustrated in
86 CHAPTER 5. OPTIMIZATION OF TEXTURE

Ring-pack FMEP (kPa) 40

30

20

10
300
3
250
2
200
1
1500.5
Honing Spacing (m) Honing Depth (m)

Figure 5.18: Map of FMEP for full ring-pack with varying honing spacing
and depth.

Fig. 5.19. The explanation for this increase in boundary friction is that the

140
D = 1.0 m
120 D = 2.0 m
Boundary FMEP (kPa)

100

80

60

40

20

0
0 10 20 30 40 50 60 70 80 90

Crank Angle ( )

Figure 5.19: Illustration of the boundary friction contribution to TCR


FMEP for the first 90 CA with S = 150 m and D = 1-2 m.

deeper honing grooves form bigger channels that allow lubricant to leak out
5.2. EVALUATION OF RESULTS 87

of the contact and less effectively generate hydrodynamic pressure. With


deep honing grooves the contact cannot as effectively form a hydrodynamic
film and the film thickness is reduced. This effect is illustrated in Fig. 5.20.
The figure shows the hydrodynamic pressure profile for 1.0 and 2.0 m deep

120
D = 1.0 m
Hydrodynamic Pressure (MPa)

100 D = 2.0 m

80

60

40

20

0
0 0.5 1 1.5 2 2.5
Contact width (mm)

Figure 5.20: Difference in hydrodynamic pressure at 30 CA with D = 1-2


m for the TCR. The contact inlet is at the left side of the plot.

honing grooves. Due to the deep honing grooves draining lubricant away
from the contact, the 2 m case cannot as effectively form a hydrodynamic
film. This is illustrated by the shallower gradient of pressure build-up in the
inlet region. The minimum film thickness is therefore reduced to satisfy the
force balance, increasing both the hydrodynamic peak and asperity contact
pressure.
When honing spacing (S) is evaluated, the results show that although
its influence is less than honing depth, a reduction in honing spacing -
i.e. more dense honing grooves, gives a slight reduction in FMEP. The
explanation for this is that the more frequent, but shallow, honing grooves
drag more lubricant into the contact and aid in building a hydrodynamic
film.
In Fig. 5.21 the total wear rate for the ring-pack is presented. As was the
case with honing angle, the wear rate follows a similar trend to FMEP, with
the leakage phenomenon that has just been introduced reducing film thick-
ness, which in turn increases the degree of asperity contact and therefore
wear.
88 CHAPTER 5. OPTIMIZATION OF TEXTURE

Wear rate (nm/h) 7.5

2.5

0
300
3
250
2
200
1
1500.5
Honing Spacing (m) Honing Depth (m)

Figure 5.21: Map of wear for full ring-pack with varying honing spacing and
depth.

Finally, Fig. 5.22 shows the oil consumption risk with a map of the mean
lubricant height left behind on the cylinder liner surface (above the plateau
surface - the same definition as introduced earlier in this section), with
varying S and D. The reader should note that the plot has been rotated
through 180 compared to Fig. 5.18 and Fig. 5.21 to aid in viewing. The
map shows that a reduction in the depth of honing grooves increases oil
consumption risk. This is logical, as more oil will be left behind on the liner
surface, above the plateau, with shallower honing grooves and the improved
lubrication that they provide.

5.3 Discussion
In the preceding sections an attempt has been made to show how three
honing parameters - depth (D), spacing (S) and angle (), affect the lubri-
cation of the PRCL contact and their influence on film thickness, friction,
wear and oil consumption risk. The simulations showed that a reduction
in honing angle and depth, and an increase in density, could increase mini-
mum oil film thickness and reduce frictional power losses and wear. Taking
todays surface ( = 50 , S = 200 m, D = 1 m), the simulations suggest
that an optimized surface of ( = 30 , S = 150 m, D = 0.5 m) could
reduce ring-pack FMEP from 12.49 kPa to 10.21 kPa and reduce specific
fuel consumption by 0.16% at 1200 rpm and 50% load, based on estimates
5.3. DISCUSSION 89
Average lubricant height left behind on liner surface (m)

0.2

0.15

0.1

0.05

0
150
0.5
200 1
250 2
300 3
Honing Spacing (m) Honing Depth (m)

Figure 5.22: Map of oil consumption risk with varying honing spacing and
depth.

of power loss breakdown provided by Scania.


However, the values presented here should not be read as absolute val-
ues of optimized honing texture. The representation of texture is rather
idealised and does not necessarily exactly represent the real surface topog-
raphy. Also, only one speed and load condition has been investigated, 1200
rpm and 50% load. Rather, this work aims to investigate the physical ef-
fects that changing honing parameters has on PRCL lubrication. Further, a
simulation tool has been demonstrated that allows for further investigation,
optimisation and understanding of this contact.
There are several physical phenomenon not included in this model that
must be considered when optimizing texture. One of the perceived functions
of honing texture is the entrapment and removal of wear particles from the
PRCL contact. Although a shallower honing angle and more transverse
texture has shown to be optimal for PRCL lubrication in this model, it is
envisaged that more longitudinal texture would be more effective at carrying
wear particles away from the contact. Longitudinal texture might be more
beneficial regarding wear than this model predicts, however this is only a
90 CHAPTER 5. OPTIMIZATION OF TEXTURE

hypothesis by the author.


The model also indicates that shallower honing grooves provide enhanced
lubrication of the TCR around TDC. However, there is most likely a limit
whereby if the honing grooves become too shallow they are no longer deep
enough to keep wear particles away from the conjunction. To answer this
question, and the previous one, it would be beneficial to undertake an anal-
ysis of wear particle size in the future.
Another point to consider is that when manufacturing cylinder liners
changing honing angle may have other unintended effects on cylinder liner
shape. A shallower honing angle, where the honing tool rotates faster but
axially reciprocates slower, will in all likelyhood create a rounder but less
axially straight cylinder liner. Conversely, a cylinder liner with a greater,
or steeper, honing angle is likely to be more axially straight but less round.
Therefore, to fully capture all of the effects of a change in honing angle a
global, non-axisymmetric, model of the PRCL may be required.
One other effect not captured in this model is the evaporation of lu-
bricant from the cylinder liner surface. If a large quantity of lubricant is
deposited near the top of the cylinder liner, where temperatures are highest,
then it may not stay there and be available to lubricate the PRCL contact
when the piston ring returns on the following up-stroke. Deeper honing
grooves may aid in the retention of lubricant on the cylinder liner surface
in this situation.
Chapter 6

Conclusions

What are the findings of this research?

This research started with the development of an all-regime simulation


tool of the PRCL contact. The tool was validated with both bench tests and
full single cylinder engine tests. A method was developed to analyse different
surface textures in a fast and efficient manner with the implementation of
flow factors. The following conclusions can be drawn on optimum honing
texture:

Honing texture that is more transverse (a smaller honing angle) leads


to improved hydrodynamics in the PRCL contact due to better en-
trainment of lubricant into the contact.

Reduced honing depth provides a friction reduction by reducing the


leakage of lubricant from the contact around the reversal points. How-
ever, it must be remembered that the texture is also responsible for
the entrapment of wear particles - a phenomenon not simulated in
this work - and reducing the honing depth too much may impact this
functionality.

An increase in the density of honing grooves, improving the entrain-


ment of lubricant into the contact, slightly increases film thickness
and reduces friction, within the bounds investigated in this work.

An initial estimate for the potential reduction in frictional losses,


within the bounds and with the limitations specified here, is a reduc-
tion in FMEP by 2.28 kPa and a corresponding reduction in specific
fuel consumption of 0.16%.

Furthermore, this research has shown that tribological modelling of the

91
92 CHAPTER 6. CONCLUSIONS

PRCL contact can be a useful aid for the understanding of the effect of
varying texture on PRCL lubrication.
Chapter 7

Future work

What is still to do?

This research has led to the development of a simulation tool that will
be used within Scania for future engine development. It is intended that
simulation development and the understanding that it brings will continue
further. Today the following areas are seen as needing further work.
There should be a study on practical techniques for the implementation
on real, measured, surfaces of the mixed lubrication model introduced in
Chapter 2. Although the theory and numerics for the mixed lubrication
model are well developed, there are many questions about the implementa-
tion of the model still outstanding. Questions that need answering include
how many surface measurements should be analysed to be statistically repre-
sentative and how large and at what resolution should surface measurements
be taken.
The development of a global model of piston ring conformability would
aid further understanding of PRCL lubrication. This would allow for non-
axisymmetric effects to be investigated such as cylinder liner out-of-roundness,
which may vary with honing angle as previously discussed.
Wear particle analysis should be conducted. As discussed in Chapter 5
it would be of interest to evaluate the size of the wear particles that occur
in the PRCL environment. This would be a start to understanding how
honing texture can remove wear particles from the PRCL contact.
Finally, the model should be used to conduct a more detailed investiga-
tion of all engine operating conditions and permutations of honing texture.
Furthermore, comparisons should be made between flow factors calculated
on the artificial surfaces presented in Chapter 5 and real cylinder liner sur-
faces.

93
94 CHAPTER 7. FUTURE WORK
Part II

Appended Papers

95
Paper A

A numerical model to
investigate the effect of
honing angle on the
hydrodynamic lubrication
between a combustion engine
piston ring and cylinder liner

97
99

Proceedings of the Institution of Mechanical Engineers, Part J:


Journal of Engineering Tribology, vol. 225 no. 7, pp. 683-689

A numerical model to investigate the effect of honing


angle on the hydrodynamic lubrication between a
combustion engine piston ring and cylinder liner
A. Spencer1 , A. Almqvist, R. Larsson
Division of Machine Elements, Lule
a University of Technology, SE-971 87, Lule
a
Sweden

Abstract
A numerical model has been developed to investigate the effect of cylin-
der liner honing angle on hydrodynamic lubrication between piston ring
and cylinder liner. The Reynolds equation was solved in 2D with periodic
boundary conditions. An artificial surface texture was generated, based on
a real surface measured with white light interferometry. Cavitation was
modelled with the Vijayaraghavan and Keith algorithm. Honing angles be-
tween 25o and 75o were investigated to find the effect of honing angle on
film thickness.

Keywords: cylinder liner, honing, texture

1
Corresponding author.
E-mail address: andrew.spencer@ltu.se
100 PAPER A.

A.1 Introduction
Typically, for a medium size car over an urban cycle, only 12 per cent of the
total power from the fuel is converted to useful power at the wheels [4]. The
rest of the energy from combustion of the fuel goes into cooling, exhaust,
pumping, and mechanical losses. The mechanical contribution amounts to
15 per cent [4] of the total losses and the friction between the piston assem-
bly and cylinder liner is the single largest contributor; amounting to 2040
per cent of the total mechanical losses [7]. Furthermore, the compression
rings typically contribute 45 per cent of all mechanical losses in a multi-
cylinder engine [11]. If all the mechanical losses can be reduced by 10 per
cent, then fuel efficiency could be increased by a value between 1.5 and
2.5 per cent [4, 11]. Therefore, the study of the piston ringcylinder liner
conjunction (PRCL) is highly important in reducing friction and improving
fuel economy, one of the main drivers in engine design today.
Usually, a cylinder liner is honed to apply the desired surface finish.
With modern plateau honing [58], the finished surface consists of relatively
smooth plateaux separated by grooves that lie at two opposing angles to
form a cross-hatch pattern. The flat plateaux provide a smooth surface
to allow for hydrodynamic film build-up between the piston rings and the
cylinder liner surface. The grooves serve two purposes; primarily, they act
as an oil reservoir. The only available oil for lubrication is supplied to the
contact by the movement of the piston rings on the upstroke; this oil is
retained in the grooves to lubricate the contact on the piston downstroke.
It is believed that a secondary reason for the grooves is to transport wear
debris away from the contact. Small particles should be swept downwards
in the grooves towards the crankcase and hopefully cause only minimal
scratching to the smooth plateau surfaces.
Very few numerical studies have investigated the effect of honing param-
eters, and in particular honing angle, on lubrication performance. There-
fore, this may be an area that has great potential for optimization, and
changing a parameter such as honing angle should not add any significant
costs to cylinder liner manufacture. Also, many existing PRCL models do
not consider non-Gaussian roughness patterns which are typical of a honed
surface.
A similar study, investigating the effect of honing angle on lubrication
performance, was undertaken by Michail and Barber [27]. They too im-
plemented an analytical emulation of a honed surface and used flow fac-
tors calculated with the Patir and Cheng [23] Average Flow model in their
solution of the Reynolds equation. However, their surface representation
consisted of a combination of cosine waves to form a surface with the peaks
of the waves removed to resemble the plateaux. It is suggested that the
A.2. THEORY 101

method developed in this article to resemble the honing pattern is a more


accurate representation of a real honed surface and in particular the sides
of the grooves.

A.2 Theory
A.2.1 Surface texture
This study aims to investigate the effect of different honing angles on PRCL
lubrication. Real surface topography was only available for a single exist-
ing honing angle and so to make the study of different angles possible, an
artificial surface texture is generated and used. Another advantage of using
an artificial surface is that it can be made periodic, a requirement for the
boundary conditions implemented. The surface of a real cylinder liner was
measured with a Wyko NT 1100 white light interferometer and this was
used as a basis for the artificially generated surfaces. An investigation of
the surface showed that although the diamond-like pattern on the surface
was not uniform, the diamonds had a typical, or mean, area of 0.048 mm2 ,
as highlighted in Fig. A.1. In this study, to keep textures with different

Figure A.1: White light interferometry image of cylinder liner surface

honing angles comparable, the area of each diamond will be kept constant
at this value. To generate a single groove, Eq. (A.1) is used (taken from
102 PAPER A.

Table A.1: Solution Parameters

Parameter Value
Plateau depth (m) 2
Percentage plateau (%) 70
Area of one diamond texture (mm2 ) 0.048
Compressibility Factor, (Pa) 6.9x107
Dynamic Viscosity, (Pas) 0.04
Inlet Pressure, PIN (kPa) 437.5
Outlet Pressure, POU T (kPa) 357.0
Simulation Time (s) 1.5x104
No. of time steps 50
Ring width (mm) 4
Ring effective width (mm) 2.7
Piston Speed (m/s) 5
Tangential Ring Force (N) 10
Bore Diameter (m) 0.1

reference [46]);
2
hT (x, y) = 10(x+ky) cos[2(x + ky)]. (A.1)

where x is the position in the axial direction, y is the position in the ra-
dial direction, k is the honing angle parameter, and is the honing width
parameter. Two of these variables should be explained. The width of the
honing groove is set with , which must be iterated for. The iteration pro-
cedure works as follows: an initial value of is guessed and the total area
of plateau on the generated surface is calculated as any nodes in the top 0.1
per cent as sorted by height. The percentage of the total generated surface
that is considered plateau is then calculated and this is compared with the
desired percentage Table A.1. The value of is then increased if the per-
centage plateau is too high or decreased if it is too low. This procedure is
repeated until the percentage of plateau on the generated surface is within
0.1 per cent of the desired value. The k parameter is dependent on the
desired angle of the honing groove, as described by Eq. (A.2):

1
k= (A.2)
tan
A single-diagonal groove is created using Eq. (A.1), as illustrated in Fig. A.2.
This single groove is mirrored in both the x and y planes to give one
artificial diamond. This single diamond is tiled to create a complete surface
A.2. THEORY 103

(a)

(b)

Figure A.2: (a) Artificial honing groove (honing angle = 50o ) and (b)
cross-section through artificial honing groove along dotted line illustrated
in (a)

for the ring to slide over (Fig. A.3). The surface is long enough for the ring
to be able to slide far enough, through enough time steps, for a steady-state
periodic solution to be reached. The artificial surface only needs to be one
diamond wide in the circumferential direction; periodic boundary conditions
are implemented in the solution; so, this is all that is required. The depth
104 PAPER A.

Figure A.3: Generated surface texture

of the artificial honing scratches is also based on the real measured surface.
Fig. A.4 shows a typical cross-section. It can be observed that the honing
depth has some variation. In this simulation, a value of 2 mm has been used
as a representative depth.

Figure A.4: Cross-section of measured liner surface

A.2.2 Cavitation algorithm


The Vijayaraghavan and Keith [33] cavitation algorithm was used to solve
the Reynolds equation, implementing cavitation boundary conditions at film
rupture and reformation. The approach is similar to the Elrod algorithm
[19] but with two advantages; compressibility is considered in the full film
zone and the discretization is rigorously derived, rather than being the result
of considerable experimentation. The Reynolds equation is written as:
c h3 g d c h3 g d
   
dc h d c hU d
+ + = 0. (A.3)
dt dx 2 12 dx dy 12 dy
A.2. THEORY 105

where is the bulk modulus (given in Table A.1) and g a switch function
becoming 1 in the full-film zone and 0 in the cavitated zone. The type of
zone is calculated from the value of ; above 1, the region is full-film and
below 1, it is cavitated. The film pressure is found from
P = Pc + gln (A.4)
where Pc is the lubricant cavitation pressure and P the lubricant film pres-
sure. Once the discretization has been applied, Eq. (A.3) can be written in
the form
aw i1,j + ae i+1,j + an i,j1 + as i,j+1 + ap i, j = rhs, (A.5)
where
(chi1,j )
aw =U C1 (hw gw )
4x
(ahi+1,j )
ae =U C1 (he ge )
4x

an = C2 (hn gn )

as = C2 (hs gs ) (A.6)

(bhi,j ) hp
ap =U + C1 [(he + hw ) gp ] + C2 [(hn + hs ) gp ]
4x t

rhs = C1 [he ge (he + hw ) gp + hw gw ] C12 [hs gs

hp i,j

(hn + hs ) gp + hn gn ]
t
with the Couette flow coefficients;
gi+1,j + gi+1,j
a =
2
gi+1,j + gi+1,j
b =2 (A.7)
2
gi+1,j + gi+1,j
c =
2
Poiseuille flow coefficients;

C1 =
12x2
(A.8)

C2 =
12y 2
106 PAPER A.

Table A.2: Subscript nomenclature

Subscript Description
e one node downstream
n one node to the left
prev previous iteration
s one node to the right
w one node upstream
* previous time step

The subscript nomenclature is given in Table A.2.

A.2.3 Boundary conditions


A pressure is specified at the inlet and outlet points based on typical gas
pressures encountered (Table A.1). Periodic boundary conditions are im-
plemented in the circumferential (perpendicular to entraining motion) di-
rection. The inlet is assumed to be always fully flooded.

A.2.4 Film thickness


The ring is assumed to have a parabolic shape. The total film thickness is
a combination of this shape, the contribution from the surface texture, and
a minimum film thickness, given by Eq. (A.9).

h = hmin + ax2 + hT . (A.9)

A linear system of equations is formed and is solved for directly. Once


has been found, the switch function, g, is updated at every node based on
the calculated value of . If is greater than 1, then g is set to 1; if is
less than 1 g is set to zero. This process is repeated, is solved for, and the
value of g is updated based on the new result, until convergence is reached.
Convergence is assumed to have been reached when the following becomes
true RR RR
P dxdy PP REV dxdy
RR < 105 . (A.10)
PP REV dxdy
The problem was solved over a grid of 1000 x 50 nodes to produce a stable
solution. If too coarse a grid was chosen, then anomalous pressure spikes
would occur in the solution, as illustrated in reference [36]. Refining the
mesh further created no noticeable change in the solution in regard to the
total load supported by the contact with a given separation.
A.3. RESULTS 107

A.2.5 Force balance and time dependence


To balance the forces (or pressures) that the piston ring is subjected to, an
iterative procedure is used. The procedure finds the value of hmin required
for Eq. (A.11) to be true. Everything on the right-hand side is known and
the mean hydrodynamic pressure on the left hand side is a function of hmin
in Eq. (A.9).
Ft
PHY D = max (Pin , Pout ) + d  (A.11)
2 l

where PHY D is the mean hydrodynamic pressure. The f zero algorithm in


Matlab is used to find the required hmin to balance the forces. This is a
combination of the bisection, secant, and inverse quadratic interpolation
methods. Once the force balance has been satisfied, the solution increments
one time step. A new surface texture is calculated as the ring has slid over
a proportion of the texture pattern and the force balance is then solved
again. The time is incremented until the ring has slid over five full tex-
ture diamonds, enough for a periodic value of hmin to become apparent.
Table A.1 lists all the parameters used in the solution. The simulation is
run with parameters that represent typical conditions at the mid-stroke of
the engine cycle. The speed is kept constant and thus the result is at a
stationary condition. The repeating pattern is, however, necessary so that
a time-dependent solution can be reached. The conditions used were chosen
as the model presented is currently only capable of simulating hydrodynamic
lubrication; the mixed and boundary lubrication regimes at Top Dead Cen-
tre (TDC) and Bottom Dead Centre (BDC) were intentionally avoided.

A.3 Results
The honing angle of the cylinder liner measured with white light interfer-
ometry (Fig. A.1) was 50. Honing angles of 25 from this existing case
have been investigated, in increments of 5. A typical pressure distribution
is illustrated in Fig. A.5. The simulation shows that a steady-state solution
is obtained after the ring has transversed approximately two complete tex-
ture patterns, or 20 time steps. Fig. A.6 shows the minimum film thickness,
hmin, as a function of time as the ring slides over the diamond texture.
The smooth case is also plotted, which shows an unchanging film thickness
after the first time step, as would be expected. From Fig. A.6, it can be
seen that a fairly steady state of film thickness occurs after 20 time steps,
with oscillations occurring periodically as the ring passes over each honing
diamond. This pattern was observed for all the honing angles investigated.
Therefore, in Fig. A.7, the average minimum film thickness is calculated
108 PAPER A.

Figure A.5: Calculated lubricant pressure across ring width.

for each honing angle from the 20th time step onwards, so that only the
steady-state points are considered.

A.4 Discussion
The first issue to highlight is that the variation in film thickness with honing
angle is minimal. A 20 nm variation was observed across the range of honing
angles 0.4 per cent of the total minimum film thickness of 5 mm. A curve
has been fitted to the calculated average results; however, the values for
45 and 50 deviate significantly from this. However, the results generally
show that the film thickness is greater with a smaller honing angle. This

Figure A.6: Minimum film thickness as a function of time steps.


A.5. CONCLUSION 109

Figure A.7: Average minimum film thickness versus honing angle.

is supported by the conclusions in reference [59] which also show that a


smaller angle leads to a greater oil film thickness. The simulations presented
here assume that the inlet to the contact is always fully flooded. This is
unlikely to be the case, particularly during the mid-stroke as simulated
here, when large film thicknesses of 5 mm are predicted. In reality, there
may not be enough lubricant present on the liner to form films with this
much thickness. Periodic boundary conditions have been implemented in
the radial direction. This means that any lubricant escaping from either
side of the contact, which may be significant in the honing grooves, flows
into the contact from the opposite side. Hence, all the lubricant entering
the contact will leave from the trailing edge. This assumption was made
because the contact is many times wider than it is long. In the real contact,
assuming axi-symmetric conditions, the only leakage that can occur from
the side of the contact is at the ring gap.

A.5 Conclusion
It has been shown that the effect of the honing angle in the middle of the
piston stroke is negligible compared to the large hydrodynamic film already
developed. A simulation run closer to TDC or BDC, thereby running at
conditions closer to mixed lubrication, might yield a greater effect from the
honing angle parameter as the surface grooves become a bigger percentage of
110 PAPER A.

the total film thickness. It would be also be of value to investigate angles far
away from 50, such as 140, which can be machined as a product of helical
slide honing. Such an angle has been shown to reduce oil consumption
during bench tests [58].

A.6 Acknowledgements
The authors thank Stiftelsen for Strategisk Forskning (SSF) and ProViking
for funding this study and Scania AB for facilitating discussions and pro-
viding technical data.
Paper B

A semi-deterministic
texture-roughness model of
the piston ring - cylinder
liner contact

111
113

Proceedings of the Institution of Mechanical Engineers, Part J: Journal


of Engineering Tribology, June 2011, vol. 225 no. 6, pp. 325-333

A semi-deterministic texture-roughness model of the


piston ring - cylinder liner contact
A. Spencer1 , A. Almqvist, R. Larsson
Division of Machine Elements, Lule
a University of Technology, SE-971 87, Lule
a
Sweden

Abstract
Many simulations already exist to model the piston ring-cylinder liner con-
tact; however, very few models have been used to investigate the optimum
surface texture. An axi-symmetric, time dependent two-dimensional semi-
deterministic texture-roughness model of the piston ring to cylinder liner
contact with periodic boundary conditions and mass preserving global cavi-
tation has been developed. The cylinder liner texture, generated by honing,
was considered deterministically on the global scale, after an investigation
comparing deterministic and homogenized solutions.
The surface texture of a real cylinder liner was measured with white
light interferometry and an algorithm developed to generate an artificial
periodic texture representative of the real surface. The effect of cylinder
liner plateau roughness has been incorporated on the local scale by homog-
enization of the Reynolds equation and calculation of flow factors from real
surface topography. Using the homogenization technique to incorporate the
effect of surface roughness leads to a more efficient solution than mesh re-
finement of the deterministic problem as the roughness does not need to
be resolved on the global solution domain, allowing for significantly less de-
grees of freedom in the global problem. The lubricant boundary pressures
have been calculated using results from a numerical ring-pack model and
the lubricant viscosity has been adjusted based on the cylinder liner wall
temperature.
It was found from the result of a comparison between deterministic and
homogenized solutions that surface texture should be modelled on the global
and not on the averaged roughness scale as is the case with many previous
investigations.

1
Corresponding author.
E-mail address: andrew.spencer@ltu.se
114 PAPER B.

Keywords: honing, cylinder liner, piston ring, texture, homogenization


B.1. INTRODUCTION 115

B.1 Introduction
Power losses from an internal combustion engine can account for a large
percentage of the total power out- put of the engine, varying from 10 per cent
at full load to 100 per cent at idle [5]. Frictional losses account for a large
percentage of these and the piston ringcylinder liner (PRCL) conjunction is
the single largest contributor to frictional losses in an IC engine, accounting
for 20-40 per cent of the total frictional losses [7]. The compression rings
are responsible for 4-5 percent of all losses in a multi-cylinder engine [11]. If
parasitic losses can be reduced by 4 per cent, then a fuel efficiency increase of
1 per cent can be realized [11]. Therefore, studying this tribological contact
has great potential for reducing friction and improving fuel economy in an
engine, something that is at the forefront of engine design today.
Typically, a machining process known as honing is used to apply the
desired finish to the cylinder liner surface. The grooves that the honing pro-
cess leaves behind are believed to be important in controlling the amount
of oil available in the contact, by both retaining oil on the liner surface and
improving the distribution of oil. Another perceived function of the hon-
ing texture is to allow wear debris, generated during boundary lubrication
around top dead centre (TDC), to be channelled away from the conjunction
so as to cause only minimal damage and scratching to the smooth plateaux
which are said to be important for fluid film generation. In the case in-
vestigated here, a top compression ring, the contact width is 4 mm and
the surface plateaux have an approximate average width of 0.2 mm. In
this study the effect of applying the honing to hydrodynamic lubrication is
investigated.
Most existing PRCL models do not consider the effect on lubrication
of non-Gaussian roughness patterns typical of a honed surface. Several
existing simulations [12, 13], implement the Patir and Cheng average flow
model [23] to simulate the effect of the surface finish on hydrodynamic
lubrication. These simulations lump together both surface texture generated
by honing and plateau roughness in the same set of flow factors. This article
investigates whether this is a valid approach to take or whether these scales
should be separated.
Very few numerical studies have investigated the effect of honing texture
on lubrication performance in isolation. One which has been undertaken by
Michail and Barber [59] implemented an analytical emulation of a honed
surface. Their study used Patir and Cheng flow factors and took the honed
surface to be a combination of cosine waves with the peaks removed to
simulate the plateaux. It is suggested that the artificial surface texture
presented in this article could be a more realistic representation of a honed
surface than their approach.
116 PAPER B.

The limited work undertaken in modelling the geometry of the surface


texture means that this is an area that has great potential for optimization,
and changing the details of the honing texture should not add any significant
costs to cylinder liner manufacture.

B.2 Model Development


In this section the development of the numerical model is described.

B.2.1 Geometry and global problem


A small portion of the ring is modelled in two dimensions with periodic
boundary conditions applied in the circumferential direction. This means
that the pressure and film thickness on the left hand side of Fig. B.1 is the
same as that on the right-hand side. It is assumed that the ring is symmetric

Figure B.1: Global Geometry


B.2. MODEL DEVELOPMENT 117

about its axis to make this assumption valid. Clearly, this is not the case at
the ring gap, but as the ring has a much greater circumference than width
it should be valid for the majority of the contact. The ring geometry giving
a lubricant film thickness, h, is mode led with the equation:

x2
h = hmin + . (B.1)
2R
where hmin is the minimum film thickness, x is the position in the axial
direction, and R is the ring radius. Eq. (B.1) approximates a radius for
small values of x. In this case, the contact width for the top compression
ring is 4 mm, and so x takes a value of 2 mm. R is taken to be 0.4 m.

B.2.2 Global surface texture


An artificial cylinder liner surface texture was generated, based on real
cylinder liner surface topography. The surface texture of a cylinder liner
was sampled using white light interferometry with a Wyko NT1100 to in-
vestigate the typical honing pattern, illustrated in Fig. B.2. The direction

Figure B.2: Measured Cylinder Liner Surface with WLI

of entraining motion in the figure is leftright. The honing angle is 50 .


Although the diamond patterns generated on the surface by the honing
operation are not uniform, a typical diamond, as highlighted in the image,
118 PAPER B.

has an area of 0.048 mm2 . An algorithm was generated to produce an


artificial surface texture with variable honing angle, groove depth, and per-
centage of plateau while keeping the diamond area constant at 0.048 mm2 .
The area was held constant so that different artificial textures are compara-
ble and realistic. To generate a texture groove in the surface, the following
equation was used to describe the height of the honing texture, hT .
2
hT (x, y) = 10w(x+ky) cos[2(x + ky)]. (B.2)

Two variables require some explanation. The coefficient is used to adjust


the width of the honing groove. This is done iteratively to achieve the
desired percentage of plateau. The parameter k is a function of the angle
of the honing groove and is derived as:
1
k= (B.3)
tan
where is the honing angle. Eq. (B.2) is used to create a diagonal groove,
illustrated in Fig. B.3. This equation was chosen, as opposed to the cosine

Figure B.3: Artificial Honing Groove

assumption used by Michail and Barber [59], because it is considered to be


a closer approximation to the real surface topography. Furthermore, the
plateaux area and honing groove depth can be adjusted without influenc-
ing the overall shape of the grooves considerably. With the cosine wave
assumption, removing an increasing amount of the cosine peaks to increase
B.2. MODEL DEVELOPMENT 119

the plateaux area changes the shape and gradient of the sides of the honing
grooves. To create the diamond shape as seen on the real surface this single
groove is mirrored in the xdirection and then in the ydirection. This
single diamond is tiled to create a complete surface that has the same width
as the piston ring (Fig. B.4). The surface is only one diamond wide in the

Figure B.4: Artificial Surface

circumferential direction. Periodic boundary conditions are implemented in


the solution as illustrated in Fig. B.1; in effect, the strip in Fig. B.4 repeats
infinitely in the circumferential direction.
Information regarding the depth of the honing scratches was also ob-
tained from the Wyko measurements. It was observed that the honing depth
has some variation but an average of 2 m was found which was therefore
chosen as a representative depth. The honing angle of the cylinder liner
measured was 50 which was also chosen for use in these simulations.

B.2.3 Reynolds equation

The homogenized, iso-viscous, time-dependent Reynolds equation is solved


to model the lubrication in the PRCL conjunction:

dh
(A0 (x, y) p0 (x, y)) = B0 (x, y) + , (B.4)
dt
120 PAPER B.

where p0 is the homogenized film pressure, h is the average film thickness at


a given separation, = 6U , = 12, and A0 and B0 are the flow factors:
   
a11 a12 b12
A0 = and B0 = (B.5)
a21 a22 b22

These are described in detail by Almqvist et al. [46]. Experimental vali-


dation of using the homogenization technique to incorporate the effects of
surface roughness is not presented here but has been previously dealt with
by Sahlin et al. [30]. The factors a12 and a21 are the cross-flow terms,
describing the magnitude of the pressure-induced flow in the direction per-
pendicular to the entraining motion. Due to the symmetry in the problem
(seen in Fig. B.1 and Fig. B.4), it should be possible to neglect these terms.
The same is true for b22 , describing the shear-induced flow normal to the
piston motion; this term can also be neglected. This assumption is validated
in section B.2.4 when the flow factors are calculated and the cross-flow terms
are found to be many orders of magnitude smaller than the other terms.
To model cavitation in the contact, the Ausas et al. algorithm is used
[60]. The Vijayaraghavan and Keith [33] and Elrod [19] algorithms were
also tested but were found to be unstable around the honing grooves in the
cavitated regions. No such difficulties were encountered with the Ausas et
al. technique. Eq. (B.4) can be written as:

d h
(A0 p0 ) = (B0 (x) ) + (B.6)
dt
where:
p 0 = 1 (Full Film region)
(B.7)
p = 0 < 1 (Cavitated region)
Here, is the film fraction. The problem is non-dimensionalized and dis-
cretized as is described by Ausas et al. [34]. The solution procedure, a
GaussSeidel relaxation scheme, is the same as is described in reference [34]
and so will not be repeated here.
The inlet and outlet pressure are set to the inlet and outlet gas pressures,
respectively; these are the pressure in the combustion chamber and the inter-
ring gas pressure. The contact is assumed to be fully flooded at all times as
an oil availability model has not been used in this study.

B.2.4 Flow factors


Two sets of flow factors have been calculated. One set for a single artificial
diamond of the artificial surface texture, as illustrated in section B.2.2, and
another for a real measured area of plateau roughness as measured using a
B.2. MODEL DEVELOPMENT 121

Figure B.5: Plateau surface roughness

Wyko NT1100, as illustrated in Fig. B.5. The flow factors A0 for the surface
shown in Fig. B.5 are illustrated in Fig. B.6. The value of the flow factor (y-
axis) is plotted against the surface separation divided by the separation at
which mechanical contact occurs; that is, a value of 1 on the x-axis is where
mechanical contact occurs. It is observed that the cross-flow terms, a12 and
a21 , are miniscule in comparison with the a11 and a22 terms especially in
the full film region and so the assumption in section B.2.3 is valid.
The flow factors are calculated for 55 different separations, from contact
to a mean separation of 8 m and then for any given separation the flow
factor is interpolated for using a cubic spline. The calculation process and
contact mechanics model used is well described in references [30] and [29]
and is not repeated here.

B.2.5 Model input parameters


The engine simulated in this article is a heavy-duty truck engine. Two im-
portant input conditions for the model are provided from a separate simula-
tion developed with the commercial AVL Piston and Rings software; these
are the cylinder liner temperature and combustion chamber and inter-ring
gas pressures. The cylinder liner temperature is used to calculate the vis-
122 PAPER B.

1.6

1.4

1.2
Flow Factors

0.8
a11
0.6 a12
a21
0.4 a22
0.2

0
0 0.5 1 1.5 2 2.5 3 3.5 4
/h1

Figure B.6: Flow factors for plateau surface roughness

cosity of the oil on the liner. It is assumed that the temperature of the
lubricant in the contact is the same as the liner temperature at that point
and the Reynolds temperature-viscosity relationship is used:

= 0 e(T T0 ) . (B.8)

where is the lubricant dynamic viscosity, 0 is the lubricant dynamic vis-


cosity at the reference temperature, is the Reynolds temperature - viscos-
ity coefficient, T is the lubricant temperature (taken to be the temperature
of the cylinder liner), and T0 is the lubricant reference temperature, 40 C
in this case. The gas pressures are required for two purposes. First, the gas
pressure acts on the back of the ring, producing a force radially onwards
which must be supported by the lubricant film or asperity contact. It is
assumed that at any point in the cycle the highest of either the combustion
or inter - ring pressure acts on the back of the ring. This assumption is
made as, neglecting inertial forces; the ring will be forced against either the
top or the bottom edge of its groove by the highest pressure - sealing it and
preventing the lower pressure from acting upon it. The other requirement
for the gas pressures are the Reynolds equation boundary conditions. It is
assumed that the inlet and outlet lubricant pressures are the combustion
chamber and inter-ring gas pressure. Fig. B.7 shows the gas pressures for
three different engine running speeds, all at full load. The piston speed, and
B.2. MODEL DEVELOPMENT 123

250
1200RPM Combustion Chamber
1200RPM Inter-ring
200 1600RPM Combustion Chamber
1600RPM Inter-ring
Pressure (bar)

1900RPM Combustion Chamber


150
1900RPM Inter-ring

100

50

0
0 90 180 270 360 450 540 630 720

Crank Angle ( )

Figure B.7: Gas pressure vs crank angle

hence the entraining motion of the contact is modelled with Eq. (B.9):

LN sin cos
U= 1 +  (B.9)

60 1/2
2a 2

L sin2

where L is the engine stroke, N is the engine speed in r/min, is the crank
angle with 0 being taken as TDC, and a is the conrod length.

B.2.6 Force Balance and time dependence


An iterative procedure is used to calculate the value of hmin required for the
lubricant film pressure to support the opposing load, a combination of the
gas pressure acting on the back of the ring and ring tension. This balance
is defined as:
2Ft w
Z Z
pdxdy = max (pin , pout ) wl + (B.10)
d

where PT OP is the combustion chamber gas pressure, PBOT T OM is the inter-


ring gas pressure, w is the periodic contact width, l is the ring width, Ft
is the ring tension, and d is the bore diameter. The f zero algorithm in
Matlab [61] is used to find the required hmin to balance these forces. This is
a combination of the bisection, secant, and inverse quadratic interpolation
124 PAPER B.

methods. Once the force balance has been satisfied, the solution increases
one time step, equating approximately 0.5 crank angle depending on engine
speed. A new gas pressure, piston speed, and temperature are found and
the force balance is solved again. As the Reynolds equation implemented
is time dependent more than one complete engine cycle must be simulated
for the entire solution to converge. The discretized time-dependent term
needs a solution from the previous time step. For the first crank angle
solved for this information is not available. An initial film thickness must
be guessed; in the simulations presented here, a guess of 1 m is made at
360 crank angle. After one complete engine cycle, the film thickness and
change in film thickness with time (gradient) are compared with that found
on the previous cycle and the simulation is stopped when these are within
a specified tolerance. Typically this requires 1.1 engine cycles.

B.2.7 Friction
Friction is calculated for the entire engine cycle. The friction is composed
of two components, the viscous full film friction and the boundary contact
friction. The boundary friction is calculated by:
Z
Fbound = Pcp dA (B.11)

In this simulation the coefficient of friction between the piston ring and
cylinder liner, , is taken as 0.1. A is the area of contact between the ring
and liner and PCP is the mean contact pressure, found from the contact
mechanics model. The hydrodynamic friction is calculated with additional
flow factors c11 , d11 , and d12 using [29].
    
1 h p0 p0
Z
F0 = U + 6c11 + d11 + d12 dx (B.12)
h 2 dx dy

B.2.8 Grid size and convergence


The effect of number of nodes on the solution was investigated. With surface
texture described on the global scale (Fig. B.4) the number of nodes must
be big enough to describe it in sufficient detail. A large number of nodes was
first chosen (1200 x 60) and then was reduced until the results produced were
no longer considered acceptable. It was found that 242 x 12 nodes produced
a supported load within 0.6 per cent and a maximum pressure within 0.1 per
cent of the 1200 x 60 solution. This was considered an acceptable error, and
the solution time with the smaller number of nodes was almost two orders
of magnitude smaller. As has been mentioned previously, 55 flow factors
were calcu- lated for a range of separations between 0.05 and 8 m. These
B.3. TEXTURE INVESTIGATION 125

limits were chosen so that all points throughout the engine cycle would fall
within this range (minimum film thickness varies between 0.05 and 5.2 m).
The only points not falling within this range, due to the curvature of the
ring, are the outer edges during the mid-stroke when large film thickness
occurs. However, above 5 m the flow factors calculated change very little
as the homogenized surface features have little effect. Twenty-five of the
flow factors were calculated with separations where contact occurs and 30
at separations where no contact occurs. These numbers were decided upon
through previous experience but were more than sufficient to define the
shape of the flow factor plots.

B.3 Texture Investigation


A model has been described in section B.2 that has the capability to model
surface texture, the honing grooves, both deterministically and with flow
factors. Surface roughness can also be modelled with flow factors, however
not deterministically as the number of nodes to resolve surface roughness
across the contact would become so large that the solution time would
become too great.
Existing simulations often treat surface roughness and texture on the
same scale with Patir and Cheng flow factors. However, the conjunction
includes many scales, the biggest of which (the texture) is reasonably close
in magnitude to the contact width. In the specific contact investigated there
are approximately 20 repetitions of the texture pattern across the contact
width. If the wavelength of the surface texture is not of a separable-enough
scale from the contact width then an unacceptable error will result from
including it within homogenized flow factors. There is no definitive rule as
to what constitutes separable scales. Almqvist and Dasht [39] compared
deterministic solutions with varying roughness wavelengths with the ho-
mogenized solution. They found that with a sinusoidal roughness profile
the error in supported load with a given separation fell from 7 per cent with
a roughness wavelength of 1/24 of the contact width to 0.5 per cent with
a wavelength of 1/28 of the contact width. Another factor to consider is
that any cavitation due to the texture pattern will not be considered if the
features are homogenized, as only cavitation on the global scale is simulated
in this work. An investigation has been carried out to see whether it is an
acceptable assumption to include the effect of texture within the flow factors
in this particular case.
For a fixed minimum film thickness (1 m) the stationary case problem
is solved with a deterministic texture (section B.2.2) and a homogenized
texture (section B.2.4). A smooth surface, with neither deterministic nor
126 PAPER B.

homogenized texture, is also solved for comparison. Fig. B.8 shows the
results of this study. As the pressure profile varies slightly across the periodic

107
3
Smooth
2.5 Homogenized
Deterministic
2
Pressure (Pa)

1.5

0.5

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/L

Figure B.8: Comparison of homogenized, deterministic and smooth solu-


tions

width of the deterministic contact several different cross sections are plotted
for the deterministic case.
It can be seen that while the homogenized profile is close to the deter-
ministic profile when compared with the smooth solution, the homogenized
solution is generally of a smaller magnitude than the deterministic solu-
tion. The predicted supported load for the deterministic case is 4.8 per cent
higher than for the homogenized case.
This can be explained because the homogenized solution assumes that
the texture repeats an infinite amount of times across the contact width;
however, in reality it only repeats 20 times. Therefore, the homogenized
solution underestimates the pressure build-up. It is suggested that while
it is acceptable to homogenize surface roughness that has a much smaller
wavelength than the contact width, the same is not true for the honed
surface texture which has a much larger wavelength.

B.4 Model Results


Due to the result found in section B.3 the model presented includes surface
texture on the global scale and plateau roughness (Fig. B.5) is incorporated
B.4. MODEL RESULTS 127

with flow factors on the local scale. The top compression ring is modelled
for the entire engine cycle for three different running conditions, 1200, 1600,
and 1900 r/min, at full load.
The results for minimum film thickness and friction as a function of
crank angle are presented in Fig. B.9 and Fig. B.10, respectively. It can

6
1200RPM
1600RPM
5
1900RPM
Film Thickness (m)

0
0 90 180 270 360 450 540 630 720

Crank Angle ( )

Figure B.9: Film thickness vs crank angle

be seen that with a lower engine speed the film thickness is reduced and
the friction increases. This is to be expected with lower engine speeds as
the gas pressure is higher, the temperature is higher, lubricant viscosity is
lower, and entraining motion speed is reduced.
A period of boundary lubrication occurs, with large increase in friction,
shortly after combustion at TDC (0 crank cngle) due to the very large
increase in combustion chamber pressure (Fig. B.7) loading the back of the
ring coupled with low sliding speed. It is crucial to incorporate plateau
roughness in the model in order to implement the boundary lubrication
model. If the plateaux were considered smooth, then no contact would ever
occur as the Reynolds equation is inappropriate due to the appearance of
fluid stress singularities at the point of contact.
It should be noted that fully flooded conditions are always assumed. In
reality this is unlikely to be the case, particularly during the mid-stroke
when a large film thickness is predicted. Simulations in the past have been
developed that allow for starved inlet conditions, such as those by Ma et
al. [24] and Liu and Tian [25]. With an oil availability model the film
128 PAPER B.

1,400
1200RPM
1,200 1600RPM
1900RPM
1,000
Friction (N)

800

600

400

200

0
0 90 180 270 360 450 540 630 720

Crank Angle ( )

Figure B.10: Friction vs crank angle

thicknesses would be reduced and the friction increased; however, for the
purposes of comparing deterministic and homogenized texture models fully
flooded conditions were considered adequate.

B.5 Conclusion
An investigation into modelling surface texture in the PRCL conjunction has
been carried out. Two dimensions, contact, roughness, texture, and time
dependence were included in the model. It was found that it is more accurate
to include the surface texture in the global scale with a deterministic solution
rather than within flow factors, and in doing so separate out the effect of
roughness and texture. This is contrary to the method employed in many
existing simulations, which lump together texture and roughness in one set
of flow factors.
This work only deals with different aspects of the simulation of the
PRCL contact and as yet no experimental validation has been conducted
for this particular case. However, using the homogenization technique and
the same method to simulate surface roughness has previously been well
proven experimentally by Sahlin et al. [30].
This approach also allows the effect of texture and roughness to be
investigated independently, and as they are each products of a different
stage of the honing process this may be beneficial in optimizing surface
B.6. ACKNOWLEDGEMENTS 129

texture.

B.6 Acknowledgements
The authors would like to thank Stiftelsen for Strategisk Forskning (SSF)
and ProViking for funding this work and Scania AB for discussions and for
providing technical data.
Paper C

The influence of AFM and


VSI techniques on the
accurate calculation of
tribological surface
roughness parameters

131
133

Tribology International, January 2013, vol. 57, pp. 242-250

The influence of AFM and VSI techniques on the accurate


calculation of tribological surface roughness parameters
A. Spencera1 , I. Dobrydenb , N. Almqvistb , A. Almqvista , R. Larssona
a
Division of Machine Elements, Department of Engineering Sciences and
Mathematics, Lulea University of Technology, SE-971 87, Lule
a Sweden
b
Division of Physics, Department of Engineering Sciences and Mathematics,
Lule
a University of Technology, SE-971 87, Lule
a Sweden

Abstract
Vertical Scanning Interferometry (VSI) may induce optical artefacts in sur-
face topography measurements. The influence of these optical artefacts on
the calculation of Rk surface roughness parameters, contact stiffness and
flow factors were studied. Two surface measurement techniques were used:
Atomic Force Microscopy (AFM) and VSI. Calibration grids were used to
make it easier to isolate the causes of these artefacts, while a real engineering
surface was used to compare these two techniques in an industrially applied
case. It was found that the optical artefacts have a large influence on all
the roughness parameters, contact stiffness and flow factors calculated on
the calibration grids. However, for the engineering surface the differences
between AFM and VSI measurements were much smaller.

Keywords: Surface Roughness, AFM, VSI, Flow Factors

1
Corresponding author.
E-mail address: andrew.spencer@ltu.se
134 PAPER C.

C.1 Introduction
It is necessary to characterise engineering surface topography to ensure that
any component leaving a manufacturing process will function to the required
performance level in the tribological contact that it is subjected to. To
achieve this, the surface must be accurately characterised and the appropri-
ate surface parameters calculated and compared with a pre-specified toler-
ance. Traditionally in engineering situations a stylus profilometer has been
used to measure a line profile on the surface [62, 63] and the average rough-
ness amplitude (Ra ), or some stratified surface parameters (Rk /Rpk /Rvk ),
calculated.
Nowadays however, with the advent of new, more advanced measuring
equipment 3D surface measurement techniques are more frequently being
used. These techniques are often based on optical rather than contact meth-
ods and bring great benefits if used correctly. On the one hand they capture
more data from the surface, however often these techniques bring about or
may cause some optical artefacts and errors which influence the surface data
captured. Little is understood about the new artefacts or errors that these
techniques bring to the surface measurements. It has already been indicated
in literature that the surface measurements performed with 3D surface mea-
surement techniques may differ from the ones performed by Atomic Force
Microscopy (AFM) [64, 65]. The latter technique is often used as a reference
technique to compare with. This research will focus on comparing the AFM
and Vertical Scanning Interferometry (VSI) techniques.
The investigation of the differences between AFM and VSI in such sur-
face characterisation, often focusing on standard surface roughness param-
eters, is a well researched area. It has previously been established that
roughness parameters, such as root mean square roughness, Rq , and av-
erage roughness, Ra , reveal a difference depending on whether the surface
has been measured with AFM or VSI [64, 65, 66, 67]. The differences in
measured roughness parameters were explained by the occurrence of optical
effects, introduced by the VSI technique, such as multiple light scattering,
crater effects, a response to different light angles, sensitivity to vibrations,
surface reflectivity [67] and diffraction effects at sharp edges [64]. It was also
found that sub-micron details can be smoothed due to optical system or de-
tector limitations [65]. Objective lenses with magnification of 40x or lower
were found to be unsuitable for accurate roughness measurements [65, 66].
These observed optical artefacts have been studied and described in
more detail in [68, 69, 70]. The occurring phase jumps and the batwing
effect were explained in terms of the interference between wave reflections
incident on the top and bottom levels of a surface asperity. The 2 error
in phase was attributed to dispersion caused by chromatic aberration. It
C.1. INTRODUCTION 135

was established that multiple scattering usually leads to overestimation of


the surface roughness by VSI. Rhee et al. [68] suggested a diffraction model
which partially explains the observed spikes occurring near edges and valleys
on the surface asperities.
AFM measurements ought to be performed with care as well. A too high
applied force by a tip may lead to scratching of the surface and excessive
wear of the tip while scanning the surface. Wear of the tip will cause a
decrease in resolution. The tip shape and tip radius define the accuracy
and resolution ability of the AFM in surface characterisation. In the case
when the angle of imaged asperities is close to the tip angle or the asperity
is even sharper than the tip, a convolution, or dilation, will take place and
the imaged features will possess the same slopes as the tip angle [71]. To
improve the obtained image and acquire real surface feature dimensions it
is necessary to perform a deconvolution, or erosion, operation [72, 73, 74].
Also, scanner drift must be taken into account, especially in the case of
measurements on relatively large areas and with the long acquisition time
typical for an AFM measurement. Piezo non-linearity and hysteresis may
distort the measurements too. This can be compensated for by the use
of a closed-loop scanner. The latter may reduce the total non-linearity of
the system to about 1% [75]. Another process which may lead to image
distortion is cross-talk between the x, y and z axes. This can be reduced
by either the use of close-loop scanners or separated orthogonal scanners
[75]. When all these issues are taken into account the AFM technique is a
very accurate tool for performing surface characterisation and the obtained
surface parameters can be used as a reference.
Although there are several investigations previously undertaken to com-
pare these two techniques, several experimental aspects are still unclear.
Thus, AFM is a precise tool to accurately characterise the surface rough-
ness but it is relatively slow in comparison to an optical technique. On the
other hand the VSI technique is much faster but as reported can produce a
poorer quality surface measurement in several specific cases. It is necessary
to separate and investigate all observed effects leading or causing to differ-
ences in the roughness parameters between these two techniques and clearly
answer in which case AFM and VSI should be used in measurements.
Also, 3D surface measurements are increasingly being used not only to
calculate roughness parameters, but also as an input of surface topography
into tribological simulations. Often contact mechanics simulations utilise
these 3D surface measurements, and flow factors, such as those described by
Patir and Cheng [23, 76] and Sahlin et al. [29, 30] are calculated from a 3D
surface measurement too. Therefore it is also important to understand how
the errors introduced by the measurement technique affect the calculation
of these parameters.
136 PAPER C.

C.2 Surfaces under investigation

Five different surfaces were used in this study, four commercially available
calibration grids and one engineering surface from a tribological contact.
The calibration grids used in this investigation were: a TGG 1 grid (NT-
MDT) consisting of triangular steps with an edge angle of 70 and 3 m
period formed on an Si substrate. Two grids, TGZ 2 and TGZ 3 (NT-MDT)
have rectangular steps formed on a layer of SiO2 and step heights of 112 nm
1.5 nm and 545 nm 2 nm respectively. The fourth grid has square holes
formed on an Si wafer with a period of 10 m. The engineering surface was
taken from a combustion engine cylinder liner. The material is grey cast
iron and is plateaux honed to give a cross-hatch like surface finish. In this
study an area of the smoother plateaux was investigated, as illustrated in
Fig. C.1. When measuring the engineering surface it was obviously desirable

1 1
(m) 0 (m) 0
-1 -1
80
40 60
(m) 20 80 (m) 40 60 80
20 40 60 20
20 40
0 0 (m) 0 0 (m)

(a) VSI (b) AFM

Figure C.1: Topography of the engineering surface measured with VSI and
AFM techniques.

to measure in the same location with both the VSI and AFM measurement
techniques. Therefore a system was devised in order to try and achieve this.
The surface was indented in four locations with the desired measurement
area at the centre of these four points. When the sample was placed in
the VSI or AFM it could be moved in the x and y direction and aligned
with the four indents on the sample so that the measurement took place in
the centre of them. The scratches were made sufficiently far away from the
desired measurement region so as not to damage the measured area. Due
to the small area being measured, approximately 100x100 m, it was not
possible to measure exactly the same area with both the VSI and AFM,
however the measurements are from an approximately similar region. It is
estimated that the centre point for each measurement is 100 m from the
specified location.
C.3. MEASUREMENT TECHNIQUES 137

C.3 Measurement Techniques


The five surfaces were investigated using two different measurement tech-
niques, Vertical Scanning Interferometry (VSI) and Atomic Force Microscopy
(AFM).

C.3.1 VSI Measurements


A Wyko NT 1100 Vertical Scanning Interferometer was used to measure
each of the five samples. A 50X objective lense and 1X Field of View lense
were used to give a measurement area of 124x94 m with 736x480 data
points. The lateral resolution of the equipment is limited optically to 0.55
m.
The tilt was removed by fitting and then subtracting a first order poly-
nomial. One issue worth remembering is that care should always be taken
when removing tilt from the measurement as it can alter the surface profile.
Chiffre et al. [77] demonstrate the effect of removing the slope of a first
order polynomial fitted to a sinusoidal profile. The slope removal causes
an asymmetry in the sinusoidal wave, which in turn alters the values of
the roughness parameters. To restore any missing data points from the
optical measurements the Delaunay triangulation method is employed in
combination with linear interpolation [78].

C.3.2 AFM Measurements


The engineering surface was characterised by an Ntegra Prima AFM (NT-
MDT) in contact mode with the probe scanner. The probe scanner was
used due to the relatively large and heavy reference surface for the AFM
sample scanner. The scanner calibration was performed on the same TGZ 2
and TGZ 3 grids. The calibration grids were imaged in contact mode with
the 100 m sample scanner (NT-MDT Z50251CLPI). The measured areas
were about 100x100 m with 1024x1024 data points for both the grids and
the engineering surface. The close-loop option for the scanners was used
to reduce the total non-linearity and cross-talk effect. A Silicon Nitride
cantilever of type PNP-DB (Nanosensors) was used with a length of 100
m, tip height of 3.5 m, force constant of 0.48 N/m and a tip radius of less
than 10 nm. A silicon probe CSC-21 (NT-MDT) was used with a tip height
of 7 m, tip radius of less than 20 nm and force constant of 2 N/m. The
scan velocity for the square grid was 87 m/s. The TGG 1, TGZ 2 and TGZ
3 grids were imaged at a scan velocity of 50 m/s. The engineering surface
was imaged with a scan velocity of 82 m/s for the PNP-DB probe and 78
m/s for the CSC-21 probe. Two types of probes were used to compare the
obtained roughness parameters and to ensure that the PNP-DB probe is
138 PAPER C.

fully suitable for this type of measurement. It was found that the obtained
roughness parameters for the engineering surface with the use of both probes
were the same. Since two types of the scanners were used in this study it
was necessary to check that the scanners possess the same resolution and
provide the same results. The TGZ 3 grid was imaged with the use of both
the sample and probe scanners. The obtained profiles were very similar
and varied slightly by 2 nm. The PNP-DB probe was used in almost
all measurements in this study. However, this probe was not completely
suitable for measurement of TGG 1. It was found that the probe could not
reach the bottom of the valleys between the plateaux. Because the probe
and surface have similar pitch and size scale, the side of the probe likely
contacted the surface before the tip reached the valley bottom. The plateau
to valley height obtained with the use of the PNP-DB probe was about 1.43
m while the real height achieved with the CSC-21 probe was 1.65 m.
The probe quality was checked after each measurement on the engineering
surface to verify that the probe was not worn significantly or damaged and
still provided the correct surface profile. The check was performed on the
TGZ 3 grid. The measured heights and profiles were exactly the same as
before measurements on the engineering surface.

C.4 Tribological Surface Parameters


In order to investigate the surface measurements two types of tribological
surface parameters will be calculated, Rk height averaged roughness param-
eters and flow factors, calculated by the homogenization method as outlined
by Sahlin et al. [29, 30]. Rk surface roughness parameters are widely used
to characterise tribological surfaces. Flow factors, although not widely used
to characterise surfaces, can be seen as surface roughness parameters. They
may be more effective than Rk paramaters when it comes to the characteri-
sation of lubricated surfaces. The effect of measurement technique on both
of these types of parameter will be investigated.

C.4.1 Rk parameters
The Rk parameters, based on the Abbott curve (Fig. C.2), attempt to in-
dividually define the peaks, valleys and plateaux of a surface with different
numerical parameters. To obtain the Abbott curve for a 2D profile or a
3D surface the height range is first divided into bins and the percentage
of material falling into each of these bins is plotted against the bin height.
This yields the height distribution of the surface. The Abbott curve is re-
lated to the cumulative distribution of surface heights and hence can be
obtained from the height distribution. To calculate the numerical param-
C.4. TRIBOLOGICAL SURFACE PARAMETERS 139

Figure C.2: Illustration of an Abbott Curve with Rk parameter definitions.

eters, a straight line must be plotted through the shallowest 40% of the
curve. Rk is defined as the change in height of this line across the width
of the graph, between 0% and 100% asperity height distribution. Rpk is
the difference between the highest point on the surface minus the height of
the straight line at 0% asperity height distribution. Similarly, Rvk is the
difference between the height of the straight line at 100% asperity height
distribution minus the height of the lowest point on the surface.
For a cylinder liner surface, like the engineering surface used in this
study, Rk can be used to describe the roughness height after the running
in process and Rvk the oil accumulation in the honing grooves. The Rk
parameters have been calculated for both measuring techniques for all the
surfaces under investigation.

C.4.2 Flow Factors and Contact Stiffness


For the surface measurements of the five samples, so called flow factors have
also been calculated using the homogenization method incorporating contact
mechanics for the mixed lubrication regime. Flow factors are functions of
separation between a measured surface and a theoretical smooth opposing
surface. They can be thought of as correction factors, which are used to
modify the Reynolds equation to incorporate the averaged effect of surface
roughness. Traditionally, when solving the Reynolds equation a very fine
mesh would be required to resolve surface roughness. By implementing flow
140 PAPER C.

factors, the problem can be considered smooth with the effects of roughness
incorporated within the flow factors in the homogenized Reynolds equation.
This allows a much coarser mesh to be used and therefore far less expensive
computations. Because of this flow factors could potentially be thought of as
a way of describing the effect that the surface roughness has in a tribological
problem.
As the separation between the measured surface and the smooth oppos-
ing surface is reduced, at some point the two will come into contact. At
this point a boussinesq-type elasto-plastic contact mechanics model [29, 30]
is used to calculate the average contact pressure between the asperities of
the measured surface and the opposing smooth surface. This data can be
plotted as a stiffness curve and is also calculated and investigated here.
A brief outline of the equations involved will be given below, but for a
full derivation Almqvist et al. [79] should be referred to. For this analysis
the time dependent term in the Reynolds equation will be neglected.
The homogenized steady state Reynolds equation can be written as;
(A0 (x1 , x2 ) p0 (x1 , x2 )) = B0 (x1 , x2 ) (C.1)
where p0 is the homogenized film pressure, = 6U and A0 and B0 are
flow factors:    
a11 a12 b1
A0 = and B0 = (C.2)
a21 a22 b2
To calculate the individual flow factors, three partial differential equations
with periodic boundary conditions must be solved over the measurement
domain, i.e., the cell of periodicity, Y = (0, l1 ) (0, l2 );
y h3 y 1

= 0 in Y, (C.3a)

1 (x, 0, y2 ) + 1 = 1 (x, l1 , y2 ) ,
1 (x, y1 , 0) = 1 (x, y1 , l2 ) .

y h3 y 2

= 0 in Y, (C.3b)

2 (x, 0, y2 ) = 2 (x, l1 , y2 ) ,
2 (x, y1 , 0) + 1 = 2 (x, y1 , l2 ) .

y h3 y 0

= y (he1 ) in Y, (C.3c)

0 (x, 0, y2 ) = 0 (x, l1 , y2 ) ,
0 (x, y1 , 0) = 0 (x, y1 , l2 ) .
C.4. TRIBOLOGICAL SURFACE PARAMETERS 141

Where h describes the clearance between the rough surfaces and 1 , 2 and
0 are local scale variables. Once 1 , 2 and 0 have been calculated they
can be integrated over to give the flow factors in Eq. (C.2) needed to solve
the homogenized Reynolds equation;

1 1
Z
a11 (x) = h3 dy, (C.4a)
l2 Y y1
1 2
Z
a12 (x) = h3 dy, (C.4b)
l1 Y y1
1 1
Z
a21 (x) = h3 dy, (C.4c)
l2 Y y2
1 2
Z
a22 (x) = h3 dy, (C.4d)
l1 Y y2
 
1
Z
b1 (x)
B0 = = he1 h3 y 0 dy (C.4e)
b2 (x) l1 l2 Y

Three of the calibration grids dont vary in the y1 -direction. In addition,


the entraining direction in the Reynolds equation is aligned with the x1 -
direction and the analysis can be reduced by one dimension. In this case,
Eq. (C.1) can be written as;
 
p0
a11 (x1 , x2 ) = b1 (x1 , x2 ) (C.5)
x1 x

Where a11 is the modification of the pressure induced flow due to surface
roughness and b1 is the modification of the shear induced flow due to the
surface roughness.
When the surfaces come into contact a bousinesq-type FFT accelerated
contact mechanics approach is utilised to calculate the deformation of the
surfaces, as explained by Almqvist et al. [80]. In this analysis, for each
of the surfaces flow factors, a11 and b1 , were calculated for 50 different
separations, 30 when the surfaces were not in contact (the hydrodynamic
lubrication regime) and 20 in contact (mixed lubrication). The contact
mechanics model was used to calculate the deformed surface profile and then
flow factors calculated for the mechanically deformed, mixed lubrication,
surfaces. In order to perform the contact mechanics simulations, material
properties are required. For the calibration grids, the material was silicon,
with an elastic modulus of 185 GPa and a Poissons ratio of 0.26. For the
engineering surface, the material was grey cast iron with an elastic modulus
of 92 GPa and a Poissons ratio of 0.24.
142 PAPER C.

C.5 Results and Discussion


It has been discussed in the introduction that optical artefacts may occur
while measuring surfaces with the VSI technique. An attempt has been
made to separate and investigate any artefacts by performing characterisa-
tion of well defined surface profiles. The measurements done on the stepped
grids with heights of 112 nm and 545 nm using AFM showed good agree-
ment with the specification given by the manufacturer of the surface. It
was expected that the VSI measurements on the 112 nm grid would give an
overestimated step height value and it can be easily seen in Fig. C.3 that
the height profile is much bigger than in the AFM measurement. Also, the
step profile with VSI exhibits more of a sinusoidal shape. The VSI profile
looks quite smooth and this may have been caused by optical and detector
limitations. The narrower top and bottom shape of the steps is probably
caused by an overlapped batwing effect on the edges as the result of the
multiple interference following diffraction from the sharp edge.

0.15

0.1

0.05
z (m)

-0.05

-0.1
AFM VSI
0 5 10 15 20 25 30
x (m)

Figure C.3: Cross-section of 112 nm grid measured with AFM and VSI
techniques, VSI measurement shows increased height with sinusoidal-like
profile.

It was unexpected that the measured height of the steps by VSI was
much less in comparison to the AFM measurement for the case of the 545
nm stripe grid, Fig. C.4. A narrower top shape of the steps in the VSI
measurement is observed as in case of 112 nm stepped grid. It appears that
it is induced for similar reasons as described above for the 112 nm grid.
There are clear spikes (or kinks) in the valleys of the steps. It is assumed that
C.5. RESULTS AND DISCUSSION 143

they might be caused by multiple scattering and the interference between


reflected waves. The kink at the bottom looks similar to the left upper edge
of a previous step. It is possible that accurate height determination by VSI
is affected by a resolution limit somewhat. Since the period of the plateaux
is only 1.5 m, the optical artefacts that occur due to interference might
overlap leading to a sinusoidal profile.

0.4

0.3

0.2

0.1
z (m)

-0.1

-0.2

-0.3
AFM VSI
-0.4
0 5 10 15 20 25 30
x (m)

Figure C.4: Cross-section of 545 nm grid measured with AFM and VSI
techniques, VSI shows reduced height of measured steps and kinks in the
valleys.

The grid with a periodic square pattern was investigated next, the square
holes have a depth of approximately 100 nm and a period of 10 m in both
the y1 and y2 directions, a cross-section is illustrated in Fig. C.5. This grid
is very similar to the one investigated by Gao et al. [70]. The obtained grid
profile is analogous to that observed in [70] also. The observed spikes at
the edges and valleys are due to the batwing effect. Clearly, the top and
bottom of the holes are overestimated by VSI.
To continue the investigation further the triangular grid was imaged by
both techniques, shown in Fig. C.6. The VSI technique exhibited a good
ability to measure such features. The step height was a little bigger with
the VSI measurement than the AFM measurement and this is due to the
expected batwing errors on the sharp peaks. The kinks at the bottom of
the valleys look very similar to the top of the previous steps. However, they
are not completely symmetrical and are possibly due to multiple reflections
from the side walls.
It is believed that because of the resolution limitation overlapping of the
144 PAPER C.

0.3

0.2

0.1

0
Z (m)

-0.1

-0.2

-0.3
AFM VSI
-0.4
0 5 10 15 20 25 30
X (m)

Figure C.5: Cross-section of square grid measured with AFM and VSI tech-
niques. The height overestimation and batwings are clearly seen in the VSI
measurement.

1.5

0.5
Z (m)

-0.5

AFM VSI
-1
0 5 10 15 20
X (m)

Figure C.6: Cross-section of triangular grid measured with AFM and VSI
techniques. VSI measurement shows increased stripe height and kinks in
the valleys.

batwing effects occur, especially in the case of 1.5 m separation between


neighbouring edges of the calibration grids of Fig. C.3 and Fig. C.4. The
overlapping did not happen only in case of 10 m period and only on the
C.5. RESULTS AND DISCUSSION 145

Table C.1: Rk parameters calculate for surfaces

Sa (nm) Sk (nm) Spk (nm) Svk (nm)


Surface
AFM VSI AFM VSI AFM VSI AFM VSI
112 nm 50.4 71.3 30.2 77.3 99.9 11.1 105 148
545 nm 232 89.1 53.5 103 17.2 169 514 24.5
Square 39.0 91.6 10.9 64.3 45.1 220 101 521
Triangular 450 539 723 667 1002 1407 43.2 74.3
Eng. Surf. 135 144 342 394 492 804 744 852

bottom part where the flat area has a length of 6 m. This overlapping
might be the reason why 112 and 545 nm stepped grid profiles look very
similar.
The observed feature, the kinks, might be explained by multiple scat-
tering from the side walls or as a result of interference phenomena.

C.5.1 Rk Parameters
The roughness parameters calculated from the AFM and VSI surface mea-
surements are given in Table C.1.
In general the VSI technique overestimated the value for average surface
roughness, Sa . This is true for the 112 nm, square and triangular grids and
the engineering surface. However, in the case of the 545 nm grid the VSI
measurement gave a much smaller Sa parameter than the AFM measure-
ment. It is quite clear from Fig. C.4 that the average surface roughness
would be lower due to inaccuracies in the measurement. VSI significantly
overestimates the Sa parameter for the square grid with a value 135% greater
than from the AFM measurement. This is caused by the batwing effect as
discussed previously. The same is true of the 112 nm height grid with a
value 41.5% greater than from the AFM measurement, which is caused by
overlapping batwing effects at the edges. In the case of the triangular grid
VSI works more effectively than with the other grids and just overestimates
the Sa value by 19.8%. Although the difference between VSI and AFM
in measuring Sa parameters for the calibration grids is significant, for the
engineering surface (Fig. C.1) VSI exhibited fewer optical artefacts than for
the calibration grids, the difference only being 6.7%.
Comparing the Sk stratified surface parameters, other differences be-
come apparent between the AFM and VSI measurements. The core rough-
ness, Sk , is overestimated with the VSI technique for the 112 nm, 545 nm
and square grids, but the values are approximately the same for the trian-
146 PAPER C.

gular grid and engineering surface.


The VSI technique always overestimates the Spk peak roughness, how-
ever there is one exception, the 112 nm striped grid. The AFM measured
value of Spk is signifcantly greater than the VSI value, 99.9 nm in compar-
ison with 11.1 nm. This is due to an overestimation of the Spk value by
the AFM technique. The Spk value, as discussed previously, is a measure
of the peak height, the distance between the highest peak and the upper
level of the core roughness zone. On the AFM image there were a handful
of spikes, caused by dust or some other contamination. Although they oc-
cupy an insignificant amount of the measurement area and have very little
affect on the calculation of other roughness parameters, contact stiffness
or flow factors, they have a huge affect on the calculated Spk parameter.
This shows that the Spk parameter is highly sensitive to small errors in the
measurement and must be used with care.
Finally, examining the Svk parameter, there are two major differences
betweem the VSI and AFM measurements. For the 545 nm grid, the AFM
value is huge compared to the VSI value, 514 nm compared to 24.5 nm for
the VSI grid. This is because for the AFM measurement, the core region
is considered to be the top plateau surface, with the 545 nm deep grooves
the valleys that Svk is calculated over, hence a value of 514 nm, which is
approximately the stripe depth. This is the correct value for the parameters
for this surface. However, with the VSI grid, the profile is more sinusoidal
due to the previously discussed measurement errors. Therefore, the core
region sits at a lower level, giving the surface both peaks and valleys, and
a lower Svk is calculated. The second difference with Svk parameters comes
with the square grid, where this time the VSI measurement gives a much
higher value. This is more simply explained by the batwings occuring at the
bottom of the valleys and causing a greater depth than is there in reality,
shown in Fig. C.5.

C.5.2 Flow Factors and Contact Stiffness


The flow factors presented here are plotted against mean separation, ,
which is defined in Fig. C.7.
The pressure induced flow factor, a11 , is cube rooted when plotted on
the graphs to make it comparable in magnitude to the shear induced flow
factor, b1 , to allow them to be plotted on the same axis. One issue when
evaluating the data is how to define the surface separation. In these plots
, is the separation from the flat counter surface to the mean plane through
the measured surface. However, due to the optical artefacts, the mean plane
on the undeformed VSI and AFM surfaces will be in different locations. It
is assumed that under very high loads and deformations the separation is
C.5. RESULTS AND DISCUSSION 147

Hydrodynamic Lubrication

Mixed Lubrication - Surfaces in Contact

Surfaces
Mean plane of undeformed rough surface
Undeformed rough surface

Figure C.7: Definition of Mean Separation, , for both hydrodynamic and


mixed lubrication scenarios.

comparable, therefore the VSI separation is corrected so that at the highest


load the deformation is the same as with the AFM surface. This makes the
difference in separation easier to evaluate from the figure as the surfaces are
unloaded.
The effect of optical artefacts from the VSI technique on flow factor and
contact stiffness surface parameters is first evaluated for the square grid, as
illustrated in Fig. C.8.
Fig. C.9 shows the contact stiffness, mean contact pressure plotted
against separation, for the square grid. It is observed that the VSI sur-
148 PAPER C.

0.4 0.1
(m) 0 (m) 0
-0.4 -0.1
30 30
20 20
30 (m) 10 30
(m) 10 20 20
0 0 10(m) 0 0 10(m)

(a) VSI (b) AFM

Figure C.8: Square grid measured with both VSI and AFM.

face comes into contact with the flat counter surface first and there is a
more gradual increase in pressure indicated by a shallower gradient.

1000
900
800 VSI
Contact Pressure (MPa)

700 AFM

600
500
400
300
200
100
0
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
Separation (m)

Figure C.9: Sample stiffness for the square grid showing a lower calculated
stiffness from the VSI measurement.

The difference in contact stiffness can be explained by the optical arte-


facts present in the VSI measurement. The artefacts representing batwing
like artificial peaks lead to the surface coming into contact with the flat
plane sooner and provide a less stiff initial contact, whereas with the AFM
measurement the two surfaces come into contact later but with a much
higher initial stiffness. This is shown by a higher gradient of the AFM
stiffness curve where contact first occurs.
The a11 flow factor, as defined in Eq. (C.4a), is used to modify the
pressure induced flow in the Reynolds equation due to the effect of surface
C.5. RESULTS AND DISCUSSION 149

topography. If the flow is restricted in the x-direction, then the flow factor
will be of a lower value. The b1 flow factor, as defined in Eq. (C.4e), is used
to modify the shear induced flow due to the effect of surface topography.
Fig. C.10 shows the pressure and shear flow factors for the square grid.

107
5
4.5
AFM - (a11 )3
4
3.5 VSI - (a11 )3
Flow Factor

3
2.5 AFM - b12

2
VSI - b12
1.5
1
0.5
0
-0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5
Separation (m)

Figure C.10: Flow factors calculated for the square grid, showing more
restricted flow at higher separations and less restriction at lower separations
in comparison to the AFM measured surface.

For larger separations the flow factors calculated for the VSI surface are
of a lower value, meaning that the flow is more restricted, this is due to
the batwing effect creating additional peaks rather than a smooth plateau
surface. As the surfaces come closer together, the flow is gradually restricted
for the VSI surface. The AFM surface however, reaches a point where there
is sharp change in gradient and the flow factor has almost zero gradient
afterwards. This occurs when the flat counter surface comes into contact
with the flat plateaux of the square grid AFM measurement, the same point
at which there is a sharp increase in asperity contact pressure, Fig. C.9. It
can be concluded from these results that at large and small separations the
VSI and AFM techniques produce similar results, however between the two,
in the mixed lubrication regime around where contact occurs between the
surfaces, the behaviour is different. The transition is much smoother with
the VSI measured surface, due to the batwing optical artefacts, however this
is not true for the AFM measurement which has a much sharper transistion
and is more representative of what the surface really looks and in theory
behaves like.
The stiffness curves for the triangular grid, Fig. C.11 are very similar.
150 PAPER C.

The only difference is that the stiffness under low load is less for the VSI
measured surface, as illustrated by a shallower gradient of the stiffness curve.
This is caused by the steeper peaks of the VSI being less stiff than the
broader peaks of the AFM measurement.

1000
900
VSI
800
Contact Pressure (MPa)

700 AFM

600
500
400
300
200
100
0
0.4 0.5 0.6 0.7 0.8 0.9 1 1.1
Separation (m)

Figure C.11: Sample stiffness for the triangular grid showing similar stiffness
for both AFM and VSI measured surfaces.

The difference in flow factors is more pronounced for the triangular grid,
Fig. C.12, than was seen with the square grid, Fig. C.10. The flow factors
for the VSI measured surface are much lower, indicating a reduction in flow,
in comparison to the AFM measured surface. This can be explained because
the peaks are much steeper on the VSI measured surface, and there is no
way to flow around them as they completely cross the width of the surface.
This is unlike the batwings seen on the square grid VSI measurement, which
do not create a complete barrier to flow in the y1 direction.
Comparing the stiffness curves for the engineering surface, Fig. C.13, the
VSI measurement shows the same trend as has been previously illustrated
by the square grid stiffness curve, Fig. C.9. Although it is not as clearly
observed with this measurement of an engineering surface, this can most
likely be explained as being caused by the same phenomenon as on the
square grid, that optical artefacts on the surface cause sharper peaks that
create a less stiff surface when low loads are applied.
Examining the flow factors for the engineering surface, Fig. C.14, there
is little difference between the flow factors calculated with the VSI and AFM
measurement techniques, they are almost identical. This is very encourag-
ing, as for this particular surface it shows that the flow factor parameters
C.5. RESULTS AND DISCUSSION 151

106
1.2

1 AFM - (a11 )3

0.8 VSI - (a11 )3


Flow Factor

0.6 AFM - b12

0.4 VSI - b12

0.2

0
0.8 0.9 1 1.1 1.2 1.3 1.4 1.5
Separation (m)

Figure C.12: Flow factors calculated for the triangular grid, showing large
difference between values calculated for VSI and AFM measurements.

2000
1800
AFM
1600
Contact Pressure (MPa)

1400 VSI

1200
1000
800
600
400
200
0
-2 -1.5 -1 -0.5 0 0.5 1
Separation (mm)

Figure C.13: Sample stiffness for the engineering surface showing a lower
calculated stiffness from the VSI measurement.

can be considered reasonably independent of the measurement technique.


152 PAPER C.

106
1
0.9
AFM - (a11 )3
0.8
0.7
VSI - (a11 )3
Flow Factor

0.6
0.5 AFM - b12
0.4
VSI - b12
0.3
0.2
0.1
0
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
Separation (m)

Figure C.14: Flow factors calculated for the engineering surface showing
little difference between the values calculated from the VSI and AFM mea-
surements.

C.6 Conclusion
In this research the optical artefacts introduced by the VSI measurement
technique have been investigated for a range of calibration grids and a real
engineering surface. The AFM measurement technique was used as a refer-
ence to compare the measurements to.
Several optical artefacts, including batwing effects and phase shifts, were
observed on several of the surfaces. The calibration grids made it easier to
isolate the causes and effects of these artefacts, which was more difficult
with a real engineering surface.
The effect of these artefacts on the calculation of both Rk surface rough-
ness parameters, contact stiffness and flow factors calculated through the
homogenization process was investigated. It was found that the optical
artefacts could have a large influence on all of the roughness parameters,
especially on the striped and square calibration grids, where the artefacts
dominated the underlying surface features on the VSI measurement. An
analysis of the contact stiffness showed that when the surfaces come into
contact under low loads the increase in contact stiffness is more gradual with
the VSI measurements due to the optical artefacts creating artificial peaks.
For the flow factors there was little difference between the measurement
techniques for the engineering surface, however for the calibration grids the
extra peaks caused by the artefacts caused an additional constriction for the
C.7. ACKNOWLEDGEMENTS 153

flow across the surface, leading to a lower value of calculated flow factor.
These preliminary data suggest that it may be acceptable to use the
VSI measurement technique for certain engineering surfaces for the calcu-
lation of roughness parameters, contact stiffness and flow factors. This is
largely due to real engineering surfaces having shallower gradients than the
extreemes found on the calibration grids. However, care must be taken
with surface measurements with a small, well defined surface features, such
as that which might be produced by laser etching on an engineering sur-
face. With these measurements the VSI technique is prone to overestimate
roughness parameters and flow factors, and a lower contact stiffness at the
moment of contact. However, the exact effects depend on the wavelength
and depth of the features on the surface and must be investigated carefully
for each case.

C.7 Acknowledgements
The authors would like to thank Stiftelsen for Strategisk Forskning (SSF)
and ProViking for funding this work and Scania AB for discussions and
for providing technical data. Also The Kempe Foundations SMK -2546 is
thanked for funding the SPM. The authors thank Dr. Per Gren of the Divi-
sion Of Experimental and Fluid Mechanics at LTU for valuable discussion
and suggestions and Evelina Enqvist of the Division of Machine Elements
at LTU for valuable feedback and suggestions on the manuscript.
Paper D

Experimental and Numerical


Investigations of Oil Film
Formation and Friction in a
Piston Ring-Liner Contact

155
157

Proceedings of the Institution of Mechanical Engineers, Part J: Journal


of Engineering Tribology, February 2013, vol. 227 no. 2, pp. 126-140

Experimental and Numerical Investigations of Oil Film


Formation and Friction in a Piston Ring-Liner Contact
E. Y. Avan,a1 , A. Spencerb , R. S. Dwyer-Joycea , A. Almqvistb , R.
Larssonb
a
The Leonardo Centre for Tribology, Department of Mechanical Engineering,
University of Sheffield, Sheffield, UK
b
Division of Machine Elements, Department of Engineering Sciences and
Mathematics, Lulea University of Technology, SE-971 87, Lule
a Sweden

Abstract
The piston ring-cylinder liner contact is a major source of the total para-
sitic losses in an internal combustion engine. The lubrication process of this
contact highly influences the amount of friction, oil consumption and wear
that occurs. In this work, a reciprocating test rig combined with an ultra-
sonic film thickness measurement system was developed and then used for
tribological investigation of the piston ring-cylinder liner contact under ide-
alised cold conditions. A special piston ring and cylinder liner holder were
designed and five sensors were glued on to the back side of the liner speci-
men. Ultrasonic reflections captured by the sensors, used to obtain the film
thickness, and friction were continuously recorded as the piston ring section
reciprocated over the liner. Several experiments were performed at differ-
ent speed and load conditions. Furthermore, a numerical model has been
developed to predict film thickness and friction in all lubrication regimes.
The experimentally measured film thickness and friction were compared
with the output from the numerical model and good correlation was found.
The parameters affecting the accuracy of the ultrasound measurements and
numerical simulations of film thickness and friction are then discussed.
Keywords: Piston ring, cylinder liner, friction, reciprocating test ring, oil film
thickness, cavitation, mixed lubrication, ultrasonic film thickness measurement

1
Corresponding author.
E-mail address: e.y.avan@shef.ac.uk
158 PAPER D.

D.1 Introduction
The piston ring-cylinder liner conjunction in an internal combustion en-
gine is very important for the automotive industry in the drive to increase
engine efficiency and achieve the emission reduction targets proposed by
authorities. Piston rings operate in a range of tribological conditions, from
the boundary to hydrodynamic lubrication regime. They should provide
a mechanical seal between the combustion chamber and engine crankcase,
conduct the heat of the piston to the water-cooled cylinder liner, and also
distribute the lubricating oil along the cylinder liner surface. The optimum
lubrication of the piston rings is necessary to reduce the friction and also
limit oil consumption.
Andersson [4] showed the distribution of fuel energy usage for a medium
sized passenger car and reported that the piston assembly is a major source
of losses and responsible for about 40-50% of the total mechanical losses
during an urban cycle. The work by Spearot [8] showed that the piston ring
contribution to the entire friction loss is 19% in a light duty vehicle. Many
laboratory tests have been carried out to measure frictional losses between
the piston assembly and cylinder liner using the floating liner method
[52, 53, 81] or indicated mean effective pressure (IMEP) method [50, 51].
Although these tests are very useful and representative of engine conditions,
intensive modifications and instrumentation of the test engine inevitably
increase the time and the cost of laboratory tests. They also do not give a
clear picture of the contributions to frictional losses of the individual parts
(i.e. piston rings or piston skirt). Particularly in a fired engine, there are
many factors effecting the lubrication of the piston assembly such as blow-
by, dynamics of the piston and ring and thermal deformations.
Researchers designed and developed more simplified test rigs to inves-
tigate the piston ring lubrication mechanism and the related friction phe-
nomena. The designs of these test rigs are wide ranging and depend on
the focus of the research. Although whole piston and cylinder liner assem-
blies have been used in certain designs, i.e. in the work by Tan and Ripin
[82], a typical test rig configuration consists of segments of a piston ring
and cylinder liner where one of them reciprocates and the other is kept sta-
tionary [83, 84, 7, 26]. These components are tested for different operating
parameters, i.e. speed, load, viscosity and lubricant rate. Since most of
these parameters are controllable, bench tests provide detailed information
about how the different parameters influence the piston ring lubrication.
However, there are several drawbacks of bench tests. The stroke is often
an order of magnitude less than that in an engine. Therefore, because of a
shorter stroke the maximum speed and acceleration seen by the piston ring
is signifcantly less. These issues are investigated by Lee and Chittenden
D.1. INTRODUCTION 159

[85].
Akalin and Newaz [26] developed a reciprocating test rig to simulate the
engine piston ring and liner contact. A ring holder was modified from the
ring holder design developed by Hartfield et al. [86]. The friction force be-
tween the piston ring and liner was measured using strain gauges placed on
a cantilever beam connecting the ring holder to a loading arm. They also
developed a mixed lubrication model to predict the lubrication and friction
characteristics of the piston ring and liner contact. Their results highlighted
that temperature, surface roughness and running speed were important pa-
rameters for identifying the lubrication regime. However, normal load had
only a marginal effect on the friction coefficient under the simulated mixed
lubrication condition. In general, the analytical results and the bench test
results were well matched. Like Akalin, Bolander et al. [7] developed a test
rig to correlate with a numerical model of the piston ring-liner interface.
However, in their design a three axis force transducer was used to measure
the normal, tangential and side loads generated on the piston ring segment.
Depending on the test rig operating conditions, the entire range of lubri-
cation regimes, from boundary to full-film hydrodynamic lubrication, were
observed at different points in the stroke. As expected the highest friction
occurred in the mixed and boundary lubrication regimes. In these studies,
the simultaneous and comparative measurement of lubricant film thickness
has not been achieved.
The lubricant film formed between the piston ring and cylinder liner is
very thin and the measurement of this essential variable is difficult. In the
literature, several methods have been applied to measure oil film thickness
in piston rings such as capacitance [87, 88, 89], resistance [90] and the laser
induced florescence method [91, 92]. All of these methods have had some
degree of success; however, these methods require the need to penetrate the
cylinder liner in order to access the ring-liner conjunction.
The ultrasound technique is based on sensing the reflections from the
ring-liner contact; therefore the ultrasonic sensors do not have to be mounted
flush with the inner liner surface. This provides localised non-invasive mea-
surements. Recently, the method was applied to a motored engine to mea-
sure the film thickness between the piston ring and cylinder liner [56]. How-
ever, there occured a resolution problem due to the sensor size (i.e. the
ultrasound pulse was emitted over a larger area than the piston ring con-
tact). In this study, smaller piezo-electric sensors were used to enhance the
spatial resolution and the piston ring-contact was simulated using a Plint
high frequency reciprocator. It is acknowledged that other techniques can
more realistically recreate the piston ring-cylinder liner contact found in
an engine, however the primary goal of this work is to compare ultrasound
film thickness measurements with a numerical model rather than accurately
160 PAPER D.

recreate engine running conditions. Therefore, the plint reciprocator was


used to provide a controlled, easily accessible and repeatable experiment
with known and measurable boundary conditions to be used as input into
the numerical model. The average minimum film thickness was measured
at five different locations and the ring-liner friction force was measured over
the entire cycle. Furthermore, a numerical model applicable for simula-
tion of ring-liner contacts operating in boundary to full-film lubrication is
used to predict the pressure profile, film thickness and friction force. The
experimental results were compared to the numerical predictions.

D.2 Test Apparatus


In this section the test equipment, instrumentation and test specimens will
be introduced as well as the preparation of the ultrasound sensors.

D.2.1 Piston Ring-Liner Simulator


A Plint TE-77 high frequency reciprocating tribometer was instrumented
with an ultrasonic pulsing system. The TE77, schematically shown in
Fig. D.1, is a flexible tribometer that has been used in several other piston
ring-liner wear studies [86, 93]. The machine configuration involves a special
adapter oscillating mechanically over a fixed liner section. The adapter re-
taining a section of piston ring is loaded against the liner section by a spring
balance through a lever and stirrup mechanism. The normal force is trans-
mitted directly onto the moving specimen by means of a needle roller cam
follower on the adapter and the running plate on the loading stirrup. The
oscillations are produced by a variable speed motor with an eccentric cam,
scotch yoke and guide block arrangement. The stroke of the reciprocating
motion is 15 mm. It should be noted that this is an order of magnitude less
than the stroke of the engine that the components are taken, which is 160
mm. A stable oscillating frequency is maintained by a tacho generator feed-
back system. The whole assembly is mounted on flexible supports, which
allows free movement in the horizontal reciprocating axis. A stiff piezo-
electric force transducer connected to the assembly measures the friction
force in the reciprocating direction. The force transducer was calibrated
using a pulley fixture system which applied a calibrating load aligned with
the normal axis of ring-liner contact, parallel to the liner surface. Due to
the liners horizontal position, it is useful to identify the dead centres where
the oscillating ring stops. The dead centre closest to the friction sensor is
named top dead centre (TDC) and the one closest to the adapter mounting
hole is named bottom dead centre (BDC) in this study.
D.2. TEST APPARATUS 161

Figure D.1: Plint TE-77 simulated piston ring-liner contact schematic.

Test Specimens
A pair of cast-iron liner and ring segments from the same heavy duty diesel
engine were used to create the contact. The liner specimen was cut from
a production cylinder liner with a bore diameter of 130 mm. The liner
specimen, 50 mm in length and 20 mm in width, has a cross-hatched surface
finish typical of a plateau honing process, as illustrated in Fig. D.2. The
used top compression ring from the same engine was sectioned into a length
of 45 mm. The width of the piston ring is 3 mm and it has an asymmetric
barrel shaped face as illustrated in Fig. D.3 with a chrome surface coating.
Both the cylinder liner and piston ring have been run-in in a fired engine
prior to these tests. The cylinder liner has a mean roughness value Ra of
0.224 m and the piston ring has an Ra value of 0.066 m.

Conformability of the Ring-Liner Contact


Consistent conformability between the piston ring and cylinder liner is nec-
essary from test to test in order to produce repeatable results. In the engine,
the ring is able to conform to the liner due to its inherent tension and free-
dom to move. However, in simulated test rigs, the uncompressed piston ring
(the diameter of the ring is bigger than the diameter of the liner) would mean
162 PAPER D.

Figure D.2: 3D image of cylinder liner surface used in the tests, measured
with a Confocal Microscope.

that in its free state the piston ring would only make contact at the edges of
the liner sample. To prevent this, a special ring holder manufactured from
an original production piston was designed and attached to the adapter (see
Fig. D.4). The conformability of the ring-liner contact was adjusted by two
slotted plates located at either side of the ring holder and a grub screw in

45
40
35
30
Height (m)

25
20
15
10
5
0
0 0.5 1 1.5 2 2.5 3

Width (mm)

Figure D.3: Piston ring profile measured by a profilometer with a Gaussian


filter applied.
D.2. TEST APPARATUS 163

Figure D.4: A photograph showing: (a) Piston Ring sections, (b) Cylinder
liner section with ultrasonic transducers attached to the back side, (c) Oil
bath and cylinder liner holder, (d) Ring adapter.

the centre which pushed the ring from behind. The liner specimen was held
in the lubricant bath and secured by six grub screws allowing for axial and
lateral alignment of the liner. Fujifilm Prescale pressure measuring film was
used to check the conformability of the contact. The Fujifilm paper indi-
cating the stages of the ring-liner contact from the initial to final set-up are
shown in Fig. D.5. The process of obtaining conformability is clear from the

Figure D.5: Fujifilm Prescale pressure measuring film used during the con-
formability adjustment, from the initial attempt (left) through to the con-
formal contact (right).

initial to final set-up. The ring-liner contact was initially unconformal such
164 PAPER D.

that a more dense pink colour appeared at the right hand-side of the liner,
indicating a higher contact pressure. After a few adjustments, a more con-
formal contact where the colour is evenly distributed over the liner surface
has been obtained.

D.2.2 Ultrasonic Sensors

10 MHz piezoelectric crystals with a width of 1.3 mm and a length of 2.5


mm were used to generate ultrasonic waves. The bonding surface should
be free from pits and irregularities and the back of the liner sample was
therefore ground to provide a flat mounting surface for the piezo-crystal
sensors. An epoxy resin formulated specifically for bonding strain gauges
was applied and five piezo-crystal sensors were glued on to the back side of
the liner by means of a guidance template indicating the sensor positions.
The electrodes of the sensor were connected to a small coaxial cable, with
a diameter of 0.4 mm, and covered by a protective layer of silicone.
These five ultrasonic sensors placed at the back of the liner were almost
equally distributed along the axial length of the stroke. The central sensor
was aligned to the middle of the stroke by adjusting the liner holders grub
screws in the axial direction. Thus, the ultrasonic sensors were kept in the
region swept by the piston ring and they were able to record the oil film
data at five locations between the dead centres. The sensor positions over
the liner are given in Table D.1 where the measurements are taken relative
to TDC. The velocity of the piston ring as it passes over each sensor is also
given for a range of operating speeds.

Table D.1: Sensors location over the stroke

Dist. from TDC Speed, in Hz (Velocity in m/s)


TDC 0.0 mm 7.5 Hz 10 Hz 12.5 Hz 15 Hz 17.5 Hz
1 1.4 mm 0.21 0.28 0.35 0.41 0.48
2 4.1 mm 0.32 0.42 0.52 0.63 0.73
Sensor

3 7.3 mm 0.35 0.47 0.59 0.71 0.82


4 10.4 mm 0.33 0.43 0.54 0.65 0.76
5 13.3 mm 0.23 0.30 0.38 0.45 0.53
BDC 15.0 mm
D.2. TEST APPARATUS 165

D.2.3 Instrumentation
An ultrasonic pulsing unit embedded into a dedicated computer was used.
This unit consists of an ultrasonic pulsing and receiving card (UPR) which
is equipped with 8 channels and a maximum achievable pulse rate of 80k
pulses/second on a solitary channel. Receiver gain range is between -40 dB
and +110 dB and the receiver bandwidth is from 0.1 to 25 MHz. Each
of the ultrasonic sensors was individually connected to the pulsing unit,
totalling 5 channels with a pulse repetition rate of 15 pulses/second for
each. The sensors were excited by short duration high voltage signals and
thus ultrasonic pulses were generated. These pulses propagated through the
liner specimen. The system operated in a pulse-echo mode, meaning that
the reflected pulses from the ring-liner conjunction were also received by
the sensors. Each of ultrasonic reflection signals was digitised at 100 MHz
with a 12 bit resolution. The digitised data was recorded to hard disk in
binary file format and then a post-processing software program translated
the data to oil film thickness. The piezo-electric transducer (Kistler type
9203) with a range of 500 N and normal sensitivity of 50 pC/N was used to
measure the friction force. The charge amplifier (Kistler type 5007) converts
the charge produced by the transducer into proportional electrical signals
with a resolution of 0.001 N. The transducer was calibrated with a known
load before the experimental stage. Friction data output was logged to the
computer hosting the ultrasonic pulsing unit.

D.2.4 Lubricant
The lubricant used in this study was a pure base oil without an additive
package. The physical properties of the lubricant are given in Table D.2.
The liner specimen was fully immersed in the oil. This does not represent

Table D.2: Properties of the lubricant used in the tests.

Density, 843.4 kg/m3


Kinematic Viscosity at 40C, 37 cSt
Kinematic Viscosity at 100C, 6.5 cSt
Longitudinal wave velocity, c 1440 m/s

the real lubricant condition in the engine, which would normally be signif-
icantly less. However, it ensures that the inlet is fully flooded allowing for
good, accurate comparisons with the numerical model. It also assists in
maintaining a stable temperature of the liner surface during the short tests.
166 PAPER D.

The oil bath temperature was logged at a stable 22C throughout the tests.
This is clearly unrepresentative of real engine running conditions, where
the liner temperature can reach approximately 200C around TDC. How-
ever, it was decided that running at a lower temperature, giving a higher
viscosity and therefore higher film thickness would go some way towards
compensating for the lower entraining speeds in the test rig compared to
the real engine operating conditions. The ASTM D31 equation was used to
calculate the lubricant viscosity at 22C;

log(log( + 0.7)) = A B log(T + 273.15), (D.1)

where is viscosity, A and B are constants and T is temperature. Using


the values in Table D.2, A = 8.8686 and B = 3.4743, giving a viscosity of
85.66 cSt or 0.072 P a s. In the contact it is predicted that the temperature
of the lubricant will be higher than that of the bulk lubricant due to shear
heating. Therefore, the viscosity will be less in the contact than the value
that has been calculated here. However, in the absence of a measured
contact temperature or thermal model the bulk lubricant temperature and
hence viscosity will be used in the numerical model.

D.3 Ultrasonic Oil Film Measurement


In this section the method for processing the ultrasound signals will be
discussed.

D.3.1 Background
The proportion of an ultrasonic pulse that is reflected from a perfectly
bonded interface is known as the reflection coefficient and varies with the
acoustic properties of the matching materials. This proportion is given by
Eq. (D.2), where z1 and z2 are the acoustic impedance of the materials on
either side of the interface.
z1 z2
R12 = (D.2)
z1 + z2
However, the ring-oil-liner layer can be modelled as a quasi-static spring.
Schoenberg [94] demonstrated that the reflection coefficient of a thin layer
was given by;
(z1 z2 ) + i/K(z1 z2 )
R= (D.3)
(z1 + z2 ) + i/K(z1 z2 )
where is the angular frequency of the ultrasonic wave (2f ) and K is the
stiffness of the interfacial layer. If the layer consists of a liquid, the stiffness
D.3. ULTRASONIC OIL FILM MEASUREMENT 167

of the layer depends on its bulk modulus and thickness (K = B/h). The
bulk modulus can be written in terms of the speed of sound, c, and density,
, of the layer material (B = c2 ). This gives,
c2
K= (D.4)
h
This stiffness can be used in the quasi-static spring model for identical
materials either side of the interface (z1 = z2 ). Eq. (D.3) becomes;
s
c2 |R|2
h= (D.5)
f z 1 |R|2

where |R| is the modulus of the reflection coefficient. This relationship gives
the layer thickness in terms of reflection coefficient and acoustic properties
of the oil and materials either side of the interface. In this case, the acoustic
impedance of cast iron piston ring and liner is 34.9 MRayls and speed of
sound in base oil can be found in Table D.2. More detail about ultrasound
film thickness measurements can be found in reference [55].
Experimentally, as the sensor is coupled to a test specimen, some of the
incident wave is transmitted forward and the remainder is reflected back.
The reflection coefficient can be obtained by,
Ar
R= (D.6)
Ai
where Ai is the incident wave amplitude and Ar is the reflected wave am-
plitude. However, it is difficult to measure the incident pulse. Hence in
practice it is convenient to record a reflection from the liner-air interface,
called the reference interface, because most of the acoustic energy emitted
from the sensor is reflected back due to a high acoustic mismatch between
the materials. Therefore the reflected wave is almost equal to the incident
wave. Eq. (D.6) shows that for such an interface, R tends to one. This
ultrasonic approach has been used previously to monitor the condition of a
lubricant film in machine elements [95, 96, 97]. The ultrasonic technique is
based on physical principles of the system. The experiments to determine
the validity of the ultrasonic film thickness measurement have been carried
out and excellent agreement was found for the quasi-static spring model
technique [98].

D.3.2 Data Capturing and Analysis


The ultrasonic reflections captured by the sensors were continuously recorded
as the test was in progress. In this section, only the data captured by sen-
sor 1 has been used to explain the data processing technique. The same
168 PAPER D.

procedure has also been applied to the other sensors to obtain the oil film
thickness. Fig. D.6a shows reflections from the liner specimen in the time
domain. The first pulse (I) is a combination of reflection from the back side
of liner and sensor initiation. The second pulse (II) is reflected from the
inner side of liner. This II pulse is isolated from the rest of the signal and
successively recorded during the test. Fig. D.6b shows a sample of these
successive signals as the piston ring reciprocates over the liner. It is seen in
Fig. D.6c that the amplitudes of the signals decrease while the piston ring
is passing over the sensing area. This is because some of the ultrasound
energy passes through the oil and is absorbed by the ring.
As mentioned previously, the reference pulse from the liner-air inter-
face was recorded before starting the test. A fast Fourier Transform (FFT)
was performed on the reference pulse and each of these successive reflected
pulses to extract the useful information from the signal (see Fig. D.6d). Thus
dividing all FFT amplitudes of the successive pulses by the amplitude of ref-
erence pulse gives the reflection coefficient spectra (i.e. R versus f ). Within
the -6dB bandwidth of the sensor, the reflection coefficient is smooth and
monotonically increasing with frequency. Despite the fact that frequency
appears in Eq. (D.5), its effect on h is cancelled by the counter-variation of
reflection coefficient with frequency. The resulting film thickness within the
bandwidth region is independent of measuring frequency [55, 97]. Therefore,
in this study, the reflection coefficient spectra within the bandwidth region
was used to determine the film thickness. Fig. D.6e shows the reflection co-
efficient against pulse number. In the figure, each trough, where the pulses
are reflected from the piston ring, corresponds to one of the ring traversals
over the sensor area. It is seen that the piston ring has passed over the
sensing area six times and there are two repetition intervals between the
troughs, short and long, since sensor 1 is close to TDC. The lubricant film
thickness can be obtained by substituting the reflection coefficient data into
the spring model, Eq. (D.5). Fig. D.6f shows the oil film thicknesses which
were obtained from the data given in Fig. D.6e. One ring passage has been
represented by approximately 100 pulses and this number depends on the
pulsing rate of the ultrasonic system and the ring reciprocating speed.
Fig. D.7 shows an oil film trace for one ring passage over the sensing
area. The width of the transducer (1.3 mm) was smaller than the ring
width (3 mm). Thus the sensor records an average of the reflection signal
over that 1.3 mm window, hence the film thickness is averaged over 1.3 mm
as the ring traverses the region. The exact ring profile is, therefore, not ex-
pected from this kind of profile measurement. A de-convolution algorithm
was used to eliminate this average effect on the minimum film thickness.
If the ultrasound wave is thought of as a series of discrete pulses over the
sensor area, then their reflection depends on the ring profile in the sensor
D.3. ULTRASONIC OIL FILM MEASUREMENT 169

2000 1500
1000

Amplitude (mV)
1500 (I)
Amplitude (mV)

1000 500
(II)
500 0
0 -500
-500 -1000
-1000 -1500
-1500 -2000
0 1000 3000 5000 7000 9000 0 5 10 15
Time (ns) Data Points 104
(a) (b)
22
1500 Ref 20 Ref
(i)
Amplitude (mV)

1000 18 (i)
(ii) 16 (ii)
Amplitude

500 (iii) 14 (iii)


0 12
-500 10
(i) (ii)(iii) 8
-1000 6
-1500 4
2
-2000 0
0 100 200 300 400 500 600 0 2 4 6 8 10 12 14 16
Time (ns) Frequency (MHz)
(c) (d)
Oil Film Thickness (m)

4
Reflection Coefficient

1 3.5
0.8 3
2.5
0.6 2
0.4 1.5
1
0.2
0.5
0 0
0 500 1000 1500 2000 2500 0 500 1000 1500 2000 2500
Pulse Number Pulse Number
(e) (f)

Figure D.6: Data analysis graphs for Sensor 1: (a) a typical waveform
showing the reflections from the liner specimen, (b) extracted pulses from
the inner side of the liner are amplified and successively recorded as the
piston ring reciprocates, (c) selected time domain signals, (d) FFTs of the
selected signals, (e) reflection coefficient, (f) oil film thickness.

window. The algorithm uses the spring model for each discrete reflection
from the ring profile in the sensor window and calculates the equivalent
reflection coefficient as the ring rests on the liner. This could be thought
of as the minimum reflection coefficient observed by the sensor if the mini-
mum film thickness is zero. Therefore, if the separation of ring and liner is
increased and this algorithm is repeated iteratively, the minimum reflection
170 PAPER D.

4
3.5

Oil Film Thickness (m)


3
2.5
2
1.5
1
0.5
0
-0.5 0 0.5 1 1.5 2 2.5 3 3.5
Distance from TDC (mm)

Figure D.7: Film thickness recorded for one ring passage.

coefficient observed by the sensor as a function of the separation of the ring


and liner is found. The true minimum film thickness can then be deduced
from the observed reflection coefficient. There was -/+2 % error present
in the reflection data due to electrical noise in the system. For a reflection
coefficient of 0.7, this yields -/+4% error in film thickness measurement.

D.4 Numerical Model

A numerical model of the experiment has been developed in order to predict


both the film thickness and friction that are also measured experimentally.
As it is assumed that the piston ring-cylinder liner contact runs in the mixed
lubrication regime for at least some of the stroke, a model must be developed
that calculates both the hydrodynamic film pressure and asperity contact
pressure.
As the contact profile is converging-diverging, some cavitation will occur
along the trailing edge of the ring. In order to solve the Reynolds equation
incorporating cavitation, a modified version of the Giacopini et al. [35] mass-
conserving cavitation algorithm was used. More precisely, a two dimensional
time dependent solution of an averaged form of the Reynolds equation was
D.4. NUMERICAL MODEL 171

restated as a Linear Complimentary Problem (LCP).


(A0 p0 ) (B0 ) h
t

+ (rB0 ) + r h = 0,
t
(D.7)
p0 > 0,
r > 0,
p0 r = 0.

In this form, the film thickness is replaced with flow factors, A0 and B0
are calculated according to the method found in [79], p0 is the averaged
film pressure, r is the complementary variable and and are constants
is
(defined in the nomenclature). The average film thickness parameter, h,
defined as;
= h0 (x, y, t) + 1
Z
h hr (x, y)dy, (D.8)
l1 l2
where h0 is the separation between the piston ring and cylinder liner, hr is
the liner roughness (see Fig. D.2) and l1 and l2 are the length and width of
the roughness measurement. The boundary conditions were defined as p0 =
0 and r = 1 at the inlet, outlet and sides of the contact which corresponds
to fully flooded and zero pressure.
The LCP problem, Eq. (D.7), was discretised using the finite difference
method, with central differences in space for all terms except for (rB0 )
which was upwind differenced, to properly consider the hyperbolic nature
of the problem inside cavitation zones. An explicit method built on the
forward euler method was used to discretise the problem in time. The
solution domain, Eq. (D.7), was divided into 50 by 10 nodes (50 in the
entraining direction, 10 across the width), which was found to make the
film thickness and friction virtually independent of grid size. Such a course
grid representation is made possible by incorporating the effect of surface
roughness in the flow factors A0 , B0 and h. These are calculated using the
technique described in [79] over the liner surface illustrated in Fig. D.2, the
equations solved and the calculated values are given in the appendix. The
problem was divided into 100 time steps. Increasing the number of time
steps was found to only marginally affect the solution.
Based on the assumption that the piston ring surface is much smoother
than the cylinder liner, the piston ring was modelled as perfectly smooth.
When the surfaces come into contact, the deformation, asperity contact
pressure and real area of contact is found using a boussinesq-type elasto-
plastic contact mechanics model. The technique is described in detail by
172 PAPER D.

Sahlin et al. [29, 30] and will not be repeated here. Asperity contact pressure
as a function of separation is given in the appendix, Fig. D.18.
Each time step is associated to a specific plint spindle angle () and the
velocity of the piston ring was calculated from Eq. (D.9):

U = N s cos () , (D.9)

where N is the rotational speed of the Plint machine in Hz and s is the


stroke. Once the velocity is known a force balance equation is solved for the
film thickness which, in turn, balances the applied load with the hydrody-
namically supported load and the load supported by asperity contact (from
the contact mechanics model [29, 30], Fig. D.18). Once this is complete the
problem can be incremented one time step and the process repeated. As
the explicit method is applied to solve Eq. (D.7) numerically, the solution
at the current time step depends on the previous one and the solver must
be run through approximately 1.1 full cycles for the transients to fade out
and to reach convergence with the previous cycle. Convergence is assumed
to be reached when the film thickness and derivative of film thickness is
within 1% of the previous cycle.
The friction force (ftot ) is calculated as the sum of viscous friction force
fhyd and boundary friction force fbd . Boundary friction is calculated as;
Z
fbd = PCP dA (D.10)

where is the dry friction coefficient and PCP is the average asperity con-
tact pressure, found from the contact mechanics model. The dry friction
coefficient, taken as 0.192, was found by running a reciprocating test, in the
test apparatus described in Section D.2, with no lubricant present. Hydro-
dynamic friction is calculated as;

 
1
Z
Fhyd = U + 6c11
h
   (D.11)
h p0 p0
+ d11 + d12 dx
2 dx dy
is the average separation and c11 , d11 and d12 are flow factors
where h
calculated as described in [29, 30].

D.5 Results and Discussion


In this section the film thickness measured with ultrasound will be presented
as well as comparisons of film thickness and friction with the numerical
model.
D.5. RESULTS AND DISCUSSION 173

D.5.1 Measured Film Thickness


If an array of the successive pulses of interest (i.e. pulse (II) in Fig. D.6a)
individually captured by the sensors is analyzed, the film thickness values
at the fixed sensor locations can be obtained. These film thickness values
can be superimposed on a single graph. Fig. D.8 shows the measurements of
film thickness (OFT) as the ring reciprocates at 2.5 Hz under a normal load
of 80 N. In the figure, the ring starts its travel from TDC to BDC, thus it is
initially captured by sensor 1 (close to TDC). As the ring moves from one

2
1.8
S1
1.6
1.4 S2
OFT (m)

1.2
S3
1
S4
0.8
0.6 S5
0.4
BDC TDC BDC TDC
0.2
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
Pulse Number

Figure D.8: Oil film thicknesses measured by sensors as the piston ring
reciprocates over the liner specimen.

sensing area to the next, the other sensors detect the ring respectively. This
roughly provides an overview of lubricant film formation over the stroke. It
is notable that the film thickness data from cycle to cycle was very repeat-
able. The horizontal axis in the figure is given in terms of pulse number,
however this could be converted to time if the pulse rate (indicating how
many ultrasound pulses are sent through the liner specimen in a second) is
known. In this case the pulse rate is 2000 pulses per second. This was also
confirmed because it can be seen in the figure that two and half cycles were
observed in one second at 2.5 Hz (2.5 rev/sec).
If the oil is partly depleted under the sensing area, this leads to more
ultrasound reflected from the interface due to an air-oil mixture in the con-
tact. This results in bigger reflection coefficients and the impression of
greater film thickness measurements being recorded than is actually the
174 PAPER D.

case. This is more visible for the starved condition where a smaller amount
of oil into the contact is provided (see Fig. D.9a). Under the same loading
and reciprocating speed, the reflection coefficients are bigger in the starved
condition than in the lubricated condition, Fig. D.9b. During testing, there

Starved Contact
1

S1
Reflection Coef.

0.95 S2

S3

S4
0.9
S5
Down-stroke Up-stroke
0.85
0 0.2 0.4
Time (second)

(a)
Lubricated Contact
0.85
0.8 S1
Reflection Coef.

0.75 S2
0.7
S3
0.65
S4
0.6
S5
0.55 Down-stroke Up-stroke
0.5
0 0.2 0.4
Time (second)

(b)

Figure D.9: Measured reflection coefficients for one complete cycle: (a)
starved and (b) flooded contact conditions; higher reflection coefficients are
observed in starved conditions due to the presence of an air-oil mixture in
the contact.

were some fluctuations in the ultrasonically measured ring profiles, espe-


cially in the case of the ring travelling from BDC to TDC (see Fig. D.8).
This is because the interface between the ring and liner is not homoge-
D.5. RESULTS AND DISCUSSION 175

nously filled with lubricant due to cavitation occurring. The ring profile is
not symmetric, but has a different diverging shape in each direction which
influences the lubricant condition in the contact and how much cavitation
occurs. The ring was installed in the ring holder with the greater converging
shape facing towards BDC (see Fig. D.11). Because of this profile, normally
a higher film thickness would be expected on the down stroke from TDC
to BDC. However, the non-symmetric barrel shaped piston ring leads to a
different size of the cavitation region according to the direction of the stroke
[32]. The cavitation region on the down stroke is considerably smaller than
that on the up stroke because the diverging part of the ring is much smaller
in down stroke than in the up stroke.
This explanation is illustrated further with the results of the numerical
simulations, see Fig. D.10a and Fig. D.10b. It can be seen that the cavitated
region is much smaller on the down stroke as opposed to the up stroke. Since
there is a far larger cavitating region on the up stroke, as is mentioned the
ultrasound recorded an anomalously high film thickness in this direction
due to an inhomogeneous film layer (i.e. air-oil mixture). This can be seen
in Fig. D.11 where oil film thickness is bigger on the up stroke, compared
with the down stroke. However, fluctuations in the measured signal are
clearly discernible. In this study, the down stroke data exhibiting far less
fluctuations has been used for comparison with the numerical model. It
was concluded that the up stroke data cannot be considered reliable due to
excessive cavitation occurring.
An example of the minimum oil film thicknesses (MOFT) measured by
the sensors for only down strokes is given in Fig. D.12 where the MOFT
data represents the mean value of a series of cycles and the standard de-
viation gives an indication of cycle to cycle variation. Fig. D.13 shows the
effect of speed and load on film thickness formation in the ring-liner contact.
It is seen that the minimum film thickness decreases as the normal load is
increased. This is more clear as the reciprocating speed increases. Increas-
ing the ring sliding speed enhances the wedge effect of the converging ring
profile, providing better load carrying capacity and a thicker lubricant film.
The hydrodynamic effect is more obvious for the tests performed at a low
load (60 N). It is also possible to see this hydrodynamic effect for each data
set as the measured minimum film is thicker at S3 where the sliding speed
is greatest.

D.5.2 Comparisons between Experimental and Numerical


results
Several short duration tests were carried out at different speed and load
conditions. The results from the numerical model will be compared with
PSfrag

176 PAPER D.

12

10

Film Pressure (MPa) 8

4
Cavitated
2 region
Full film region
0
0 0.5 1 1.5 2 2.5 3
Width (mm)
(a)

15
Film Pressure (MPa)

10

Cavitated Region Full film region


0
0 0.5 1 1.5 2 2.5 3
Width (mm)
(b)

Figure D.10: Results from the numerical simulations showing the full film
and cavitated regions for (a) the Down stroke and (b) the Up stroke.

the values measured experimentally, firstly in terms of friction and then oil
film thickness.

Friction

In this section, the total friction force calculated from the simulations, a
combination of both boundary and hydrodynamic friction, is compared with
the friction force measured by the force transducer in the experiment. Cross-
correlation was performed on the raw data collected by the force transducer
to identify each friction cycle during the test period. Following this the
mean friction cycle was calculated from all of the cycles collected over the
D.5. RESULTS AND DISCUSSION 177

B T B T
D D D D
C C C C

2
1.8 Down-stroke Up-stroke
1.6
1.4
OFT (m)

1.2
1 Fluctuations
0.8
0.6
0.4
0.2
0
Time

Figure D.11: A close view of ultrasonically measured ring profiles; fluctua-


tions in the up stroke.

0.7
Average

MIN OFT
0.6
MIN OFT (m)

0.5

0.4

S1 S2 S3 S4 S5
Sensors

Figure D.12: Minimum film thickness for down stroke (Normal load: 80 N,
reciprocating speed: 2.5 Hz).

test duration. Fig. D.14 gives a comparison of the experimental and sim-
ulated friction for two different operating conditions, both with the same
speed of 15 Hz but with a low load of 40 N and a high load of 100 N.
Fig. D.15 again compares experiment and simulation, but this time with a
constant load of 60 N and a low speed of 10 Hz and a high speed of 17.5
Hz. Analysing the results, it is noticeable that there is a large difference
in friction values around TDC, or 0 plint spindle angle. In the numerical
178 PAPER D.

180N 120N 60N


1.6
Operating speed
1.4 increasing
MIN OFT (m)
1.2
1
0.8
0.6
0.4
0.2
S1 S3 S5 S1 S3 S5 S1 S3 S5 S1 S3 S5
Ultrasonic Sensors
Figure D.13: The effect of load and speed on the minimum film thickness
on the down stroke.

20
Exp., 40N,15Hz
15
Cal., 40N,15Hz
10 Exp., 100N,15Hz
Friction (N)

5 Cal., 100N, 15Hz

0
-5
-10
-15
-20
0 50 100 150 200 250 300 350
Crank Angle (deg)

Figure D.14: Comparison of calculated and experimentally measured fric-


tion for a constant speed of 15 Hz and loads of 40 N and 100 N.

simulation, boundary lubrication is predominant until approximately 35


plint spindle angle when the sliding speed becomes sufficiently high to allow
for the surfaces to become fully separated by hydrodynamic effects. At this
point the boundary friction contribution dramatically reduces and only the
viscous friction component is left, giving the sudden drop in friction force.
However, in the experiment the friction force drops far more quickly at ap-
proximately 15 plint spindle angle. It is proposed that this is due to the
dynamics of the test rig, leading to stick-slip occurring. When the sliding
speed approaches zero at TDC, the ring comes into asperity contact with
D.5. RESULTS AND DISCUSSION 179

15
Exp., 60N, 10Hz
10 Cal., 60N, 10Hz
Exp., 60N, 17.5Hz
5 Cal., 60N, 17.5Hz
Friction (N)

-5

-10

-15
0 50 100 150 200 250 300 350
Crank Angle (deg)

Figure D.15: Comparison of calculated and experimentally measured fric-


tion for a constant load of 60 N and speeds of 10 Hz and 17.5 Hz.

the liner and friction increases dramatically. As the ring sliding speed in-
creases again, the ring sticks and then suddenly slips due to the difference
in static and dynamic coefficients of friction. This can be evidenced by the
rippling that occurs as the friction suddenly drops.
It should be noted that the plint spindle angle here does not correlate
directly to the crank angle in a real engine with respect to the lubrication
regime. The stroke and therefore the acceleration is much greater in a real
engine and it is suggested that this will lead to full film lubrication occurring
in a real engine at a much earlier crank angle compared to the plint spindle
angle discussed here.
During the midstroke the lubrication is predominantly hydrodynamic
and there is a much closer match between simulation and experiment. Any
difference here is most likely attributed to an inaccurate value of lubricant
viscosity being used in the simulation, which is an estimate at 22 C based
on the provided lubricant data.
At BDC, or 180 plint spindle angle, as the friction increases again there
is less difference between experiment and simulation as was present at TDC,
with the values matching well. This is probably because, due to the asym-
metric ring profile, there is a larger film present on the downstroke and so
at BDC the film thickness is greater (due to the squeeze effect). Therefore,
the issues encountered at TDC with the very low film thickness are less
prevalent.
180 PAPER D.

Film thickness
In this section, the minimum oil film thickness calculated with the numeri-
cal model is compared to the values measured at the five sensor locations.
Fig. D.16 illustrates this data at a constant speed of 10 Hz with a low,
medium and high load. From Fig. D.16 it can be seen that the experimen-

2
1.8
Cal., 40N,10Hz
1.6
Cal., 120N,10Hz
MIN OFT (m)

1.4
1.2 Cal., 200N,10Hz

1 Exp., 40N,10Hz
0.8 Exp., 120N,10Hz
0.6 Exp., 200N,10Hz
0.4
0.2
0 20 40 60 80 100 120 140 160 180
Crank Angle (deg)
Figure D.16: Comparison of calculated and experimentally measured film
thickness at a constant speed of 10 Hz and loads of 40 N, 120 N and 200 N.

tal data correlates reasonably well with simulation. Ultrasonically measured


MOFT varied between 0.25 and 1.8 microns. The deviation bars of the min-
imum film thickness were less than 0.05 microns for high load but tended to
increase with low load. In general, the numerical simulation overestimates
the minimum oil film thickness when compared to the ones acquired with
the ultrasound technique. It is suggested that the predominant reason for
this is due to the value of viscosity used in the simulation, as was mentioned
previously when discussing the friction results. It is unknown exactly what
the effective viscosity was in the contact. For example, it is inevitable that
some shear heating of the lubricant will take place during sliding and this
increase in temperature will reduce the viscosity of the lubricant. A lower
viscosity would cause lower film thicknesses, as has been shown numerous
times experimentally.
Another source of error is the assumption and modelling of the contact
geometry. Although the ring shape was measured accurately (see Fig. D.3)
and the ring adapter was aligned to make sure it sat flat and parallel with
the liner sample, it is still possible that there was a small misalignment.
Although small, such misalignment leads to a slightly different contact ge-
ometry in turn, causing a slightly different film thickness. This is not consid-
ered in the numerical model which assumes perfect alignment and contact
D.6. CONCLUSION 181

geometry accordingly.
Also, it must be remembered that the ultrasound technique requires as
input a value of the density and speed of sound in the lubricant as found
inside the contact. The values used in the calculation of the oil fim thickness
are those corresponding to the lubricant bulk properties. This is another
potential source of error.

D.6 Conclusion
The piston ring-cylinder liner contact has been simulated in a reciprocating
test rig apparatus measuring the oil film thickness between a section of a
piston ring and a cylinder liner using an ultrasound technique. Moreover,
a numerical method was developed to enable predictions of the oil film
thickness values obtained with the ultrasound technique. The following
conclusions are drawn:

1. Ultrasound proves an effective way of measuring film thickness, a use-


ful aid for understanding the lubrication process in a tribological con-
tact and for validating simulations.

2. Cavitation occurs in the contact and this can adversely effect the
ultrasonic measurement of film thickness, however this is detectable
in the reflection coefficient profile and in this case it only significantly
affected the tests in one sliding direction.

3. The numerical prediction of oil film thickness and friction force matched
well with the experimental data, however the result is sensitive to the
value of viscosity used and more work is required to deduce means of
acquiring viscosity data corresponding to the values found inside the
lubricated conjunction.

D.7 Acknowledgements
Scania AB is thanked for providing components, technical data and discus-
sions. The authors from the Leonardo Centre would like to acknowledge the
EPSRC Encyclopaedic Program Grant for funding this work. Support of all
the program industrial partners is acknowledged. The authors from LTU
would like to thank Stiftelsen for Strategisk Forskning (SSF) and ProViking
for funding their contribution to this work and the Swedish Research School
in Tribology for funding Andrew Spencers placement at Sheffield Univer-
sity.
182 PAPER D.

D.8 Appendix

The calculation of the flow factors A0 and B0 is described below. A0 and


B0 can be written as:

   
a11 a12 b1
A0 = and B0 = (D.12)
a21 a22 b2

To calculate the individual flow factors, three partial differential equations


with periodic boundary conditions must be solved over the measurement
domain, i.e., the cell of periodicity, Y = (0, l1 ) (0, l2 );

y h3 y 1

= 0 in Y, (D.13a)

1 (x, 0, y2 ) + 1 = 1 (x, l1 , y2 ) ,
1 (x, y1 , 0) = 1 (x, y1 , l2 ) .

y h3 y 2

= 0 in Y, (D.13b)

2 (x, 0, y2 ) = 2 (x, l1 , y2 ) ,
2 (x, y1 , 0) + 1 = 2 (x, y1 , l2 ) .

y h3 y 0

= y (he1 ) in Y, (D.13c)

0 (x, 0, y2 ) = 0 (x, l1 , y2 ) ,
0 (x, y1 , 0) = 0 (x, y1 , l2 ) .

Where h describes the clearance between the rough surfaces and 1 , 2 and
0 are local scale variables. Once 1 , 2 and 0 have been calculated they
can be integrated over to give the flow factors in Eq. (D.12) needed to solve
D.8. APPENDIX 183

the homogenized Reynolds equation;


1 1
Z
a11 (x) = h3 dy, (D.14a)
l2 Y y1
1 2
Z
a12 (x) = h3 dy, (D.14b)
l1 Y y1
1 1
Z
a21 (x) = h3 dy, (D.14c)
l2 Y y2
1 2
Z
a22 (x) = h3 dy, (D.14d)
l1 Y y2
 
1
Z
b1 (x)
= he1 h3 y 0 dy (D.14e)
b2 (x) l1 l2 Y

These flow factors were calculated for the roughness measurement shown in
Fig. D.2 and the results for a11 , a22 and b1 , the flow factors of most interest,
are presented in Fig. D.17 as a function of separation. The asperity contact

107
10
9 1/3
a11
8 1/3
a22
7 b1
Flow Factor

6
5
4
3
2
1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Separation (m)

Figure D.17: Flow factors calculated for the cylinder liner surface.

pressure, PCP , calculated using the boussinesq-type elasto-plastic contact


mechanics model as a function of separation for the cylinder liner surface
shown in Fig. D.2 is presented in Fig. D.18.
184 PAPER D.

70
Mean Asperity Contact Pressure (MPa)

60

50

40

30

20

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Separation (m)

Figure D.18: Contact stiffness calculated for cylinder liner surface.


Paper E

An experimental and
numerical investigation of
frictional losses and film
thickness for four cylinder
liner variants for a heavy
duty diesel engine

185
187

Proceedings of the Institution of Mechanical Engineers, Part J:


Journal of Engineering Tribology, December 2013, vol. 227 no. 12, pp.
1319-1333

An experimental and numerical investigation of frictional


losses and film thickness for four cylinder liner variants
for a heavy duty diesel engine
A. Spencera1 , E. Y. Avanb , A. Almqvista , R. S. Dwyer-Joyceb , R.
Larssona
a
Division of Machine Elements, Department of Engineering Sciences and
Mathematics, Lule
a University of Technology, Lule a SE-971 87 Sweden
b
The Leonardo Centre for Tribology, Department of Mechanical Engineering,
University of Sheffield, Sheffield, UK

Abstract
The piston ring pack is the single greatest contributor to mechanical losses
in a Heavy Duty Diesel Engine (HDDE), accounting for 1.1-6.8% of the to-
tal losses. Therefore, the piston ring-cylinder liner contact is potentially
the most rewarding area to study when attempting to reduce mechani-
cal losses in a HDDE. In this work, four different HDDE cylinder liner
variants have been tested to evaluate the lubricating conditions that occur
when a section of top compression ring is reciprocated against them in a
lubricated environment. Two of the cylinder liners were traditional grey
cast iron and plateau honed with different honing angles, one had ANS

Triboconditioning applied and the last was plasma sprayed with a stain-
less steel and ceramic coating, then honed. An experimental test rig was
used where friction and film thickness was recorded, by means of an ultra-
sonic technique. A numerical model was also developed to calculate the
friction and film thickness. Comparisons are made between the simulation
and experiment, and the four cylinder liner variants are also evaluated. It
was found that both simulation and experiment could differentiate between
all surfaces and the results from the model and experiment also correlated
well with each other. A lower plateau average surface roughness, as exhib-

ited by the ANS Triboconditioning and plasma liners, led to a significant
reduction in friction.

1
Corresponding author.
E-mail address: andrew.spencer@ltu.se
188 PAPER E.

Keywords: cylinder liner, ultrasound, tungsten disulphide, piston ring, honing


E.1. INTRODUCTION 189

E.1 Introduction
Fuel efficiency is one of the most important areas of automotive vehicle
research and development today, with rising fuel costs, energy security and
environmental concerns being at the forefront of customers and legislators
minds. Heavy Duty Diesel Engines (HDDE) are the primary source of
mechanical power generation in todays trucks and buses and this is likely to
continue for the foreseeable future. In 2010 the global demand for diesel fuel
was 16 million barrels of oil per day, in 2040 this is predicted to be 26 million
barrels per day, an increase of 60% [1]. EU transport still depends on oil
and oil products for 96% of its energy needs [2]. Although larger trucks and
truck efficiency will increase, growing worldwide GDP will increase the need
for road transportation and road congestion will increase fuel use further. In
the 2011 European Commission White Paper on transport [2], a reduction
of at least 60% of greenhouse gas emissions from transport by 2050, with
respect to 1990 levels, was called for. The report concludes that acting on
vehicles efficiency through new engines, materials and design will help in
the reduction of oil dependence, the competitiveness of Europes automotive
industry as well as health benefits, especially improved air quality in cities.
However, it is predicted that a considerable extent of the shipments over
short and medium distances, i.e. below 300 km, and that 50% of shipments
over 300 km will remain on trucks even in 2050. Also, there will be a greater
push in cities to encourage the use of public transport with large fleets of
urban buses. Therefore, the efficiency and frictional losses in a vehicles
powertrain, particularly those in trucks and buses powered by a HDDE, are
areas of great interest.
Only about 40% of the fuel energy consumed in a HDDE is converted
to mechanical power [3]. The rest is lost to a combination of thermal and
mechanical losses. Although the thermal losses are the greatest component,
the mechanical friction losses are significant at between 4-15% of the total
fuel energy [3]. Of these mechanical losses, the piston ring pack is the
greatest single contributor, amounting to 1.1-6.8% of the total losses in a
HDDE [3]. Hence the ring pack is possibly the most interesting area for
consideration when trying to reduce mechanical friction in an engine.
Cylinder liner surface topography can be an important factor in the
amount of piston ring-cylinder liner (PRCL) friction generated. Various
authors have investigated the effect of surface roughness and topography
on PRCL lubrication [27, 59]. Johansson et al. [99] showed through sim-
ulation that cylinder liner surfaces with a lower core roughness parameter,
Rk , exhibit both lower film thickness and friction. In fired engine tests with
a floating liner, Sato et al. [100] showed that a smoother surface reduced
PRCL friction. However, with smoother surfaces more vertical flaws, or
190 PAPER E.

scuffing, became apparent on the liner and this led to increased oil con-
sumption.
In this work, four distinct cylinder liners are investigated, each with dif-
ferent surface roughness and one with a low friction surface coating. The
samples are investigated with both numerical simulations and in an exper-
imental test rig, to aid in validating the numerical solution. Both friction
and film thickness are simulated and measured. The friction is the primary
parameter of interest, however the film thickness aids in understanding the
friction results and allows a more in-depth comparison of the numerical and
experimental solutions.
This work has two objectives: firstly to see whether advanced surface
topographies, with lower roughness and surface coatings, can reduce friction
over more traditional topographies, and secondly to see whether numerical
simulations can accurately predict the friction between the different sur-
faces. If possible, the numerical model could then be used to optimise
the surface roughness and investigate possibilities of even lower friction. It
is the authors belief that this is the first time that both experimentally
and numerically obtained film thickness has been used to evaluate different
cylinder liner surface topography.

E.2 Experimental setup


The experimental setup and ultrasound measurement technique is the same
as that used previously by the authors [101]. This work is referenced for
a detailed description of the setup of the test rig and processing of the
ultrasound data, however a more brief description of the setup will also be
given here.

E.2.1 Test setup, specimens and lubricant


A Plint TE-77 high frequency reciprocating tribometer, schematically shown
in Fig. E.1, was used in this work. The machine configuration consists of
sections of piston ring and liner where the ring specimen reciprocates over
the stationary liner sample. The normal load is applied via a spring-balance
through a lever and stirrup mechanism. The normal force is transmitted
directly onto the ring section by means of a needle roller cam follower on a
carrier head and a running plate on the loading stirrup. The whole assembly
is mounted on flexible supports providing for free movement in horizontal
directions, and connected to a stiff force transducer (Kistler type 9203) that
measures tangential force in both directions.
A section of run-in cylinder liner and top compression ring from the same
heavy duty diesel engine were used to create the contact. Four different
E.2. EXPERIMENTAL SETUP 191

Figure E.1: Plint TE-77 simulated piston ring-liner contact schematic.

cylinder liner samples were used with different surface topographies and
material properties, these are discussed later in section E.4. Each cylinder
liner sample measured 50 mm in length and 20 mm in width and was cut
from a complete cylinder liner with a bore diameter of 130 mm. The piston
ring was sectioned into a length of 45 mm. The width of the piston ring is 3
mm and it has an asymmetric barrel shaped face, as illustrated in Fig. E.2,
with a chrome surface coating. A lubricant bath was modified to hold the
liner specimen. Six grub screws were used to secure the liner specimen in
place; this allows for alignment of the liner in both the axial and lateral
directions. To retain the ring section, a special ring holder attached to the
carried head was manufactured from an original production piston. A ring
section was clamped to the ring holder using two slotted plates either side
of the ring holder and a grub screw in the centre. This clamping system
bent the ring section and allowed it to conform over the liner section. The
conformability between the ring and liner was checked using pressure paper
and a good conformal contact was obtained [101].
192 PAPER E.

45
40
35

Ring height (m) 30


25
20
15
10
5
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Ring width (mm)

Figure E.2: Top Compression Ring Profile. When fitted in an engine the
right hand side faces towards the combustion chamber. In these tests, the
right hand side faces towards the reciprocating arm.

E.2.2 Test conditions


For each of the four liner surfaces, tests were run for a range of speeds and
loads. The load was varied from 40 N to 200 N in steps of 20 N and the
speed was varied from 2.5 Hz to 17.5 Hz in steps of 2.5 Hz, giving 63 test
points in total (7 speeds and 9 loads).
At this point it is of interest to make a comparison with the real engine
operating conditions. The maximum piston speed at 1200 RPM is 10.5 m/s.
At a peak combustion pressure of 200 bar, if this pressure is assumed to act
entirely on the back of the top compression ring, the load on a section of
ring of the size used in these tests (20 mm x 3mm) would be 1200N. This
is the maximum speed and load that the top compression ring is subjected
to in the engine. Therefore, the speed in these tests is representative of the
reversal points, but not mid-stroke. Also, the load is representative of most
of the engine cycle, but not around the region of peak combustion pressure.
It would of course be of interest to run at higher speeds and loads, but the
conditions used here were the upper limits of the capabilities of the test rig
used in this study.
The stroke of the machine was set to 15 mm (the maximum value). The
liner specimen was fully immersed in pure base oil without an additive pack-
age. The oil bath temperature was logged at a stable 22 C throughout the
tests. It is acknowledged that the lubricant temperature is unrepresentative
of real engine running conditions. However, the primary goal of this work is
E.2. EXPERIMENTAL SETUP 193

to accurately and repeatably investigate different liner surface topographies


rather than recreate engine running conditions. It was decided that running
at a lower temperature, giving a higher viscosity and therefore higher film
thickness would go some way towards compensating for the lower entraining
speeds in the test rig compared to the real engine operating conditions. In
addition, the large quantity of lubricant in the oil bath ensures that the
inlet is always fully flooded allowing for good, accurate comparisons with
the numerical model. It also serves in maintaining a stable temperature of
the liner surface during the short tests.

E.2.3 Ultrasonic film thickness measurement


For over a decade, several methods have been developed and used to mea-
sure the thin oil film that forms under piston rings such as capacitance
[87, 88, 89], resistance [102] and the laser induced fluorescence method
[91, 92, 103]. However, these are all invasive methods meaning that they
require penetration of the cylinder liner in order to measure the piston ring
to cylinder liner film thickness. In this study, a non-invasive ultrasound
technique based on sensing the reflections from the piston ring-liner contact
was used. This technique has been recently applied to a hydraulic motor
piston ring-liner [104], a motored engine [56] and a fired engine [105] to
quantify the oil film thickness between piston ring and cylinder liner.
To generate ultrasonic waves, 10 MHz piezoelectric crystals with a width
of 1.3 mm and a length of 2.5 mm were placed on the back side of each liner
segment. Five such ultrasonic sensors were equally distributed along the
axial length of the stroke, within millimetre precision. The positions of the
sensors on the each liner sample are given in Table E.1. When these piezo

Table E.1: Sensors location over the stroke for the liner samples.

Distance from TDC (mm)


Liner Sample STD55 STD35 ANS PL
TDC 0.00 0 0 0
Sensor Number

1 1.4 1.4 1.4 1.6


2 4.1 4.3 4.2 4.7
3 7.3 7.4 7.1 7.4
4 10.4 10.2 9.7 10.0
5 13.3 13.2 13.2 13.0
BDC 15.0 15.0 15.0 15.0

elements are pulsed at high frequencies they generate and send ultrasonic
waves through the inner liner surface. When the piston ring is within the
sensing area, some of the incident wave is transmitted forward to the ring
194 PAPER E.

and the remainder is reflected back. The proportion that is reflected is


known as the reflection coefficient, R, and varies with the acoustic properties
of the materials and the stiffness of the layer, K. The response of a thin
layer embedded between two materials is governed by a quasi-static spring
model [7],
1
R= p (E.1)
1 + (2K/z)
where is the angular frequency of the ultrasound and z the acoustic
impedance of the materials. For the case of liquid layer trapped between
two flat surfaces, the stiffness of the layer can be related to its bulk modulus,
B, and thickness, h, by K = B/h. Furthermore, the bulk modulus can be
replaced using the relation, B = c2 , where and c are the speed of sound
and density of the liquid layer (i.e. lubricant) respectively. This gives,

c2
K= (E.2)
h
If Eq. (E.2) is substituted into Eq. (E.1) and rearranged, this gives Eq. (E.3)
where the layer thickness can be described in terms of reflection coefficient
and acoustic properties of the oil and materials either side of the interface;
s
c2 |R|2
h= , (E.3)
f z 1 |R|2

where f is the frequency of the ultrasonic pulse. In this work the acoustic
impedance of the cast iron piston ring and liners, z, is 34.9 MRayl and speed
of sound in base oil can be found in Table E.2.
During the tests, the ultrasonic reflections were recorded as the piston
ring passed over the sensing area. The reflection coefficients were created by
normalizing measured reflections with a reference reflection and then the
spring model Eq. (E.3) was employed for each reflection coefficient to deter-
mine the oil film thickness. Practically, the air-liner interface was used as
a reference interface because an almost complete reflection of incident wave
occurs for this interface due to a high acoustic mismatch. Before testing,
the inner liner surface was therefore cleaned (no oil present on the surface)
and the reference signal for each individual sensor was recorded. More detail
about ultrasound film thickness measurement and signal processing can be
found in [55] and [101].
In the experiments, ultrasonic data from the sensors was simultaneously
streamed into an ultrasonic pulsing unit via 5 channels. The unit consists of
an ultrasonic pulsing and receiving card, a data acquisition card and digitis-
ing card. Each channel was pulsed at a repetition rate of 15k pulses/second.
The data was digitised at 100 MHz with a 12 bit resolution.
E.3. NUMERICAL MODEL 195

E.3 Numerical model


A numerical model of the experiment has been developed in order to predict
both the film thickness and friction that also are measured in the previously
described experimental setup. As it is assumed that the piston ring-cylinder
liner contact runs in the mixed lubrication regime for at least some of the
stroke, a model must be developed that calculates both the hydrodynamic
film pressure and asperity contact pressure.
The contact profile is first converging then diverging. This leads to that
cavitation will occur along the trailing edge of the ring. In order to solve
the Reynolds equation incorporating cavitation, a modified version of the
Giacopini et al. [35] mass-conserving cavitation algorithm was used. More
precisely, a two dimensional time dependent solution of an averaged form
of the Reynolds equation was restated as a Linear Complimentary Problem
(LCP).


(A0 p0 ) (B0 ) h
t

+ (rB0 ) + r h = 0,
t
(E.4)
p0 > 0,
r > 0,
p0 r = 0.

where p0 is the averaged film pressure, r is the complementary variable and


and are constants defined in the nomenclature. The advantage with the
approach proposed by Giacopini et al. [35] is that the two unknowns become
complementary throughout the whole domain. Existence and uniqueness
of the solution follows by the rigorous mathematical analysis in Bayada
et al. [106]. The advantage with the LCP formulation is that standard
techniques can be used to solve the problem numerically, e.g. Lemkes
pivoting algorithm. See e.g. the book by Cottle et al. [37]. This alleviates
the problematics associated with discrete formulations that changes at the
boundaries between the cavitated and the full film zones. Moreover, this
solution technique finds the solution in a finite number of steps, hence issues
related to iterative processes are avoided.
If r = 0 then the contact is fully flooded and a positive hydrodynamic
pressure exists in the lubricant. If r > 0, there is no hydrodynamic pressure
in the lubricant, physically the lubricant is cavitating here. r can be defined
as follows;

r =1 (E.5)
c
196 PAPER E.

where is the density of the lubricant and c is the density of the lubricant
at the cavitation pressure, which in this work is assumed to be atmospheric
pressure. With Eq. (E.4) in this form, the film thickness is replaced with
coefficients, A0 and B0 where,
   
a11 a12 b1
A0 = and B0 = (E.6)
a21 a22 b2
A0 and B0 are calculated for each of the surfaces under investigation. An
explanation of the method to calculate these coefficients is given later in
Section 4.1. The average film thickness parameter, h, is defined as;

= h0 (x, , t) + 1
Z Z
h hr (x, y)dy, (E.7)
Y l1 l2
where h0 is the global separation between the piston ring and cylinder liner,
hr is the liner roughness, l1 and l2 are the length and width of the rough-
ness measurement and Y is the domain of the local scale, which is further
discussed in section E.4. The boundary conditions were defined as p0 = 0
and r = 1 at the leading edge, trailing edge and sides of the contact which
corresponds to fully flooded and zero pressure. Zero pressure boundary con-
ditions were implemented as the ring is not subjected to gas pressures at
either edge of the ring, as would be the case in a real engine. These bound-
ary conditions are not representative of real engine operating conditions,
but are used in these tests as they are repeatable and known, as opposed
to a starved inlet where the degree of starvation would be hard to measure
experimentally.
The LCP problem, i.e. Eq. (E.4), was discretised using the finite dif-
ference method, with central differences in space for all terms except for
(rB0 ) which was upwind differenced, to properly consider the hyperbolic
nature of the problem inside cavitation zones. An explicit method built
on the forward Euler method was used to discretise the problem in time.
The solution domain, Eq. (E.4), was divided into 50 by 10 nodes (50 in the
entraining direction, 10 across the width), which was found to make the
film thickness and friction virtually independent of grid size. Such a coarse
grid representation is made possible by incorporating the effect of surface
roughness in the coefficients A0 , B0 and h. The problem was divided into
100 time steps. Increasing the number of time steps was found to only
marginally affect the solution.
At the reversal point the value of r at all grid points was set to 1. The
reason for this is as follows. As the ring slides a cavitated region is formed in
the diverging outlet section of the ring. As the ring changes sliding direction
at the reversal point, this cavitated region is now found in the inlet region.
The ring must then slide far enough for new oil to enter the contact before
E.3. NUMERICAL MODEL 197

a wedge of oil can be formed and hydrodynamic lift generated. This is


a real physical phenomenon, however in the Plint test rig there is a large
quantity of oil in the oil bath, unlike on the cylinder liner surface in a real
engine. It is in these authors opinion that at the reversal point, in this
lubrication situation, this oil rapidly replenishes into the diverging portion
of the ring and fills what becomes the inlet region of the ring at reversal.
There is no direct experimental evidence for this, however the rapid increase
in ultrasonically measured film thickness as the ring moves away from the
reversal point indicates that the inlet must be fully flooded with lubricant
immediately after reversal. The physics of this oil flow is not modelled
by the Reynolds equation. By setting r to 1 at the reversal points, the
contact is artificially fully replenished with oil immediately. This is thought
to properly simulate the mechanism described above.
Each time step is associated to a specific Plint spindle angle () and the
velocity of the piston ring was calculated from Eq. (E.8):
U = N s cos () , (E.8)
where N is the rotational speed of the Plint machine in Hz and s is the
stroke length, which in these tests was 15 mm. Once the velocity is known
a force balance equation is solved for the film thickness;
FHYD (h) + FCP (h) = FLOAD (E.9)
The hydrodynamically supported load, FHYD , comes from integrating the
hydrodynamic pressure found from Eq. (E.4). The asperity contact load,
FCP , is found from integrating the asperity contact pressure, PCP , from the
contact mechanics simulation result, presented in Fig. E.4, and FLOAD is
the applied load in the test.
The LCP problem (Eq. (E.4)) is solved using Lemkes Complementary
Pivot algorithm with the LCP solver included in OpenOpt [107]. Once the
force balance is solved the problem can be incremented one time step and
the process repeated. The solution at the current time step depends on
the previous one and the solver must be run through approximately 1.1 full
cycles for the transients to fade out and to reach convergence with the pre-
vious cycle. Convergence is assumed to be reached when the film thickness
and the derivative of the film thickness are within 1% of the previous cycle.
The friction force (ftot ) is calculated as the sum of viscous friction force
fhyd and boundary friction force fbd . Boundary friction is calculated as;
Z
fbd = PCP dA, (E.10)

where is the dry friction coefficient and PCP is the average asperity con-
tact pressure, found from the contact mechanics model, see Fig. E.4. The
198 PAPER E.

boundary friction coefficient, taken from Table E.3, was found by running
a reciprocating test in the test apparatus described in section E.2.1, with
a small quantity of lubricant present and at a very low speed to ensure
that no hydrodynamic lubrication effect occurred. The values 30 each side
of the reversal points (0 and 180) are not included in the average, which
translates to the first (and last) 3.75 mm of the stroke. The justification
for this is that, in these friction tests, at the reversal points the stick-slip
phenomenon occurs between the ring and liner as the ring starts moving
from stationary. This leads to a noisy friction value before the measure-
ment settles to a constant value. Of course, these reversals also occur in the
actual tests, however due to the much higher speeds involved and some flex
in the text rig, the ring breaks cleanly away from stationary along the liner
surface and stick-slip does not occur, or at least to a much smaller degree.
For these simulations we require the friction coefficient of boundary lubri-
cated sliding, not of the stick-slip phenomenon. The values are presented
in Table E.3. Hydrodynamic friction is calculated as;
Z  
U h p0
fhyd = + dA, (E.11)
h 2 x

where A is the area of the contact.

E.3.1 Lubricant properties


The ASTM D341 equation was used to calculate the lubricant viscosity at
22 C;
log(log( + 0.7)) = A B log(T + 273.15), (E.12)
where is viscosity, A and B are constants and T is temperature. Using
the values in Table E.2, A = 8.8686 and B = 3.4743, giving a viscosity of
85.66 cSt or 0.072 P a s, at 22C. In the contact it is predicted that the
temperature of the lubricant will be higher than that of the bulk lubricant
due to shear heating. Therefore, the viscosity will be less in the contact
than the value that has been calculated here. However, in the absence of
a measured contact temperature or thermal model, such as in the one by
Morris et al. [43], the bulk lubricant temperature and hence viscosity will
be used in the numerical model.

E.4 Surfaces under investigation


Four different cylinder liner variants are investigated and compared in this
work. 3D surface measurements of the four cylinder liners are shown in
Fig. E.3. The surface measurements, taken with a Talysurf CLI 2000, were
E.4. SURFACES UNDER INVESTIGATION 199

Table E.2: Properties of the lubricant used in the tests.

Density, 843.4 kg/m3


Kinematic Viscosity at 40C, 37 cSt
Kinematic Viscosity at 100C, 6.5 cSt
Longitudinal wave velocity, c 1440 m/s

taken just of an area of plateau on the cylinder liner surface - any large
honing grooves were avoided. These surfaces will be used to calculate ho-
As investigated previously by Spencer
mogenized coefficients A0 , B0 and h.
et al. [108], due to the large wavelength of the honing grooves in relation
to the measurement and contact size, they should not be considered on the
local scale and homogenized together with the surface roughness. However,
this is somewhat of a limitation of this model, as any effect that the honing
grooves have on the lubrication will not be simulated. Each of the measure-
ments is 150x150 m and 139x139 data points, with the exception of the
ANS liner which measured 100x100 m and 93x93 nodes. The reason for the
ANS liner measurement being of a smaller size than the other measurements
was that the density of the deep honing grooves made it impossible to select
an area of 150x150 m without including some of them, in the measurement
aperture. Therefore, a smaller area was chosen to keep the measurement
representative of the plateau surface only. The average roughness parame-
ter, Ra , Elastic Modulus and Poissons ratio for these four surfaces as well
as the opposing piston ring surface are summarised in Table E.3.

The first surface, denoted as STD55, is a standard production grey cast


iron cylinder liner with a honing angle of 55 and Ra = 0.196m. The second
surface, STD35 is also a grey cast iron cylinder liner, but with a honing
angle of 35 and a slightly rougher surface finish of Ra = 0.263m. The third
surface, denoted ANS is a grey cast iron cylinder liner treated with ANS

Triboconditioning . This process is incorporated into the honing process
with the use of a special honing tool and fluid [48, 49]. A layer of Tungsten
Disulphide (WS2 ) is deposited on the surface and the topography is also
modified to give a more run-in like surface finish, where Ra = 0.132m. The
final surface, denoted PL, is a grey cast iron liner that is plasma sprayed
with a stainless steel and ceramic composite before being honed with a very
small grit size tool which leaves an approximately 100m thick coating of
stainless steel on the surface with a low roughness of Ra = 0.138m. The
surface is also relatively porus with deep holes in the stainless steel rather
than deep honing grooves of the STD55, STD35 and ANS surfaces.
200 PAPER E.

1 1
0 0

hr (m)

hr (m)
-1 -1
-2 -2
-3 -3
-4 -4
-5 -5
-6 -6
-7 -7
150 150
125 125
0 75
100
m ) 0 75
100
m )
25
50
y1 (75m 100 50 ( 25
50
y1 (75m 100 50 (
) 125
25 y2 ) 125
25 y2
150 0 150 0

(a) STD55. (b) STD35.

1 1
0 0
hr (m)

hr (m)
-1 -1
-2 -2
-3 -3
-4 -4
-5 -5
-6 -6
-7 -7
100 150
75 125
0 50 m ) 0 75
100
m )
25
y1 (50m 25 ( 25
50
y1 (75m 100 50 (
) 75 y2 ) 125
25 y2
100 0 150 0

(c) ANS. (d) PL.

Figure E.3: Surface images.

E.4.1 Calculation of flow factors and asperity contact pres-


sures
When solving the force balance, Eq. (E.9), asperity contact pressure as a
function of separation is required. Also, as discussed in section E.3, ho-
mogenized coefficients are used to include the effects of surface roughness
when solving Eq. (E.4) over the smooth global domain and must be calcu-
lated too. The calculation of these parameters for the surfaces illustrated
in Fig. E.3 is now discussed.
When the surfaces come into contact, the deformation and asperity
contact pressure are found using an fft-accelerated, boussinesq-type elasto-
plastic contact mechanics model. This model, unlike the more common
approach by Greenwood and Tripp [20], numerically deforms a real, mea-
sured, surface topography to give the average asperity contact pressure as
a function of separation. The technique is described in detail by Sahlin et
al. [29, 30] and will not be repeated here. The piston ring, Ra = 0.066m,
is rather smooth compared to the liner surfaces and is therefore assumed to
perfectly smooth during the numerical simulations. The deformation and
asperity contact pressure results are thus obtained by loading a smooth
E.4. SURFACES UNDER INVESTIGATION 201

Table E.3: Roughness parameters and dry friction coefficients for each cylin-
der liner sample.

Sample Ra (m) Fric. Coef. () El. Mod. (GPa) Poiss.s ratio


STD55 0.196 0.1274 140 0.3
STD35 0.263 0.1249 140 0.3
ANS 0.132 0.1142 140 0.3
PL 0.138 0.1115 140 0.3
Pist. Ring 0.066 N/A 220 0.3

surface against the roughness measurements depicted in Fig. E.3. These


surface measurements, described in section E.4, were interpolated onto a
512x512 grid before being input into the contact mechanics model. To cal-
culate the asperity contact pressure, the elastic modulus and Poissons ratio
of each of the surfaces is needed. These values are given in Table E.3. In the
model plastic deformation is assumed to occur if the local asperity contact
pressure exceeds 1% of the elastic modulus. The resulting asperity contact
pressure as a function of average separation between the surfaces is given
for the standard liner surface in Fig. E.4.
To take into account the effect of the surface roughness, the homogenized
are calculated. For each of the surfaces in Fig. E.3,
coefficients A0 , B0 and h
the following Partial Differential Equations (PDEs),

y h3 y 1

= 0 in Y, (E.13)

1 (x, 0, y2 ) + 1 = 1 (x, l1 , y2 ) ,
1 (x, y1 , 0) = 1 (x, y1 , l2 ) .

y h3 y 2

= 0 in Y, (E.14)

2 (x, 0, y2 ) = 2 (x, l1 , y2 ) ,
2 (x, y1 , 0) + 1 = 2 (x, y1 , l2 ) .

y h3 y 0

= y (he1 ) in Y, (E.15)

0 (x, 0, y2 ) = 0 (x, l1 , y2 ) ,
0 (x, y1 , 0) = 0 (x, y1 , l2 ) .

are solved with periodic boundary conditions in both the y1 and y2 directions
over the measurement domain, i.e., the cell of periodicity, Y = (0, l1 )(0, l2 ),
202 PAPER E.

where h describes the clearance between the rough surfaces and 1 , 2 and
0 are local scale variables. Instead of solving these equations for each pair
of values x = (x1 , x2 ) , they are solved for a range of separations, ,
defined as the distance between the rough surface and the smooth counter-
surface (the piston ring). Mathematically:
h = hr + , (E.16)
where hr represents the surface roughness (Fig. E.3). In order to obtain
sufficiently well resolved homogenized coefficients, is taken as a range of
distances between 0.1-55 m.
Once 1 , 2 and 0 have been calculated they can be integrated/averaged
to give the homogenized coefficients A0 , B0 and h in Eq. (E.4) needed to
solve the averaged form of the Reynolds equation. The explicit expressions
read:
for the coefficients A0 , B0 and h
1 1
Z
a11 (x) = h3 dy, (E.17)
l2 Y y1
1 2
Z
a12 (x) = h3 dy, (E.18)
l1 Y y1
1 1
Z
a21 (x) = h3 dy, (E.19)
l2 Y y2
1 2
Z
a22 (x) = h3 dy, (E.20)
l1 Y y2
 
1
Z
b1 (x)
= he1 h3 y 0 dy, (E.21)
b2 (x) l1 l2 Y
is defined in Eq. (E.7). The results of these calculations for a11 , a22
while h
and b1 for the STD55 surface are shown in Fig. E.4.

E.5 Results and Discussion


In the following sections, friction and film thickness results from both the
numerical model and experimental setup will be presented and compared.
Finally, there will be a deep-dive into selected results to investigate the
similarities and differences between the numerical and experimental film
thickness and friction in a time resolved format.

E.5.1 Friction
In Fig. E.5, the friction coefficient averaged through the reciprocating cycle
is presented for each of the four surfaces. On each plot both the experimental
E.5. RESULTS AND DISCUSSION 203

1.0 350
a11
0.9
Normalised Flow Factor

a22

Contact Pressure (MPa)


300
b1
0.8 Contact Pressure
250
0.7
0.6 200

0.5 150
0.4
100
0.3
50
0.2
0.1 0
101 100 101 102
Separation (m)

Figure E.4: Asperity Contact Pressures and Flow Factors for the STD55
liner.

values and the simulated values are given. Fig. E.5a gives the friction map
for the STD55 surface. It can be observed that the correlation between
the measured and numerically predicted film thickness is very good, with
the two surfaces sitting almost exactly on top of each other. As would
be expected, the friction coefficient is highest at 200 N and 2.5 Hz, i.e.
the highest load and lowest speed. The physical explanation is that the
entraining speed is insufficient to generate an oil film to support such a high
load. A friction coefficient of around 0.1 at this point suggests a majority of
boundary lubrication, as the dry friction coefficient was found to be 0.1274
for the standard surface (see Table E.3). However, it should be remembered
that this is a reciprocating test and therefore the lubrication regime will vary
throughout the stroke. This will be investigated further in the following
sections where film thickness is evaluated and the friction contributions
divided into boundary and viscous components. Looking at the rest of
the friction map, as the speed increases, or the load reduces, the average
friction coefficient drops in value. This indicates moving further to the
right on the Stribeck curve as lower load, or higher speed, allows for a
thicker oil film to be generated and the majority lubrication regime becomes
either mixed or full film rather than boundary. The lowest average friction
coefficient occurs at the lightest load and highest speed; 40 N and 17.5 Hz.
Here the thickest film is observed and the asperity friction contributes the
least to the total friction. Overall, the simulation slightly overestimates the
friction coefficient. There are a number of reasons for this, e.g., the viscosity
or boundary friction coefficient used in the simulation could be slightly
higher than what it actually was during the experiment. It is reasonable to
204 PAPER E.

Avg. Friction Coefficient

Avg. Friction Coefficient


0.11 0.11
0.1 0.1
0.09 0.09

Experiment

Experiment
Simulation

Simulation
0.08 0.08
0.07 0.07
0.06 0.06
0.05 0.05
0.04 0.04
0.03 0.03
0.02 0.02
0.01 0.01
2.5 60 40 2.5 60 40
5.0 100 80 5.0 100 80
Sp7.5 120 ) Sp7.5 120 )
eed10.0
(Hz12.5
140 d (N eed10.0
(Hz12.5
140 d (N
) 15.0
17.5 200
180
160
Loa ) 15.0
17.5 200
180
160
Loa

(a) STD55. (b) STD35.


Avg. Friction Coefficient

Avg. Friction Coefficient


0.11 0.11
0.1 0.1
0.09 0.09

Experiment

Experiment
Simulation

Simulation
0.08 0.08
0.07 0.07
0.06 0.06
0.05 0.05
0.04 0.04
0.03 0.03
0.02 0.02
0.01 0.01
2.5 60 40 2.5 60 40
5.0 100 80 5.0 100 80
Sp7.5 120 ) Sp7.5 120 )
eed10.0
(Hz12.5
140 d (N eed10.0
(Hz12.5
140 d (N
) 15.0
17.5 200
180
160
Loa ) 15.0
17.5 200
180
160
Loa

(c) ANS. (d) PL.

Figure E.5: Average friction coefficient maps.

assume that the shear heating of the lubricant during the test, could raise
the oil temperature and lower its viscosity, thereby reducing the viscous
friction compared to that predicted by the numerical simulation. The only
exception to the simulation having slightly higher values for the STD55
liner (Fig. E.5a) is at 40 N load, where the experimental friction coefficient
seems to increase from 60 N and become slightly larger than the simulated
value. It is suggested that this could be caused by the friction force sensor
detector limit in the experiment. At 40 N, the friction force is very small
(40 N multiplied by a friction coefficient of around 0.02) and it may be too
small for the friction force sensor to be accurately recorded.
Moving onto the STD35 sample and the results presented in Fig. E.5b.
The simulated friction map, for this surface, is found to have marginally
higher values than the STD55 liner. This occurs because the surface is
slightly rougher exhibiting an Ra value of 0.263 m as opposed to 0.196 m
for the STD55 surface. This means that for the same operating conditions
the surfaces come into direct contact sooner, which results in a higher per-
centage of boundary friction. However, when comparing the friction maps
for the measured data, the STD35 liner exhibits lower friction as the load
increases in comparison to the STD55 liner. There are two proposed ex-
planations for this. Firstly, the reduction in honing angle (from STD55 to
STD35) could have increased the oil film thickness, as shown by Jocsak et
E.5. RESULTS AND DISCUSSION 205

al. [28] and thereby reduced both the viscous shear of the lubricant and
likelihood of boundary lubrication, which in turn cancels out the increase
in surface roughness. The surface roughness measurements, depicted in
Fig. E.3, only contain an area of the plateau region with no major honing
grooves. This means that the effect from the honing angle is not consid-
ered fully in the numerical simulations. This is the reason why the effect
of honing angle, appears in the measured friction data and not in the data
predicted by the simulation. Secondly, and along the same lines, the surface
roughness could have helped the formation of an oil film by contributing to
the converging gap in the inlet. This would also help increase the oil film
thickness and cancel out the effect of a rougher surface. This would also not
be predicted by the simulations, because the roughness is only considered
in the local scale and no inter-asperity cavitation, and therefore no lift, is
generated.


The third friction map, shown in Fig. E.5c, presents the data for the ANS
(Triboconditioning ) surface. This surface, along with the final plasma
coated surface, PL, with friction map in Fig. E.5d, shows significantly lower
average friction coefficients than the STD55 and the STD35 liner surfaces,
across the whole friction map. This is predominately due to the reduction in
surface roughness, which at 0.132 m is merely half that of the STD55 liner
and half of the STD35 liner. The boundary friction coefficient, although
lower due to the Tungsten Disulphide coating, is broadly similar to that of
the STD55 liner (0.1142 for the ANS liner compared to 0.1274 for the STD55
liner). The simulation predicts friction coefficients less than the STD55 and
the STD35 surfaces, however the experimental friction coefficients are even
lower. The slight increase in friction at 40N load is, as previously suggested,
most likely due to the friction sensor detector limit.
The final friction map, i.e. the one in Fig. E.5d, presents data for the
plasma coated liner (PL). Here the simulation and experimental data match
very well. The friction coefficient is very low, like with the ANS liner most
likely because of the small Ra value of 0.138 m. The boundary friction
coefficient of 0.1115 is less than the STD55 and STD35 liners but not signif-
icantly so. The smoother surface provides for full film lubrication at lower
speeds and higher loads and delays the onset of boundary lubrication as the
speed reduces or the load increases.

E.5.2 Film thickness


In this section, the numerically simulated film thickness and the film thick-
ness measured with ultrasound will be compared. Fig. E.6, presents the nu-
merically and experimentally obtained film thickness data, for the STD55,
STD35 and ANS liner. There is no data for the plasma liner. This is be-
206 PAPER E.

cause the ultrasound technique was not able to measure the film thickness
through the plasma coating due to attenuation of the ultrasound signal in
the coating.
In each subfigure, seven surfaces are plotted, one for each of the sliding
speeds, with position on the x and load on the y axes. Only one sliding
direction can be shown, in all cases this is from the left to the right of
the image in Fig. E.1, with the ring orientated in the direction shown in
Fig. E.2. The stroke in the other direction, from right to left, was not
considered due to potential anomalies in the ultrasound measurement due
to cavitation. This effect is discussed in great detail in a previous work
by the same authors [101]. The simulated film thickness maps, contain
50 steps for each stroke. As was discussed in Section 2.2 and illustrated
in Fig. E.1, the ultrasound technique measured the film thickness at five
discrete locations on the cylinder liner surface, therefore the experimental
plots have five points for each stroke. On the z axis the Minimum Oil
Film Thickness (MOFT) is given. At a given sensor position this is the
smallest film thickness that occurs between the cylinder liner surface and the
piston ring as the ring passes over the sensor. A deconvolution operation is
performed to get the true minimum rather than the average over the sensor
width. This has already been explained in [101]. In the simulations, the
definition of MOFT is the distance between the mean plane of the surface
roughness and the (smooth) ring profile. All plots show the same trend
in film thickness, where the MOFT increases towards the mid-stroke - as
the entraining speed increases, and then reduces - as the ring comes to a
halt at the end of the stroke. The film thickness maps, from simulations
and experiments, are all skewed so that the maximum MOFT occurs just
past mid-stroke, where the entraining speed is greatest. This is due to the
time dependent nature of the oil film build-up, and time dependence is also
considered when solving the modified Reynolds equation, Eq. (E.4), during
the numerical simulation.
At low speeds the simulated and experimental film thicknesses, for all
three liner surfaces, compare well. According to the ultrasound measure-
ments, the ANS coated liner surface seems to generate a lower film thickness
than the other two liners and this is also what the numerical simulation pre-
dicts. There is a greater difference at higher speeds where measured film
thickness is significantly lower that the value computed by the model. Due
to the fact that the experimentally measured and numerically simulated
friction compare very well at this point, see Fig. E.5a, it is suggested that
this difference occurs because the ultrasonic film thickness measurement is
underestimating the film thickness. At high speeds more cavitation occurs
in the trailing portion of the contact, as shown by Dellis and Arcoumanis
[32] and also from the results of the numerical model. Unfortunately, the
E.5. RESULTS AND DISCUSSION 207

5.0 Hz 2.5 Hz 5.0 Hz 2.5 Hz


3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz 3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz
3.0 3.0

2.6 2.6
MOFT (m)

MOFT (m)
2.2 2.2

1.8 1.8
1.4 1.4
1.0 1.0
0.6 0.6
0.2 0.2

40 40
60 60
80 80
Lo

Lo
100 100
120 120
a

a
140 140
d(

d(
160 160
180 TB T 180 S1 S5 S1
N)

N)
TB TB TB S1 S5 S1 S5
200 TB TB 200 S1 S5 S1 S5 S1 S5
B S5
Position Sensors

(a) STD55 Simulation. (b) STD55 Experiment.

5.0 Hz 2.5 Hz 5.0 Hz 2.5 Hz


3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz 3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz
3.0 3.0

2.6 2.6
MOFT (m)

MOFT (m)

2.2 2.2

1.8 1.8
1.4 1.4
1.0 1.0
0.6 0.6
0.2 0.2

40 40
60 60
80 80
Lo

Lo

100 100
120 120
a

140 140
d(

d(

160 160
180 TB T 180 S1 S5 S1
N)

N)

TB TB TB S1 S5 S1 S5
200 TB TB 200 S1 S5 S1 S5 S1 S5
B S5
Position Sensors

(c) STD35 Simulation. (d) STD35 Experiment.

5.0 Hz 2.5 Hz 5.0 Hz 2.5 Hz


3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz 3.4 17.5 Hz 15.0 Hz 12.5 Hz 10.0 Hz 7.5 Hz
3.0 3.0

2.6 2.6
MOFT (m)

MOFT (m)

2.2 2.2

1.8 1.8
1.4 1.4
1.0 1.0
0.6 0.6
0.2 0.2

40 40
60 60
80 80
Lo

Lo

100 100
120 120
a

140 140
d(

d(

160 160
180 TB T 180 S1 S5 S1
N)

N)

TB TB TB S1 S5 S1 S5
200 TB TB 200 S1 S5 S1 S5 S1 S5
B S5
Position Sensors

(e) ANS Simulation. (f) ANS Experiment.

Figure E.6: Film thickness comparison of three liner samples (for sensor
positions refer to Fig. E.1). T refers to the reversal point closest to the
load cell and B refers to the reversal point closest to the reciprocating arm.

effect of cavitation on the speed of sound and density in a medium confined


between two surfaces in relative motion has, to the authors knowledge not
yet been researched. However, these properties are needed to translate the
measured reflection coefficient into film thickness, using Eq. (E.3), and the
uncertainty in these values could lead to the errors in film thickness mea-
208 PAPER E.

surement at high speed.

E.5.3 Time resolved data analysis


Five load and speed combinations will be investigated in more detail here.
These are listed in Table E.4. Four of the test points represent the corners
of the speed/load map and the fifth is in the middle, giving three points
which are predominantly mixed lubrication, one boundary lubrication and
one full-film. The lowest load chosen is 60 N rather than 40 N, due to
the potential inaccuracy in the friction force sensor at 40 N as discussed in
Section 5.1. For each test point the STD55 and PL liner will be investi-
gated further. These liners were chosen for two reasons - firstly the STD55
liner is one of the higher friction surfaces and the plasma a lower friction
surface so differences should be apparent, and secondly because these two
liners are production surfaces, while the STD35 and ANS liners are proto-
type surfaces. The friction data for the STD55 and the PL liner will be
further investigated by looking at the boundary and the viscous friction, for
the five speed and load combinations independently. The data will also be
examined in a time resolved manner. The simulated friction is split into

Table E.4: Deep dive test points.

Test Lub. Regime Load (N) Speed (Hz)


1 Full Film 60 17.5
2 Mixed 200 17.5
3 Mixed 60 2.5
4 Mixed 120 10.0
5 Boundary 200 2.5

viscous and boundary friction components for each of the test points. The
results are given in Fig. E.7. It is interesting to see the contributions from
the viscous and boundary friction so that when the time resolved data is
discussed, the results can be better understood. The lubrication regime of
the first point, 60 N and 17.5 Hz, is predominantly hydrodynamic. The split
between viscous and boundary friction is approximately 50/50, with most of
the boundary contribution occurring near the reversal points and the mid-
stroke being mostly hydrodynamic. The second test point, i.e. 200 N and
17.5 Hz, shows a similar viscous friction component but greatly increased
boundary friction. This is not so surprising, since the high entraining speed
is maintained, leading to similar viscous friction, albeit slightly increased
due to the higher load reducing the film thickness and thereby increasing
viscous shear. The increased load also leads to more severe boundary lubri-
cation at the reversal points. In the third test point; 60 N and 2.5 Hz, the
E.5. RESULTS AND DISCUSSION 209

25.0
Test Point 5
22.5
200N 2.5Hz
Average Friction Force (N)

20.0

17.5

15.0

12.5

10.0 Test Point 1 Test Point 2 Test Point 3 Test Point 4


60N 17.5Hz 200N 17.5Hz 60N 2.5Hz 120N 10.0Hz
7.5

5.0
Viscous
2.5
Boundary
0.0
STD55 PL STD55 PL STD55 PL STD55 PL STD55 PL

Figure E.7: Breakdown of simulated friction into viscous and boundary


components.

contact is operating almost exclusively in the boundary lubrication regime.


The reason for this is the very low entraining speed inhibiting the formation
of a significant hydrodynamic pressure profile and generation of hydrody-
namic lift. The boundary friction contribution at the fourth test point; 120
N and 10.0 Hz, is similar to the one in test point 3. The increased load,
reducing film thickness, is offset by a greater entraining speed which will
allow a greater hydrodynamic film to be generated. The viscous shear is
much greater, however, due to the increased entraining speed. In the final
test point; 200 N and 2.5 Hz, the friction force is composed almost entirely
from boundary friction, due to the very high load and low entraining speed.
The average friction force for the STD55 liner of just over 20 N, is almost
equal in magnitude to the applied load of 200 N multiplied by the friction
coefficient of 0.1274. This indicates very little contribution to load support
by hydrodynamic film generation.
In the majority of the test cases, it can be observed that the friction force
is dominated by boundary friction. However, it must be remembered that
the power loss, which is what we are really interested in, is equal to friction
force multiplied by velocity. As the majority of the boundary friction occurs
around the reversal points (where the velocity is very low) and most viscous
friction occurs at mid-stroke (when the velocity is, relatively, much higher)
then the viscous contribution to power loss is much greater.
Fig. E.8, shows the friction and film thickness for the first test point;
60 N and 17.5 Hz. In this plot, and in all following plots, the solid lines
are for the STD55 liner and the dashed lines are for the plasma liner. The
x-axis gives the Plint crank angle. There is little difference between the
STD55 and PL liner surfaces. A physical explanation to is that the surface
topography only very weakly influences the total friction in the full film
lubrication regime. The film thickness is greater between 0-180 due to
the longer converging gap in that direction (see Fig. E.2). The second test
210 PAPER E.

Minimum Film Thickness (m)


30
STD55 - Experiment
STD55 - Simulation
2.0
20 STD55 - h0

Friction Force (N)


PL - Experiment
10 PL - Simulation 1.5
PL - h0

0
1.0

10
0.5
20

30 0.0
0 45 90 135 180 225 270 315 360
Plint Spindle Angle ( )

Figure E.8: Test Point 1 - 60 N and 17.5 Hz.

point, with time resolved friction data in Fig. E.9, is for 200 N and 17.5
Hz, which is the highest speed and highest load. Under these conditions the
contact has entered the mixed lubrication regime. At the reversal points,
the friction is very high with virtually all of the load supported by the
asperities. For the STD55 liner, with no hydrodynamic film generated, the
friction force is almost equal to the normal load of 200 N, multiplied by the
dry friction coefficient, found in Table E.3. With a smoother surface, the PL
liner maintains an oil film with less asperity contact and some of the load
is always supported by hydrodynamic lubrication, giving a considerably
Minimum Film Thickness (m)

30
STD55 - Experiment
STD55 - Simulation
2.0
20 STD55 - h0
Friction Force (N)

PL - Experiment
10 PL - Simulation 1.5
PL - h0

0
1.0

10
0.5
20

30 0.0
0 45 90 135 180 225 270 315 360
Plint Spindle Angle ( )

Figure E.9: Test Point 2 - 200 N and 17.5 Hz.


E.5. RESULTS AND DISCUSSION 211

lower friction coefficient at the reversal points. Even at mid-stroke the


lubrication regime for the STD55 liner never becomes fully hydrodynamic
and so this explains the larger friction force compared to the PL liner here.
The simulated and measured results compare well. The only significant
difference is that the simulation overestimates the friction for the PL liner
at the reversal points, suggesting that the onset of asperity contact occurs
sooner in the simulation. The time resolved data for the third test point

Minimum Film Thickness (m)


30
STD55 - Experiment
STD55 - Simulation
2.0
20 STD55 - h0
Friction Force (N)

PL - Experiment
10 PL - Simulation 1.5
PL - h0

0
1.0

10
0.5
20

30 0.0
0 45 90 135 180 225 270 315 360
Plint Spindle Angle ( )

Figure E.10: Test Point 3 - 60 N and 2.5 Hz.

- 60 N and 2.5 Hz, which is the lowest load and lowest speed is presented
in Fig. E.10. As in the previous test point, the contact operates in mixed
lubrication although with a lower friction force than in the previous test
point. Fig. E.10, reveals that boundary lubrication occurs at the reversal
points, just as in Fig. E.9, however the reduction in friction at mid-stroke
is less significant. This is due to the seven times lower entraining speed.
The measured and simulated friction force match extremely well, with the
same trend for lower friction from the plasma liner echoed in both results.
The friction and film thickness data for the fourth test point is depicted in
Fig. E.11. This point is for medium speed and medium load; 120 N and
10.0 Hz which places the contact predominantly in the mixed regime. The
friction force plot and film thicknesses are broadly similar to that of test
point 2, shown in Fig. E.9. The friction force is lower because the applied
load is lower - 120 N instead of 200 N, but the behaviour is almost identical.
The reason is that, although the load is less, the entraining speed to generate
an oil film is also less. The final test point; 200 N and 2.5 Hz, places the
contact predominantly in the boundary lubrication regime, as illustrated in
Fig. E.12. It can be observed that the increase in film thickness throughout
the stroke is minimal. There is a minor reduction in friction force at the mid-
212 PAPER E.

Minimum Film Thickness (m)


30
STD55 - Experiment
STD55 - Simulation
2.0
20 STD55 - h0

Friction Force (N)


PL - Experiment
10 PL - Simulation 1.5
PL - h0

0
1.0

10
0.5
20

30 0.0
0 45 90 135 180 225 270 315 360
Plint Spindle Angle ( )

Figure E.11: Test Point 4 - 120 N and 10.0 Hz.

stroke, however it is still very high and comes mainly from asperity friction.
Thus the contact operates predominantly in boundary lubrication. In both
the measured and simulated result there is lower friction with the plasma
liner. The smoother surface of the plasma liner leads to lift-off occurring
sooner, as the entraining speed increases away from the mid-stroke. This is
increasing the hydrodynamic load support and reducing the asperity contact
with lowered boundary friction as a result.
Minimum Film Thickness (m)

30
STD55 - Experiment
STD55 - Simulation
2.0
20 STD55 - h0
Friction Force (N)

PL - Experiment
10 PL - Simulation 1.5
PL - h0

0
1.0

10
0.5
20

30 0.0
0 45 90 135 180 225 270 315 360
Plint Spindle Angle ( )

Figure E.12: Test Point 5 - 200 N and 2.5 Hz.


E.6. CONCLUSIONS 213

E.6 Conclusions
The friction between a piston ring and cylinder liner has been investigated
for four different cylinder liner variants. A reciprocating test rig was used to
simulate the contact and both friction and oil film thickness, measured with
ultrasound, was recorded by means of a specially designed experimental
setup. Moreover, a numerical model was developed and implemented to
calculate friction and film thickness and comparisons were made between the
experimentally measured and numerical results. The following conclusions
are drawn:

Both measured and simulated friction and film thickness, could be


used to clearly differentiate between the investigated surfaces.

The cylinder liner variants with lower surface roughness, i.e. the ANS
and plasma liners, exhibited significantly lower friction, according to
both the measurements and the simulations.

The results for friction from the numerical model and experiment com-
pared well, with similar trends and magnitudes for both.

At low speeds the measured friction and film thicknesses were very
similar. However, at higher speeds the simulation predicts a larger
film thickness than that observed in the experiment. Due to the good
correlation between the measured and simulated friction, it is proposed
that this difference is due to imprecise values of lubricant density
and speed of sound in the lubricant, caused by increased lubricant
cavitation, being used when processing the ultrasound data.

E.7 Acknowledgements
Scania AB is thanked for providing components, technical data and discus-
sions. Applied Nano Surfaces is thanked for allowing the publication of this

work and providing the ANS Triboconditioning cylinder liner. Loughbor-
ough University are thanked for performing the surface measurements of
the cylinder liner samples.

E.8 Funding
The authors from the Leonardo Centre would like to acknowledge the EP-
SRC Encyclopaedic Program Grant for funding their contribution to this
work. Support of all the program industrial partners is acknowledged. The
authors from LTU would like to thank Stiftelsen for Strategisk Forskning
214 PAPER E.

(SSF) and ProViking for funding their contribution to this work and the
Swedish Research School in Tribology for funding Andrew Spencers place-
ment at Sheffield University.
References

[1] Exxon Mobil. 2012 the outlook for energy: A view to 2040. Technical
report, Exxon Mobil, 2012.

[2] European Commission. White paper - roadmap to a single european


transport area - towards a competitive and resource efficient transport
system. Technical report, European Commission, 2011.

[3] D.E. Richardson. Review of power cylinder friction for diesel engines.
Transactions of the ASME: Journal of Engineering for Gas Turbines
and Power, 122:506519, 2000.

[4] B.S. Andersson. Company perspectives in vehicle tribology - Volvo.


17th Leeds-Lyon Symposium on Tribology - Vehicle Tribology, Tribol-
ogy Ser., Elsevier., 18:503506, 1991.

[5] J.B. Heywood. Internal Combustion Engine Fundamentals. McGraw-


Hill, 1988.

[6] O. Pinkus, D.F. Wilcock, and T.M. Levinson. Reduction in tribolog-


ical energy losses in the transportation and electric utilities sectors.
U.S Department of Energy, 1:1366, 1985.

[7] N.W. Bolander, B.D. Steenwyk, F. Sadeghi, and G.R. Gerber. Lu-
brication regime transitions at the piston ring-cylinder liner interface.
Proc. IMechE Part J: Engineering Tribology, 219:1931, 2005.

[8] J. A. Spearot. Friction, wear, health, and environmental impacts


- tribology in the new millennium. A keynote lecture at the STLE
Annual Meeting, Nashville, Tennessee, May 2000, 2000.

[9] AVL. Excite Piston & Rings, https://www.avl.com/web/ast/excite,


accessed 2013-10-09.

[10] Ricardo. RINGPAK, http://www.ricardo.com/en-gb/what-we-


do/software/products/ringpak/, accessed 2013-11-09.

215
216 REFERENCES

[11] P.C. Mishra, H. Rahnejat, and P.D. King. Tribology of the ring-bore
conjunction subject to a mixed regime of lubrication. Proc. IMechE
Part C: J. Mechanical Engineering Science, 223:987998, 2009.

[12] Y. Hu, H.S. Cheng, T. Arai, Y. Kobayashi, and S. Aoyama. Numerical


simulation of piston ring in mixed lubrication, a nonaxisymmetrical
analysis. ASME: Journal of Tribology, 116:470478, 1994.

[13] O. Akalin and G.M. Newaz. Piston ring-cylinder bore friction mod-
eling in mixed lubrication regime: Part I, analytical results. ASME:
Journal of Tribology, 123:211218, 2001.

[14] M.T. Ma, I. Sherrington, and E.H. Smith. Analysis of lubrication and
friction for a complete piston-ring pack with an improved oil availabil-
ity model Part 1: circumferentially uniform film. Proc. IMechE Part
J: Engineering Tribology, 211:115, 1997.

[15] M.T. Ma, E.H. Smith, and I. Sherrington. Analysis of lubrication and
friction for a complete piston-ring pack with an improved oil availabil-
ity model part 2: circumferentially variable film. Proc. IMechE Part
J: Engineering Tribology, 211:1727, 1997.

[16] N.W. Bolander, B.D. Steenwyk, A. Kumar, and F. Sadeghi. Film


thickness and friction measurement of piston ring cylinder liner con-
tact with corresponding modeling including mixed lubrication. Pro-
ceedings of ICEF04, 2004:811821, 2004.

[17] S.D. Gulwadi. A mixed lubrication and oil transport model for piston
ring using a mass-conserving algorithm. Journal of Engineering for
Gas Turbines and Power, 120:199208, 1998.

[18] M.T. Ma, I. Sherrington, E.H. Smith, and N. Grice. Development of


a detailed model for piston-ring lubrication in ic engines with circular
and non-circular bores. Tribology International, 30:779788, 1997.

[19] H.G. Elrod. A cavitation algorithm. Transactions of the ASME,


103:350354, 1981.

[20] J.A. Greenwood and J.H. Tripp. The contact of two nominally flat
rough surfaces. Proc. IMechE, 185:625633, 1971.

[21] M-T. Ma, E.H. Smith, and I. Sherrington. A three-dimensional anal-


ysis of piston ring lubrication. part 1 : modelling. Proc. IMechE,
209:114, 1995.
REFERENCES 217

[22] K. Liu, Y.B. Xie, and C.L. Gui. Two-dimensional lubrication study
of the piston ring pack. Proc. IMechE Part J: Engineering Tribology,
212:215220, 1998.

[23] N. Patir and H.S. Cheng. An average flow model for determining
effects of three-dimensional roughness on partial hydrodynamic lubri-
cation. ASME Journal of Lubrication Technology, 100:1217, 1978.

[24] M.-T. Ma, E.H. Smith, and I. Sherrington. Implementation of an


algorithm to model the starved lubrication of a piston ring in distorted
bores: prediction of oil flow and onset of gas blow-by. Proc. IMechE
Part J: Engineering Tribology, 210:2944, 1996.

[25] L. Liu and T. Tian. Modeling piston ring-pack lubrication with con-
sideration of ring structural response. SAE Technical Paper Series,
2005-01-1641:114, 2005.

[26] O. Akalin and G.M. Newaz. Piston ring-cylinder bore friction model-
ing in mixed lubrication regime: Part ii - correlation with bench test
data. ASME: Journal of Tribology, 123:219223, 2001.

[27] S.K. Michail and G.C. Barber. The effects of roughness on piston ring
lubrication Part 1: Model development. STLE: Tribology Transac-
tions, 38:1926, 1995.

[28] J. Jocsak, Y. Li, T. Tian, and V.W. Wong. Modeling and optimizing
honing texture for reduced friction in internal combustion engines.
SAE Technical Paper Series, 2006-01-0647:115, 2006.

[29] F. Sahlin, Larsson. R., A. Almqvist, P.M. Lugt, and P. Marklund.


A mixed lubrication model incorporating measured surface topogra-
phy Part 1: theory of flow factors. Proc. IMechE Part J: Journal of
Engineering Tribology, 224:335351, 2010.

[30] F. Sahlin, R. Larsson, P. Marklund, A. Almqvist, and P.M. Lugt. A


mixed lubrication model incorporating measured surface topography
Part 2: roughness treatment, model validation, and simulation. Proc.
IMechE Part J: Journal of Engineering Tribology, 224:353365, 2010.

[31] O. Reynolds. On the theory of lubrication and its application to mr.


beauchamp towers experiments, including an experimental determi-
nation of the viscosity of olive oil. Philiosophical Transactions of the
Royal Society of London, 177:157234, 1896.
218 REFERENCES

[32] P. Dellis and C. Arcoumanis. Cavitation development in the lubricant


film of a reciprocating piston-ring assembly. Proc. IMechE Part J: J.
Engineering Tribology, 218:157171, 2004.

[33] D. Vijayaraghavan and T.G. Keith. Development and evaluation of


a cavitation algorithm. STLE: Tribology Transactions, 32:225233,
1989.

[34] R. Ausas, M. Jai, and G.C. Buscaglia. A mass-conserving algorithm


for dynamical lubrication problems with cavitation. ASME: Journal
of Tribology, 131:17, 2009.

[35] M. Giacopini, M. T. Fowell, D. Dini, and A. Strozzi. A mass-


conserving complementarity formulation to study lubricant films in
the presence of cavitation. ASME: Journal of Tribology, 132:112,
2010.

[36] A. Spencer. Optimizing surface texture for combustion engine cylinder


liners. Masters thesis, Lule University of Technology, 2010.

[37] R.W. Cottle, J-S. Pang, and R. Stone. The Linear Complementarity
Problem. Society for Industrial & Applied Mathematics, 2009.

[38] A. Almqvist. On the effects of surface roughness in lubrication. PhD


thesis, Division of Machine Elements, Lule
aUniversity of Technology,
2006.

[39] A. Almqvist and J. Dasht. The homogenization process of the


Reynolds equation describing compressible liquid flow. Tribology In-
ternational, 39:9941002, 2006.

[40] R.A. Mufti, M. Priest, and R.J. Chittenden. Analysis of piston assem-
bly friction using the indicated mean effective pressure experimental
method to validate mathematical models. Proc. IMechE Part D: J.
Automobile Engineering, 222:14411457, 2008.

[41] H.M. Stanley and T. Kato. Fft-based method for rough surface con-
tact. Transactions of the ASME: Journal of Tribology, 119:481485,
1997.

[42] A. Almqvist, F. Sahlin, R. Larsson, and S. Glavatskih. On the dry


elasto-plastic contact of nominally flat surfaces. Tribology Interna-
tional, 40:574579, 2007.
REFERENCES 219

[43] N. Morris, R. Rahmani, H. Rahnejat, P.D. King, and B. Fitzsimmons.


Tribology of piston compression ring conjunction under transient ther-
mal mixed regime of lubrication. Tribology International, 59:248258,
2013.

[44] H. Rahnejat, S. Balakrishnan, P.D. King, and S. Howell-Smith. In-


cylinder friction reduction using a surface finish optimization tech-
nique. Proc. IMechE Part D: J. Automobile Engineering, 220:1309
1318, 2006.

[45] I. Etsion. State of the art in laser surface texturing. Transactions of


the ASME: Journal of Tribology, 127:248253, 2005.

[46] A. Almqvist, E.K. Essel, J. Fabricius, and P. Wall. Reiterated ho-


mogenization applied in hydrodynamic lubrication. IMechE Part J:
Journal of Engineering Tribology, 222:827841, 2008.

[47] A. Vokac and T. Tian. An experimental study of oil transport on


the piston third land and the effects of piston and ring designs. SAE
Technical Paper Series, 2004-01-1934:118, 2004.

[48] B. Zhmud, G. Flores, C. Verpoort, and U. Morawitz. A novel sur-


face finishing process for improving tribological properties of cylinder
bores. In Proc. 6th VDI-Conference Zylinderlaufbahn, Kolben, Pleuel,
2012.

[49] B. Zhmud, E-B. Akerlund, S. Jacobson, J. Hardell, L. Hammerstrom,


and R. Ohlson. Ans triboconditioning: In-manufacture running-in
process for improving tribological properties of mechanical parts made
of steel or cast iron. In Proc. 18th International Colloquium Tribology
- Industrial and Automotive Lubrication, 2012.

[50] H.M. Uras and D.J. Patterson. Measurement of piston and ring assem-
bly friction instantaneous imep method. SAE Technical Paper Series,
830416, 1983.

[51] R.A. Mufti and M. Priest. Experimental evaluation of piston-assembly


friction under motored and fired conditions in a gasoline engine.
ASME: Journal of Tribology, 127:826836, 2005.

[52] S. Furuhama and M. Takiguchi. Measurement of piston frictional


force in actual operating diesel engine. SAE Technical Paper Series,
790855, 1979.
220 REFERENCES

[53] S. Furuhama and S. Sasaki. New device for the measurement of pis-
ton frictional forces in small engines. SAE Technical Paper Series,
831284:3950, 1983.

[54] H. Yoshida, K. Kusama, and Y. Mizuma. Measurement of piston


friction force on firing conditions. ?, ?:17, 1987.

[55] R.S. Dwyer-Joyce, B.W. Drinkwater, and C.J. Donohoe. The mea-
surement of lubricant-film thickness using ultrasound. Proceedings
of the Royal Society A: Mathematical, Physical and Engineering Sci-
ences, 459:957976, 2003.

[56] E.Y. Avan, R.S. Mills, and R.S. Dwyer-Joyce. Ultrasonic imaging of
the piston ring oil film during operation in a motored engine - towards
oil film thickness measurement. SAE International Journal of Fuels
and Lubricants, 3(2):786793, 2010.

[57] H.M. Herbst. Theoretical modeling of the cylinder lubrication in inter-


nal combustion engines and its influence on piston slap induced noise,
friction and wear. PhD thesis, The Faculty of Mechanical Engineering
at the Technical University of Graz, 2007.

[58] G. Haasis and U.P. Weigmann. New honing technique reduces oil
consumption. Industrial Diamond Review, 3/99:205211, 1999.

[59] S.K. Michail and G.C. Barber. The effects of roughness on piston ring
lubrication Part 2: The relationship between cylinder wall surface
topography and oil film thickness. STLE: Tribology Transactions,
38:173177, 1995.

[60] R. Ausas, P. Ragot, J. Leiva, M. Jai, G. Bayada, and G.C. Buscaglia.


The impact of the cavitation model in the analysis of microtextured
lubricated journal bearings. Transactions of the ASME, 129:868875,
2007.

[61] G.E. Forsythe, M.A. Malcolm, and C.B. Moler. Computer methods
for mathematical computations. Prentice-Hall, Englewood Cliffs, New
Jersey, 1976.

[62] R. Levi. Finish on surface ground steel. International Journal of


Machine Tool Design and Research, 2:357367, 1962.

[63] J.D. Huffington. Comparative studies of surfaces by means of contact


area measurements and stylus profile traces. Wear, 16:313324, 1970.
REFERENCES 221

[64] R. Ohlsson, A. Wihlborg, and H. Westberg. The accuracy of fast 3D


topography measurements. International Journal of Machine Tools
and Manufacture, 41:18991907, 2001.

[65] C.Y. Poon and B. Bhushan. Comparison of surface roughness mea-


surements by stylus profiler, AFM and non-contact optical profiler.
Wear, 190:7688, 1995.

[66] K.H. Guenther, F.K. Bennett, F.K. Urban, and M.F. Tebet. Com-
parative Study of the roughness of oxide thin films, Atomic Force Mi-
croscopy/Scanning Tunneling Microscopy. Plenum Press, 1994.

[67] S. Adi, H. Adi, H-K. Chan, P.M. Young, D. Traini, and A. Yang,
R. Yu. Scanning white-light interferometry as a novel technique to
quantify the surface roughness of micro-sized particles for inhalation.
Langmuir, 24:1130711312, 2008.

[68] H. G. Rhee, T.V. Vorburger, J.W. Lee, and J. Fu. Discrepancies


between roughness measurements obtained with phase-shifting and
white-light interferometry. Applied Optics, 44:59195927, 2005.

[69] T.V. Vorburger, H. G. Rhee, T. B. Renegar, J.F. Song, and A. Zheng.


Comparison of optical and stylus methods for measurement of surface
texture. International Journal of Advanced Manufacturing Technol-
ogy, 33:110118, 2007.

[70] F. Gao, R.K. Leach, J. Petzing, and J.M. Coupland. Surface measure-
ment errors using commercial scanning white light interferometers.
Measurement Science and Technology, 19:113, 2008.

[71] A. Mannelquist, N. Almqvist, and S. Fredriksson. Influence of tip ge-


ometry on fractal analysis of atomic force microscopy images. Applied
Physics A: Materials Science & Processing, 66:S891S895, 1998.

[72] P. Markiewicz and M. Cynthia Goh. Atomic force microscopy probe


tip visualization and improvement of images using a simple deconvo-
lution procedure. Langmuir, 10:57, 1994.

[73] V. Bykov, A. Gologanov, and V. Shevyakov. Test structure for spm


tip shape deconvolution. Applied Physics A: Materials Science & Pro-
cessing, 66:499502, 1998.

[74] J.S. Villarrubia. Algorithms for scanned probe microscope image sim-
ulation, surface reconstruction, and tip estimation. Journal of research
of the National Institute of Standards and Technology, 102:425454,
1997.
222 REFERENCES

[75] H.U. Danzebrink, L. Koenders, G. Wilkening, A. Yacoot, and H. Kun-


zmann. Advances in scanning force microscopy for dimensional metrol-
ogy. Annals of the CIRP, 55/2/2006:841878, 2006.

[76] N. Patir and H.S. Cheng. Application of average flow model to lubri-
cation between rough sliding surfaces. ASME Journal of Lubrication
Technology, 101:220230, 1979.

[77] L.D. Chiffre, P. Lonardo, H. Trumpold, D.A. Lucca, G. Goch, C.A.


Brown, J. Raja, and H.N. Hansen. Quantitative characterisation of
surface texture. CIRP Annals - Manufacturing Technology, 49:635
642, 2000.

[78] Mathworks. Delaunay triangulation. Mathworks website,


http://www.mathworks.com/help/techdoc/ref/delaunay.html, 2010.

[79] A. Almqvist, J. Fabricius, A. Spencer, and P. Wall. Similarities and


differences between the flow factor method by Patir and Cheng and
homogenization. Journal of Tribology, 133:15, 2011.

[80] A. Almqvist, R. Larsson, and P. Wall. The homogenization process of


the time dependent reynolds equation describing compressible liquid
flow. Tribologia - Finnish Journal of Tribology, 26:115, 2007.

[81] Y. Wakuri, T. Hamatake, M. Soejima, and T. Kitahara. Piston


ring friction in internal combustion engines. Tribology International,
25:299308, 1992.

[82] Y.C. Tan and Z.M. Ripin. Frictional behavior of piston rings of small
utility two-stroke engine under secondary motion of piston. Tribology
International, 44:592602, 2011.

[83] C. Arcoumanis, M. Duszynski, H. Flora, and P. Ostovar. Development


of a piston-ring lubrication test-rig and investigation of boundary con-
ditions for modelling lubricant film properties. SAE Technical Paper
Series, 952468, 1995.

[84] L.L. Ting. Development of a reciprocating test rig for tribological


studies of piston engine moving components - part i: Rig design and
piston ring friction coefficients measuring method. SAE Technical
Paper Series, 930685, 1993.

[85] P.M. Lee and R. Chittenden. Consideration of test parameters in


reciprocating tribometers used to replicate ring-on-liner contact. Tri-
bology Letters, 39:8189, 2010.
REFERENCES 223

[86] S.E. Hartfield-Wunsch, S.C. Tung, and C.J. Rivald. Development of a


bench test for the evaluation of engine cylinder components and cor-
relation with engine test results. SAE Technical Paper Series, 932693,
1993.

[87] S.J. S
ochting and I. Sherrington. The effect of load and viscosity on
the minimum operating oil film thickness of piston-rings in internal
combustion engines. Proceedings of the Institution of Mechanical En-
gineers, Part J: Journal of Engineering Tribology, 223:383391, 2009.

[88] D.O. Ducu, R.J. Donahue, and J.B. Ghandhi. Design of capacitance
probes for oil film thickness measurements between the piston ring
and linear in internal combustion engines. Journal of Engineering for
Gas Turbines and Power, 123:633643, 2001.

[89] A. Dhar, A.K. Agarwal, and V. Saxena. Measurement of dynamic lu-


bricating oil film thickness between piston ring and liner in a motored
engine. Sensors and Actuators, A: Physical, 149:715, 2009.

[90] J.S. Courtney-Pratt and G.K. Tudor. An analysis of the lubrication


between the piston rings and cylinder wall of a running engine. Pro-
ceedings of the IMechE., 155:155293, 1946.

[91] T. Seki, K. Nakayama, T. Yamada, A. Yoshida, and T. Takiguchi.


A study on variation in oil film thickness of a piston ring package:
variation of oil film thickness in piston sliding direction. JSAE Review,
21:315320, 2000.

[92] M. Takiguchi, K. Nakayama, S. Furuhama, and H. Yoshida. Varia-


tion of piston ring oil film thickness in an internal combustion engine
- comparison between thrust and anti-thrust sides. SAE paper No.
980563, 980563, 1998.

[93] J.J. Truhan, J. Qu, and P.J. Blau. A rig test to measure friction and
wear of heavy duty diesel engine piston rings and cylinder liners using
realistic lubricants. Tribology International, 38:211218, 2005.

[94] M. Schoenberg. Elastic wave behavior across linear slip interfaces. J.


Acoust. Soc. Am., 68:15161521, 1980.

[95] T. Reddyhoff, R. Dwyer-Joyce, and P. Harper. Ultrasonic measure-


ment of film thickness in mechanical seals. Sealing Technology, 7:711,
2006.
224 REFERENCES

[96] R. Dwyer-Joyce, P. Harper, and B. Drinkwater. Oil film measurement


in polytetrafluoroethylene-faced thrust pad bearings for hydrogenera-
tor applications. Proc. IMechE Part A: Journal of Power and Energy,
220:619628, 2006.

[97] M. K. Wan Ibrahm, D. Gasni, and R. Dwyer-Joyce. Profiling a ball


bearing oil film with ultrasonic reflection. Tribology Transactions,
55:409421, 2012.

[98] J. Zhang, B.W. Drinkwater, and R.S. Dwyer-Joyce. Calibration of


the ultrasonic lubricant-film thickness measurement technique. Mea-
surement Science and Technology, 16:17841791, 2005.

[99] A. Johansson, P.H. Nilsson, R. Ohlsson, C. Anderberg, and B-G.


Rosn. New cylinder liner surfaces for low oil consumption. Tribol-
ogy International, 41:854859, 2008.

[100] O. Sato, M. Takiguchi, T. Aihara, Y. Seki, K. Fujimura, and


Y. Tateishi. Improvement of piston lubrication in a diesel engine by
means of cylinder surface roughness. SAE Technical Paper Series,
2004-01-0604:17, 2004.

[101] E. Y. Avan, A. Spencer, R. Dwyer-Joyce, A. Almqvist, and R. Lars-


son. Experimental and numerical investigations of oil film formation
and friction in a piston ring-liner contact. IMechE Part J: Journal of
Engineering Tribology, Online First:115, 2012.

[102] P. Saad, L. Kamo, M. Mekari, W. Bryzik, V. Wong, N. Dmitrichenko,


and R. Mnatsakanov. Modeling and measurement of tribological pa-
rameters between piston rings and liner in turbocharged diesel engine.
SAE Technical Paper Series, 2007-01-1400:116, 2007.

[103] R.I. Taylor and P.G. Evans. In-situ piston measurements. Proc.
IMechE Part J: Engineering Tribology, 218:185200, 2004.

[104] P. Harper, R.S. Dwyer-Joyce, U. Sjodin, and U. Olofsson. Evaluation


of an ultrasonic method for measurement of oil film thickness in a
hydraulic motor piston ring. In Proceedings of the 31st Leeds-Lyon
Symposium on Tribology, 2004.

[105] R.S. Mills, E.Y. Avan, and R.S. Dwyer-Joyce. Piezoelectric sensors to
monitor lubricant film thickness at piston-cylinder contacts in a fired
engine. Proc. IMechE Part J: Engineering Tribology, 227:100111,
2013.
REFERENCES 225

[106] G. Bayada, M.E.A. Talibi, and K.A. Hadi. Existence and unique-
ness for a non-coercive lubrication problem. Journal of Mathematical
Analysis and Applications, 327:585610, 2007.

[107] D. Kroshko. OpenOpt: Free scientific-engineering software for math-


ematical modeling and optimization. software package downloadable
from http://openopt.org, 2013.

[108] A. Spencer, A. Almqvist, and R. Larsson. A semi-deterministic


texture-roughness model of the piston ring-cylinder liner contact.
Proc. IMechE Part J: J. Engineering Tribology, 225:19, 2011.

You might also like