Download as pdf or txt
Download as pdf or txt
You are on page 1of 136

Copyright Warning

Use of this thesis/dissertation/project is for the purpose of


private study or scholarly research only. Users must comply
with the Copyright Ordinance.

Anyone who consults this thesis/dissertation/project is


understood to recognise that its copyright rests with its
author and that no part of it may be reproduced without the
author’s prior written consent.
CITY UNIVERSITY OF HONG KONG
香港城市大學

Research on Boundary Slippage in Thin Film


Hydrodynamic Lubrication
薄膜流體潤滑邊界滑移現象之研究

Submitted to
Department of Mechanical and Biomedical Engineering
機械及生物醫學工程系
in Partial Fulfillment of the Requirements
for the Degree of Doctor of Philosophy
哲學博士學位

By

GUO Liang
郭亮

December 2015
二零一五年十二月
Abstract

The traditional lubrication models are usually based on the viscous flow of the

lubricant. It is assumed that there is no relative motion at the interface of the

lubricant and the solid surfaces. This assumption has been proved by many

engineering examples and theoretical studies. In this situation, the sliding friction

coefficient is a function of the viscosity of the lubricant, and the friction force can be

decreased by reducing the viscosity of the lubricant directly. However, there is a

limitation to reduce lubricant viscosity in most engineering applications, because

reducing viscosity would lead to the drop of lubricant film thickness, which may

induce wear in contact areas. Recently, a new idea for reducing friction force has

been proposed by using boundary slippage. This requires a very low adhesion force

between the solid surface and the lubricant. To better use the slippage, the effect of

three parameters, interfacial affinity, sliding speed and lubricant’s viscosity on

hydrodynamic lubrication performance was studied through experiments and

theories in this project.

A self-developed slider-on-disc test rig was applied for this study. In order to realize

the simultaneous measurement of lubricant film thickness and friction, a dichromatic

optical interference method was developed. By simultaneously projecting two light

beams of different wavelengths on the lubricating contact, an envelope of intensity-

difference and film thickness variation equivalent to an intensity and film thickness

curve with a longer cycle can be obtained. Using the equivalent intensity curve, the

measurement range of film thickness can be enhanced compared with

monochromatic interferometry. The moving rate of the equivalent intensity curve


corresponding to a rapid change in film thickness was fairly slow, which facilitates

the on-line measurement of lubricant thickness and realizes the simultaneous

measurement of lubricant film thickness and friction.

Two parameters, contact angle and contact angle hysteresis were compared for their

correlation with the hydrodynamic lubricating effect of a slider bearing. Five very

smooth slider surfaces of different materials and three lubricants, polar and non-

polar, which provided contact angles ranging from 40o to 110o, were used in the

bearing tests. Contact angle hysteresis, but not the contact angle, was found to

closely correlate with the hydrodynamic effect. The finding is also supported by an

existing theory based on thermodynamic principles that the adhesive strength, which

can be quantified by potential energy barrier, between the contacting solid and liquid

molecules is a strong function of contact angle hysteresis but not the contact angle if

its value falls into the range of 20o to 140o.

More samples of different interfacial affinity were obtained with Silicon oil 201-500

containing different concentrations of perfluoroenathic acid (C6F13COOH) and a

steel surface. The influence of roughness on hydrodynamic lubrication could be

ignored. Experiments proved again that the contact angle hysteresis determines the

hydrodynamic lubrication behavior. Besides, monotonous correlation between the

boundary yield stress and the potential energy barrier was confirmed based on the

experimental results for the two parameters describing similar phenomena.

The effect of sliding speed of glass disc and viscosity of the lubricant on boundary

slippage was studied. Two types of lubricants, glycerol solutions and PAO oils, were

applied for the study of the viscosity effect. The strong linear relationship between
lubricant’s viscosity and boundary yield stress was found through the experimental

results. Besides, it is also proved that the boundary yield stress increases with sliding

speed. These conclusions are consistent with the slip model proposed by Spikes and

Granick.
CITY UNIVERSITY OF HONG KONG
Qualifying Panel and Examination Panel

Surname: GUO
First Name: Liang
Degree: Doctor of Philosophy
College/Department: Department of Mechanical and Biomedical Engg

The Qualifying Panel of the above student is composed of:

Supervisor(s)
Dr. WONG Pat Lam Patrick Department of Mechanical and Biomedical Engg
City University of Hong Kong

Qualifying Panel Member(s)


Dr. LI Kwok Yan Department of Mechanical and Biomedical Engg
City University of Hong Kong

Dr. WANG Zuankai Department of Mechanical and Biomedical Engg


City University of Hong Kong

This thesis has been examined and approved by the following examiners:
Dr. MA Weiyin Department of Mechanical and Biomedical Engg
City University of Hong Kong

Dr. WANG Zuankai Department of Mechanical and Biomedical Engg


City University of Hong Kong

Dr. WONG Pat Lam Patrick Department of Mechanical and Biomedical Engg
City University of Hong Kong

Dr. SINHA Sujeet Kumar Department of Mechanical Engineering


Indian Institute of Technology, Delhi
ACKNOWLEDGMENT

I would like to express my grateful thanks to my supervisor, Dr. Patrick Wong, for

his guidance and support in many ways, which has made my study in CityU

enjoyable and fruitful. I am deeply influenced by his serious attitude and passion for

research. Besides, my Ph.D has been an amazing experience working with Dr. Wong,

who provided wonderful opportunities in taking part in many conferences, which

broadened my horizons greatly. I would also like to express my sincere thanks to my

co-supervisor, Professor Feng Guo for his many useful instructions and advices to

my study. Thanks go to Dr. Xiaobo Zhou from SKF research center and Dr. George

Wan for providing valuable specimens and excellent suggestions for my research.

I am grateful to Dr. Xia Li, Dr. Xinming Li, Dr. Zhongxue Fu and Mr. Haichao Liu

for their help in supporting my experiments. I appreciate the technical support from

the laboratory technicians, particularly Mr. C. K. Chung, Mr. K.C. Kian and Mr.

W.P. Lau. Acknowledge is given to the Research Grant Council of Hong Kong for

the financial support to my research. I would like to thank City University of Hong

Kong for providing the scholarship for me.

Finally, but by no means least, thanks go to my girlfriend Yanhua Zhao and my

parents for their support to my study.


Contents
1. Introduction and Literature Review ................................................................ 1
1.1 Background ................................................................................................... 1
1.2 Research on boundary slip ............................................................................ 2
1.2.1 Research on boundary slip in history ................................................. 2
1.2.2 Research on boundary slip in recent years ......................................... 3
1.2.2.1 Techniques used for boundary slip research ......................... 3
1.2.2.2 Parameters governing boundary slippage ............................. 6
1.3 Research on boundary slip in lubrication ................................................... 10
1.4 Aims and objectives.................................................................................... 11
1.5 Chapter outline ........................................................................................... 14
2. Development of Test Rig for Thin Film Hydrodynamic Lubrication ......... 16
2.1 Required test rig for the project .................................................................. 16
2.2 Existing set up ............................................................................................ 17
2.3 Friction measurement design ...................................................................... 19
2.4 On-line optical film thickness measurement design ................................... 22
2.4.1 Optical interferometry for lubrication research ................................ 22
2.4.2 Special requirements of optical methods for the project .................. 24
2.4.3 Principle of dichromatic optical interference technique ................... 26
2.4.4 Implementation ................................................................................. 27
2.4.5 Validation for on-line film thickness measurement ......................... 31
2.5 Summary..................................................................................................... 38
3. Boundary Yield Stress Model in Hydrodynamic Lubrication ..................... 39
3.1 Introduction ................................................................................................ 39
3.2 Slippage models .......................................................................................... 40
3.2.1 Slip length model .............................................................................. 40
3.2.2 Boundary yield stress model ............................................................ 40
3.2.3 Non-linear slip length model ............................................................ 42
3.3 Derivation of modified Reynolds equation ................................................ 43
3.4 Solution of modified Reynolds equation .................................................... 47
3.4.1 Control equation for wall slip ........................................................... 47
3.4.2 Multi-linearity algorithm .................................................................. 48
3.4.3 Scalar method ................................................................................... 50
3.5 Numerical results and analyses–Effects of boundary yield stress .............. 50
3.5.1 Normalized load support .................................................................. 50
3.5.2 Normalized friction force ................................................................. 52
3.5.3 Pressure distribution ......................................................................... 53
3.5.4 Lubricant film thickness under different parameters ....................... 54
3.6 Summary .................................................................................................... 56
4. Identifying the Best Interfacial Parameter Correlated with Hydrodynamic
Lubrication....................................................................................................... 57
4.1 Introduction ................................................................................................ 57
4.2 Interfacial parameters ................................................................................. 58
4.2.1 Contact angle .................................................................................... 58
4.2.2 Contact angle hysteresis ................................................................... 60
4.2.3 Kalin’s spreading parameter ............................................................. 60
4.3 Measurement of interfacial parameters ...................................................... 61
4.3.1 Measurement of contact angle .......................................................... 61
4.3.2 Measurement of contact angle hysteresis ......................................... 63
4.3.3 Measurement of surface energy and spreading parameter ............... 64
4.4 Specimen surface and lubricant .................................................................. 65
4.5 Comparison of different parameters ........................................................... 67
4.6 Why CAH is the best? ................................................................................ 77
4.7 Summary .................................................................................................... 78
5. Correlation of Boundary Yield Stress and Interfacial Potential Energy
Barrier .............................................................................................................. 80
5.1 Introduction ................................................................................................ 80
5.2 Specimen surface and lubricant .................................................................. 81
5.3 Results and discussion ................................................................................ 85
5.4 Summary .................................................................................................... 92
6. Sliding Speed and Viscosity Effect on Boundary Slippage .......................... 93
6.1 Introduction ................................................................................................ 93
6.2 Effect of sliding speed on boundary yield stress ........................................ 95
6.3 Effect of viscosity on boundary yield stress ............................................... 96
6.3.1 Specimen surface and lubricant ........................................................ 96
6.3.2 Results and discussion ...................................................................... 98
6.4 Summary .................................................................................................. 105
7. Conclusions and Suggestions for Future Work ........................................... 106
7.1 Summary of main conclusions ................................................................. 106
7.1.1 Development of test rig for thin film hydrodynamic lubrication ... 106
7.1.2 New optical method for lubricating film thickness measurement .. 106
7.1.3 Solution of modified Reynolds equation ........................................ 107
7.1.4 The best interfacial parameter correlated with hydrodynamic
lubrication ...................................................................................... 107
7.1.5 Correlation between boundary yield stress and potential energy
barrier ............................................................................................. 108
7.1.6 Sliding speed and viscosity effect on boundary yield stress .......... 108
7.2 Suggestions for future work ..................................................................... 108
Appendix A: Envelope cycle of dichromatic interference .................................. 123
Appendix B: Ratio of changing rate ..................................................................... 124
Appendix C: Flow chart of numerical analysis ................................................... 125
Appendix D: List of my publications.................................................................... 126
Chapter 1

Introduction and Literature Review

1.1 Background

Energy conservation has come to be an urgent issue for the shortage of non-

renewable energy, such as petroleum, coal and natural gas. Friction and wear in

tribo-pairs are closely related to the performance reliability and energy efficiency of

machines [1, 2]. $41.4 billion would be saved every year at 2006 rates (equivalent to

1.55% of GDP (Gross Domestic Product) in China) if the tribology knowledge was

applied appropriately [3]. Considerable efforts have been made towards friction and

wear reduction, and the usual method of achieving this in engineering is through oil

lubrication, whereby two contact surfaces are separated by a thin oil layer. Friction

and lubrication models for lubricated contact are usually based on the viscous flow

of the lubricant and the no-slip assumption at the interface of the lubricant and the

solid surface. This assumption is valid since most engineering surfaces have good

adhesive force to lubricating oils and the adhesive force is much higher than the

viscous resistance. In this situation, the coefficient of friction is a function of the

viscosity of the lubricant, and the friction force can be reduced by reducing the

viscosity of the lubricant. However, there is a limitation to reduce lubricant viscosity

in most engineering applications, because reducing viscosity would lead to a thinner

lubricant film, which may induce more wear on the solid surface.

For well lubricated contacts, is it possible to further reduce the sliding friction? An

idea of using boundary slippage has been hypothesized in recent years [4, 5]. When

the interface affinity between the oil and the solid surface is lower than the viscous

1
resistance, the oil film may slip on solid surface. The lubrication behavior in

machines can be changed directly by the change of boundary conditions, i.e. slippage.

1.2 Research on boundary slip

1.2.1 Research on boundary slip in history

The discussion about the boundary condition (BC) in fluid mechanics can date back

to 18th century. In 1738, Bernoulli [6] proposed the no-slip boundary condition,

which was proved later by Du Buat [7] and Coulomb [8] (Fig. 1.1(a)). After that,

almost all analyses of fluid mechanic adopted this assumption.

(a) No-slip (b) Stagnant layer (c) Slip length

Fig. 1.1 Boundary conditions

However, some researchers disagreed this assumption, for example, Girard [9] and

Prony [10] believed that there is a stagnant layer between bulk liquid and solid

surface (Fig. 1.1(b)) and the thickness of this layer is related with the properties of

the solid/liquid interface. In this assumption, there is no relative movement between

the stagnant layer and the solid surface, but the bulk liquid slips at the bulk

liquid/stagnant layer interface.

2
In 1823, Navier [11] proposed the slip length model (Fig. 1.1(c)), in which the slip

velocity is believed proportional to the shear stress of the liquid at the solid/liquid

interface, i.e.

𝜕𝑢
𝑢𝑠 = 𝑏 | (1.1)
𝜕𝑧 𝑧=0

where 𝑢𝑠 is slip velocity, and the slip length b is the fictional distance between the

solid surface and the point where the velocity linearly decreases to zero.

After that, many researches have been done for identifying the boundary condition.

Bingham [12] summarized these works. For example, Stokes [13] confirmed the no-

slip BC from his experiments, in which the drag force was compared when water and

mercury flowed from a glass tube. However, Poiseuille [14], Darcy [15], Helmholtz

and Piotrowski [16] found the evidence of boundary slip from their experiments. For

the limitation of experimental conditions at that time, the accuracy was not good

enough. That is why Maxwell [17], Whetham [18], Couette [19] and Ladenburg [20]

doubted about the boundary slip conclusions.

Until the early of 20th century, there was no effective means for accurate boundary

slip measurement. It was believed that the magnitude of boundary slip would be too

small to affect the fluid behavior in macroscopic scale if boundary slip did exist.

Therefore, the no-slip boundary condition was accepted and adopted widely both in

engineering and research.

1.2.2 Research on boundary slip in recent years

1.2.2.1 Techniques used for boundary slip research

3
With the improvement of micro/nano fabrication, many techniques with high

accuracy and sensitivity were gradually developed, such as atomic force microscope

(AFM), surface force apparatus (SFA), capillary channel, particle image velocimetry

(PIV), and fluorescence recovery after photobleaching (FRAP). These state-of-the-

art techniques provide new methods for boundary slip research. Besides, molecular

dynamics simulations (MDS) were widely used in boundary slip research since

1990s. Compared with experimental methods, it is easy to just focus on the effect of

one parameter on boundary slippage and ignore the coupling effect of different

parameters by using MDS. Therefore, the boundary slip research has become a hot

topic since the end of the last century.

In terms of experimental studies, there are two main methods for boundary slip

measurement: direct and indirect slip measurement. Direct slip measurement

includes particle image velocimetry (PIV) [21-30], and fluorescence recovery after

photobleaching (FRAP) [31-37]. Figure 1.2 shows the principle of PIV technique, in

which the tracer particles are followed and their velocity is statistically calculated

based on the grabbed images at quite small time intervals. The slip magnitude can

then be obtained directly from the velocity distribution. PIV provides a direct

measure of liquid flow. Figure 1.3 illustrates the principle of FRAP. An evanescent

wave illustrates the studied liquid, which contains a quite low concentration of

fluorescent molecules. The fluorescent molecules are photobleached by using a high

intensity laser beam and then the recovery of the fluorescent to the targeted region is

monitored due to the influx of the new fluorescent molecules. FRAP is a direct

method since the pattern of the liquid in a designated region can be followed. For

indirect slip measurements, parameters related to boundary conditions are measured

4
and compared with those calculated under the no-slip boundary condition. The slip

magnitude is obtained based on the difference of the two. These techniques include

atomic force microscope (AFM), surface force apparatus (SFA) and capillary

method. SFA was applied by many groups [38-47] for boundary slippage research.

By measuring the drainage force between two approaching surfaces, the slip length

can be obtained based on the Vinogradova slip model [48-52]. AFM is widely

applied in a similar manner as SFA to test the drainage force between two surfaces

during approaching process [52-65]. Compared with SFA, AFM can be used with

different substrates, even including rough surfaces [52]. The slip length is also

calculated from the Vinogradova slip model [48-52] based on the measured drainage

force. Capillary technique is another method for boundary slippage measurement.

Figure 1.4 shows the principle of capillary technique, which is adopted widely for

boundary slip research [66-76]. By measuring the flow rate and comparing it with

the predicted value considering boundary slip, the slip length can be obtained.

Capillary technique is relatively simple, and the requirement for the equipment is

low. However, preparing a channel with atomically smooth inner surface and

constant width is quite difficult. Besides, the wettability and other parameters in

inner surface cannot be measured directly.

Fig. 1.2 Schematic illustration of particle image velocimetry (PIV) technique

5
Fig. 1.3 Schematic illustration of fluorescent recovery after photobleaching

Fig. 1.4 Illustration of capillary technique

1.2.2.2 Parameters governing boundary slippage

Many factors have been found affecting the onset and the magnitude of boundary

slippage. The main factors include wettability, surface roughness, viscosity of the

fluid, shear rate, type of the fluid molecule (polar or non-polar) and nanobubble or

gaseous films at solid/liquid interface. However, there is almost no consistent

conclusion about the parametric effect on boundary slip.

Normally, a large contact angle is indicative of a weak adhesive force between the

liquid and the solid, which can easily be overcome leading to the appearance of

boundary slip and easing the liquid movement on the surface. In the last decade,

evidences for this argument have been provided through simulations and

experiments. Barrat and Bocquet [77] carried out molecular dynamics simulations

6
(MDS) and found that the magnitude of boundary slip increases with the contact

angle, and the slip length can reach 30 molecular diameters for a contact angle of

140°. Huang et al. [78] proved with MDS of a realistic water model that contact

angle is the crucial parameter for controlling water slippage. They also obtained a

curve describing the relationship between slip length and contact angle for

hydrophobic surfaces. Baudry et al. [41] measured the drainage force between a

sphere and a plane, which can be hydrophobic or hydrophilic, immerged in glycerol

using a surface force apparatus (SFA). They reported that the slippage only existed

under non-wetting conditions. Tretheway and Meinhart [21] employed the particle

image velocimetry method to identify the occurrence of boundary slip. They

reported that there was an apparent slip velocity when water flowing through a

hydrophobic micro-channel, but not a hydrophilic channel. Cottin-Bizonne et al. [45]

measured the interaction forces in water and dodecane with a dynamic surface force

apparatus. No boundary slip was detected in all wetting conditions but slip was

sensed for water moving on strong hydrophobic surfaces. They also observed that

water slippage increases with hydrophobicity. Zhu and Granick [40] measured the

slip length in an aqueous Newtonian liquid using SFA. They reported that the slip

length increases with contact angle when the flow rate exceeds a critical level.

Bonaccurso et al. [54, 79] obtained various hydrodynamic drainage forces for water

flow on a fully wetted system and they claimed that it was a proof of water slippage

on a fully wetted surface. Joseph and Tabeling [80] observed the velocity profile of

water flow on different surfaces and concluded that the slip length of water flowing

on a hydrophobic or a hydrophilic surface is about the same. Similar results were

obtained by Henry et al. [81]. They compared the drainage forces that were obtained

with solid surfaces partially covered by surfactant molecules and found that there

7
was no difference in slip length for hydrophobic and hydrophilic surfaces. More

interestingly, Cho et al. [82] measured the hydrodynamic force in different liquids

with an alkylsilane coated surface and a colloidal probe AFM. They found that the

slip length decreased with contact angle for non-polar liquids when the magnitude of

contact angle is relatively small (from about 10°to 40°). However, for polar liquids

with relatively large contact angles (from about 60°to 100°), no correlation between

the slip length and the contact angle was found.

Intuitively, the boundary slip should be affected by surface roughness [79, 83].

However, it is hard to draw a simple conclusion for discrepancies of measured slip

behavior from the published outcomes. Both the nanoscale and microscale roughness

effect have been studied. Hao et al. [84], Hetsroni et al. [85], Kandlikar el al. [86],

McHale and Newton [87], Pit et al. [32], Zhu and Granick [43], Hervet and Leger

[88, 89], Koplik et al. [90] and Jabbarzadeh et al. [91] concluded that increasing the

surface roughness can inhibit boundary slippage from their experiments or MDS.

However, Bonaccurso et al. [79], Cottin-Bizonne et al. [92, 93] and Ligrani et al. [94]

drew an opposite conclusion that the degree of slip increases with the solid

roughness. By specially depositing nanoparticles on a sapphire plate, Schmatko et al.

[95] found that boundary slip decreases with increasing the height of corrugation,

while increasing their width enhanced boundary slip. Galea and Attard [96], and

Priezjev et al. [97] pointed out that there is no boundary slippage if the magnitude of

the roughness is at the same order as the diameter of the fluid molecules. Otherwise,

boundary slippage occurs at the solid/liquid interface. Besides, Ou et al. [74], Choi

and Kim [98] proved that the boundary slip can be strongly improved by a special

surface pattern design with super-hydrophobicity.

8
There are also no consistent conclusions about viscosity effect on boundary slippage.

Craig et al. [53] proved that the slip length increases with fluid viscosity and shear

rate based on the measurement of drainage force using an AFM. However, Cho et al.

[82] noted that there is no relation between slip length and viscosity for polar fluids.

There exist two main conclusions about the effect of shear rate on boundary slippage.

Some experimental and MDS results proved the correctness of the linear slip length

model, i.e. boundary slippage is independent of shear rate, for example, the

experimental results reported by Pit et al. [32], Cottin-Bizonne et al. [99], Joseph and

Tabeling [80], Campbell et al. [38] and Bonaccurso et al. [54]. On the other hand,

opposite conclusion was also reported, i.e. slip length is not a constant. Zhu and

Granick [40] and Craig et al. [53] found that the slip length increases with shear rate

and there is no slip if the shear rate is less than a yield value. Their conclusions were

confirmed by MDS results, such as Thompson and Troian [100], Priezjev and Troian

[101]. Based on their MDS results, the slip length is a constant when the shear rate is

small and increases nonlinearly once the shear rate exceeds a yield value.

Apparently, there is no consistent conclusion about the parametric effect on

boundary slippage. One important reason is that the correlation may be system-

dependent. The second reason is that the parametric coupling effect cannot be

ignored in some experiments. Furthermore, even some correlations between one

parameter and boundary slippage are confirmed, there are only valid under some

special working conditions.

9
1.3 Research on boundary slip in lubrication

The study of boundary slippage in lubrication can date back to the middle of the 20th

century. That the increase in the traction force is barred at a certain sliding speed in a

lubricated non-conformal contact, i.e. elastohydrodynamic lubrication (EHL), was

revealed experimentally, for example by Smith [102], and Johnson and Tevaarwerk

[103]. A limiting shear stress was proposed to account for this limiting traction force,

and lubricant boundary slippage was supposed to occur given that shear stress at the

interface reaches its limiting value. Boundary slip was also proposed as one of the

possible explanation to the mechanism of friction reduction in lubricant additives

[104, 105]. Chappuis [106] proposed a non-wetting bearing based on the boundary

slip phenomenon. Following these studies, there was little related studies appeared

until 1990s, when more approaches for modifying interface affinity became available

due to the advances in surface science. Choo et al. [107, 108] measured the friction

of hydrodynamic lubricated non-conformal contacts under light loads. The lubricants

used were n-hexadecane and glycerol. A roller was run against a stationary disc. The

surface of the disc specimen was smooth and treated with lyopholic or lyophilic

coatings. They reported that lyophobic surfaces provided much smaller friction

compared with lyophilic surfaces. This phenomenon was attributed to boundary slip.

Bongaerts et al. [109] found in the tests with a PDMS ball on a PDMS disk that

hydrophobicity did not affect the friction coefficient in the elastohydrodynamic

lubrication (EHL) regime. However, in the boundary lubrication regime, the friction

coefficient decreased with decreasing contact angle.

Most of the work on boundary slip in lubrication was based on the measurements of

the hydrodynamic or drainage force. Guo et al. [110] proposed to compare the

10
lubricating film formation capacity of different surfaces in order to differentiate the

surface effects on lubrication. They measured the film thickness using interferometry

which is able to measure thickness in micron and sub-micron scale with high

repeatability and accuracy. Their results showed that surfaces of stronger wettability

and smaller contact angle would result in greater film thickness due to a large

strength of intermolecular attraction between the liquid and the solid.

In theory, Spikes [4] incorporated the boundary yield stress model into

hydrodynamic lubrication and derived the modified Reynolds equation. Based on the

new equation, he analyzed the lubrication behavior under different boundary yield

stress for an one-dimensional case. After that, Ma et al. [111] solved a two-

dimensional case by using a multi-linearity algorithm. Later, Guo and Wong [112],

Fortier and Salant [113], and Wu et al. [114] proposed the concept of heterogeneous

slip which means the slip behavior is different at different locations on the sliding

surface. Through the design of slip area, the bearing with higher load capacity and

lower friction coefficient can be obtained.

1.4 Aims and objectives

Boundary slip has proved to be an effective way in altering lubrication behavior,

especially in reducing friction coefficient. The main work of this project is focusing

on the boundary slippage in hydrodynamic lubrication. Optical interferometry is

used for lubricant film thickness measurements. Therefore, the general lubricant film

thickness is in the range of 0.2 μm to 10 μm. The load range in this study is less than

0.5 MPa for ignoring the deformation effect. Besides, all experiments were

11
conducted in ambient environment, and all the lubricants used in this study are

Newtonian for simplifying the analysis.

Inconsistent conclusions were drawn from the aforementioned studies on the effect

of wettability, lubricant’s viscosity and shear rate on friction and slippage, which

means a complete understanding of the correlation between these parameters and

boundary slippage is still under explored. To make good use of slip in directing

lubrication system design, it is necessary to evaluate the relationship among these

parameters, slippage, and lubrication behavior. One of the aims of this research is to

find out how these parameters in determining the boundary slip in hydrodynamic

lubrication by carrying out experimental investigations.

In the present project, an optical lubricated slider test rig with conformal contact

configuration is adopted for its simple contact geometry. An inclined slider geometry

is mathematically well-defined and has been widely used in lubrication analyses.

Furthermore, the magnitude of induced pressure is not significant for a conformal

contact such that pressure effect can be ignored. An optical slider test rig was firstly

designed and commissioned by Guo et al. [115], and its feasibility and credibility

were confirmed [115]. The film thickness can be obtained by monochromatic

interferometry. However, the old version test rig is limited to the measurement of

film thickness, but not friction. Hence, the optical slider test rig is further developed,

such that it can measure the lubricant film thickness and friction force

simultaneously for the current study.

12
To find the effect of boundary slip on lubrication behavior, it is necessary to solve

the modified Reynolds equation considering boundary slip model. Right now, many

slip models have been derived and introduced into hydrodynamic lubrication. Three

models, linear slip length model, boundary yield stress model and non-linear slip

length model, are widely accepted and applied in physics and engineering. Generally,

slip length model (especially linear slip length model) is popular in physics for its

simple and direct description of the slip phenomena. From the engineering

perspective, the boundary yield stress model is more acceptable because a yield or

limiting shear stress is believed existing at the solid/liquid interface [116, 117].

Spikes [4] derived the modified Reynolds equation considering the boundary yield

stress at the stationary solid surface and obtained the solution of the one-dimensional

modified equation. Ma et al. [111] studied the two-dimensional case by using the

finite element method. The two-dimensional modified Reynolds equation is not easy

to solve by applying the widely used finite difference method because the slip

control equation is cyclic and there are infinite possibilities for the slip direction.

Tauviqirrahman et al. [118] had a try by simplifying the boundary yield stress as a

scalar instead of a vector. Therefore, the present study is first started with the

solution of the modified Reynolds equation, and the effect of boundary slip on

lubrication behavior under different working conditions is then studied.

In physics, the potential energy barrier at a liquid and solid interface, which can be

calculated from the contact angle and the contact angle hysteresis, is derived for

describing the force needed in separating liquid from solid surface. Therefore, there

should be correlation between boundary yield stress and potential energy barrier

because they describe similar phenomena. In order to control the boundary slippage

13
quantitatively by wetting design, it is necessary to find out the relationship between

the two parameters.

The main objectives of the project can be summarized as:

1. To improve the existing slider test rig to realize the measurement of lubricant

film thickness and friction simultaneously.

2. To find out the effect of boundary slip on lubrication behavior under different

working conditions in theory by solving the modified Reynolds equation

considering boundary yield stress.

3. To experimentally investigate the correlation between interface wettability and

lubrication behavior.

4. To describe the relationship between boundary yield stress and potential energy

barrier.

5. To study the effect of viscosity and sliding speed on boundary slippage in thin

film hydrodynamic lubrication.

1.5 Chapter outline

The detailed information about this research is introduced in the following chapters,

including the development of the test rig, the solution of modified Reynolds equation,

identification of the best wetting parameter correlated with hydrodynamic lubrication,

confirmation of the correlation between boundary yield stress and potential energy

barrier, detecting the effect of viscosity of lubricants and sliding speed on boundary

slip and future work.

14
In Chapter 2, the existing set-up and its running conditions are described firstly.

Then, the design for the measurement of friction force is discussed. Finally, the

dichromatic interferometry is introduced for the on-line measurement of film

thickness.

The solution of modified Reynolds equation is introduced in Chapter 3. Slippage

models and the derivation of the modified Reynolds equation are shown firstly. Then,

the scalar method adopted in this research is verified by comparing the calculations

with the published results.

In Chapter 4, the definition and measurement of contact angle, contact angle

hysteresis and spreading parameter are given, respectively. The correlation of these

parameters and lubricant film thickness is shown. The explanation for the

experimental results is also discussed.

The experiment to confirm the relationship between boundary yield stress and

potential energy barrier is introduced in Chapter 5. The details of the experimental

procedure and silicon oil-based lubricant preparation are illustrated. The

experimental results are also discussed in the end.

Chapter 6 describes the effect of viscosity of lubricant and sliding speed on boundary

slippage. The findings are compared with a new slip model proposed recently.

In Chapter 7, the main results in this research are concluded and suggestions about

the future work are given briefly.

15
Chapter 2

Development of Test Rig for Thin Film Hydrodynamic Lubrication

2.1 Required test rig for the project

Typical slip measuring techniques (SFA, AFM, PIV) have been introduced in

Chapter 1. However, almost all mentioned techniques are not valid for lubrication

research since the fluid film generated in the mentioned techniques does not bear any

load capacity. To study the boundary slip effect on lubrication, friction is one of the

key indicative parameters. The principle is originated from the concept that friction

force would drop if slip appears at the solid/lubricant interface. Based on this

concept, Choo et al. [107, 108] measured the friction of hydrodynamic lubricated

non-conformal contacts under light loads and found that lyophobic surfaces generate

much smaller friction than lyophilic surfaces. Using the similar method, Bongaerts et

al. [109] measured the friction force with a PDMS ball on a PDMS disk and reported

that hydrophobicity did not affect the friction coefficient in the elastohydrodynamic

lubrication (EHL) regime. However, in the boundary lubrication regime, the friction

coefficient decreased with decreasing contact angle. Another key parameter for

detecting slip effect is lubricant film thickness. It is because the change of boundary

condition leads to the change of load carrying capacity. In the present study, it is

desirable to obtain friction force and lubricant film thickness simultaneously because

they provide more information about the boundary slippage effect. In this chapter,

development of an existing optical slider test rig for realizing the simultaneous

measurement of film thickness and friction force is introduced.

16
2.2 Existing set up

In order to study the effect of boundary slippage on hydrodynamic lubrication, an

apparatus which can simulate the lubrication of conformal contact type is used. The

main body of the existing test rig is shown in Fig. 2.1, and this test rig was

developed by our research group in 2010 [115]. The test rig is composed of a slider

holder and adjustment part (1), a rotation part (2), an optical interference part (3), a

base frame (4), a driving part (5), a load part (6) and (8), and a friction measurement

part (7). Figures 2.2 and 2.3 show the front elevation and the top view of the system

respectively.

Fig. 2.1 Main body of the existing set up

(1-slider holder, 2-disc rotation part, 3-optical interference part, 4-base frame, 5-

motor, 6-load arm, 7-friction sensor, 8-load handle)

Figure 2.4 presents the design of the sliding contact. The sliding contact is formed

between a rotating glass disc (1) and a stationary tilted slider (2). The inclination

angle between the slider and the disc can be adjusted by eight bolts (6) located on the

load arm (5). The slider is fixed by a slider holder (3) which is connected with the

17
load arm by a universal joint (4). The lubricant film separates the slider from the

glass disc for hydrodynamic effect once the glass disc is driven by a timing pulley.

Fig. 2.2 Front elevation of the set up

Fig. 2.3 Top view of the set up

18
Fig. 2.4 Schematic illustration of the sliding contact

(1-glass disc, 2-slider, 3-slider holder, 4-universal joint, 5-load arm, 6-adjusting bolt)

2.3 Friction measurement design

In general, friction in fluid film lubrication is small and the measuring sensor must

be sensitive enough. However, the usual scheme employed in force measurement by

using a slim beam with strain gages cannot be employed here, because the rigidity of

the slider holder will be destroyed. An alternative was adopted as shown in Fig. 2.5.

The load lever has two degrees of freedom. The friction force is inferred from the

measurement taken by the force sensor (Beijing Zhangkai Instruments Co. Ltd). Its

measuring range is up to 10 N. Obviously, the design for friction force measurement

is accurate if there is no internal friction in the bearing which connects the test rig

table and load arm. To ignore the effect of internal force in the connecting bearing, it

is necessary to do calibration before each experiment. Figure 2.6 shows

schematically the calibration design for the friction force measurement. Two force

sensors are connected to the load arm symmetrically. In the calibration, sensor one is

static and sensor two moves with a quite small speed in the disc rotation direction.

The difference between the measured forces from the two sensors is regarded as the

internal force of the bearing. After the calibration, sensor two is disconnected and

only sensor one is left for the friction force measurement, as shown in Fig. 2.5. The

19
true friction force is the difference between the value detected by sensor and the

calibrated internal force.

Figure 2.7 shows the measured friction forces at different speeds for two loads. In

the study, two sliders were adopted for identifying the surface effect on

hydrodynamic lubrication. One is an untreated steel slider and the other one is coated

with a type of hydrophobic coating (EGC). The roughness, Ra, of the steel sliders

with no coating and EGC is 11.8 nm and 44 nm, respectively. For the steel slider, the

friction increases monotonously with speed and the friction coefficient reduces with

the increase in normal load. These phenomena are correlated well with the classical

hydrodynamic lubrication theory. However the trend of the friction-speed curve for

the EGC slider is quite different. The friction coefficient is very large in magnitude

under low speeds because the film thickness is about the same magnitude as the

roughness of EGC coated surface (the measured film thickness will be shown and

discussed in Chapter 3). However, it drops quickly with the increase in disc speed.

The great reduction of the friction at the beginning of the test with the EGC coated

slider by increasing the speed correlates indeed well with the Stribeck curve. The

high friction at the beginning is due to the interaction between the asperities of the

two sliding surfaces. Following the increases in the film thickness with speed, the

amount of asperity interaction becomes less and the friction is thus reduced. For

further increase in speed, the friction curve approaches to a steady value. Under high

speeds, the settling friction coefficients of the EGC coated slider can even be lower

than those of steel slider with no coating, as shown in Fig. 2.7(b).

20
Fig. 2.5 Schematic of friction measurement

Fig. 2.6 Schematic of calibration design

0.025 0.020
Friction coefficient

0.020 EGC
Friction coefficient

0.015 EGC
Steel Steel
0.015 Load: 4 N
0.010 Load: 8 N
0.010
0.005
0.005

0.000 0.000
0 5 10 15 20 0 5 10 15 20
Speed (mm/s) Speed (mm/s)

(a) (b)

Fig. 2.7 Change of measured friction coefficient with speed under two different load

21
2.4 On-line optical film thickness measurement design

2.4.1 Optical interferometry for lubrication research

The existing optical slider test rig adopts monochromatic interferometry to determine

the lubricant film thickness. The film thickness can be obtained by tracing the

intensity change at an arbitrary spot selected in the interference image of the contact

during the increase or decrease of the lubricating film thickness in the accelerating or

decelerating processes, respectively. Optical interferometry has been widely used for

measuring film thickness in lubrication. Its concept is illustrated with an optical

lubrication system in Fig. 2.8. A lubrication film is bound between a steel surface

and a chromium-coated glass disc. A light beam is projected onto the contact

through the glass disc. The resultant interferogram is formed by the interference of

the first two major reflections and with others of higher order reflections. Gohar and

Cameron [119, 120] were the first to utilize optical interferometry successfully in

lubrication research and presented the ever first optical interferogram of a typical

elastohydrodynamic lubricated (EHL) contact with a unique horseshoe-shaped film.

Their experimental set up used the concept shown in Fig. 2.8 except that the slider

was replaced by a steel ball. A major limitation of the optical technique is its

requirement that one or both surfaces be transparent, which differs from the case in

typical engineering steel-on-steel contact. However, optical interferometry provides

far more detailed information on the shape of lubricating films compared with other

methods, such as X-ray and ultrasound, which provide only the average film

thickness. Hence, optical interferometry has become the most widely used technique

in experimental EHL research since its applicability to lubrication studies was

proved by Cameron and his group at Imperial College in the 1960s. In the early work

of Cameron et al. [121], different light sources, including monochromatic,

22
dichromatic (use of two beams of different wavelengths) and chromatic (white light)

were evaluated. Monochromatic interferometry can measure a very thick film

(readily up to a few microns using a laser and theoretically to several tens of microns

for systems with high optical quality). Chromatic interferometry provides higher

resolution than monochromatic interferometry but a shorter measuring range of up to

about 1 µm. Dichromatic interferometry combines the advantages of these two

former methods.

Fig. 2.8 Schematic illustration of an optical lubrication system

In the optical EHL work of Cameron et al. [119-121], a monochromatic light source

was mostly used. Dark interference fringes were used to calculate film thickness

because of their high definition. The magnitude of optical film thickness at the

location of dark fringes is equal to a multiple of half the wavelength of the light

source. Hence, the minimum measurable film thickness is restricted to the

interference fringe of the first order. The fringe intensity method which uses the

intensity of interference to calculate the film thickness was proposed by Roberts and

Tabor in the 1970s [122]. The intensity method significantly enlarges the resolution

of an interferometry measurement. For a simple interference system with only the

two major reflections as shown in Fig. 2.8, the variation in interference intensity

23
with film thickness (I-h) follows a cosine function. However, in a conventional

optical EHL set up, the variation in interference intensity with film thickness

deviates from the ideal cosine form because of the complex optical properties of the

Cr layer and steel surface, as well as the effect of multi-reflections. The intensity

method was also adopted by Luo et al. [123] in the 1990s. They tuned carefully the

optical test rig to acquire a cosine variation of intensity with film thickness.

Subsequently, Guo and Wong [124, 125] completed a full analysis of the optical

characteristics of a conventional optical EHL set up with multi-beam interference.

The complex optical properties of all materials involved and the multi-beam

reflectance of the system were evaluated. At the end of their study, the intensity

variation with film thickness was mathematically modeled. Therefore, film thickness

can be directly interpreted from interference intensities and is not restricted to the

points of dark fringes.

2.4.2 Special requirements of optical methods for the project

Optical interferometry has recently been applied to the study of hydrodynamic

lubrication by our research groups. For the present project, the key measuring

parameters of the sliding tester include the lubricating film thickness and friction

force. Film thickness can be extracted from any of the interferograms of the

lubricated contact based on the known fringe orders, which can be obtained by

tracing the intensity changes at any spot in the interference image of the contact

during the increase or decrease of the lubricating film thickness in the acceleration or

deceleration processes, respectively. However, following an exact spot requires an

extremely stable slider contact, which is unachievable when the slider is connected

to a load cell for friction measurement. During the acceleration or deceleration

24
processes, the contact area shifts because the tension of the connecting wire to the

load cell changes. Furthermore, determining changes in the fringe order by stopping

the running disc cannot be implemented when investigating dynamic or transient

lubrication behaviors, such as rapid changes in the film thickness caused by shock.

Moreover, intensity data may be lost during measurement if the lubricant film is

rapidly formed and varying intensity rate exceeds the sampling rate of image

acquisition. Thus, film thickness must be directly measured from a single

interferogram of the contact without stopping the test.

Dichromatic interferometry using two light sources with different colors (two

wavelengths) is widely used in applied optics to evaluate aspheric surfaces for the

wide range of sensitivities of the technique [126-129]. In a typical set up, two lasers

are used successively to form interferograms with the same area of interest. Wyant

[129] showed that by combining the two interferograms formed by lasers of different

wavelengths, the resultant interferogram can be taken as the result of using a pseudo-

light source with longer wavelength. This method was mainly developed for static

measurements, i.e., component interferograms are obtained successively with

different lasers. The concept of two-wavelength interferometry was extended by

Polhemus [130] to real-time dynamic testing by simultaneously operating two

sources with different wavelengths. The measuring range of two-wavelength

interferometry was further enhanced from one or two microns to a few tens of

microns by phase-shifting [131], improving data analyses [132, 133] and, more

recently, signal processing based on arithmetic properties [134].

25
In order to fulfill the special requirements of optical method of this project, a

dichromatic interferometry technique for on-line measuring hydrodynamic

lubrication film thickness is developed. The typical thickness of the thin film

hydrodynamic lubrication regime is approximately two microns. Hence, the original

two-wavelength interferometry proposed by Wyant [129] is adopted in the current

study without implementing complex data processing or analysis algorithms.

2.4.3 Principle of dichromatic optical interference technique

Methods for conventional optical measurement of lubricating film thickness are

based on the periodic characteristics of the relationship between the interference

intensity I and film thickness h. The relationship is a cosine function if the

interference is formed by the two major reflections [123], as illustrated in Fig. 2.8, or

a skewed cosine curve if all reflectances are considered (referred to as multi-beam

intensity interference in [124]). Differences in film thickness between two points

with same intensity and phase can be expressed as,

𝜆
∆ℎ = ∆𝑁 (2.1)
2𝑛

where h is the film thickness, N is the fringe order, n is the refractive index of the

lubricant, and 𝜆 is the wavelength of the light. When considering only one period as

Δ𝑁=1, the cycle of the interference intensity variation with film thickness is,

𝜆
𝑇= (2.2)
2𝑛

The two beams with different colors (λ1 and λ2) simultaneously form interference

images. Superimposing the two intensity curves against the film thickness, i.e.

(𝐼1 − 𝐼2 ) vs. h, along the same line of measurement provides a resultant periodic

curve with an envelope cycle (see Appendix A),

26
𝜆1 𝜆2
𝑇𝑒 = (2.3)
2𝑛|𝜆1 − 𝜆2 |

where 𝑇𝑒 is the envelope cycle. The envelope is equivalent to the curve of the

intensity and film thickness variation with a longer wavelength 𝜆𝑒 [129],

𝜆1 𝜆2
𝜆𝑒 = (2.4)
|𝜆1 − 𝜆2 |

Obviously, the cycle of the envelope is much greater than that of I vs. h from a light

beam of a single color. The film thickness in the first cycle of the intensity and film

thickness variation (within the first-fringe order, N = 1) can be deduced directly from

the measured intensity. Hence, a longer wavelength results in a broader range for

film thickness measurements. Thus, dichromatic interferometry has a broader

measuring range because of its long equivalent wavelength of the envelope. The

technique is used mainly because of the slower moving rate of the envelope

corresponding to the change in film thickness compared with that of fringes with

monochromatic interference. The ratio between the two moving rates is equal to

(Appendix B),

𝑣𝑒𝑛𝑣𝑒𝑙𝑜𝑝𝑒 |𝜆1 − 𝜆2 |
= (2.5)
𝑣𝜆1 𝜆2

Therefore, the order of the envelope cycle can be easily identified. For example, if λ2

is 638 nm (red) and λ1 is 532 nm (green), the speed ratio is 0.199. Hence, the

envelope cycle moves approximately one-fifth slower.

2.4.4 Implementation

A schematic of the process of film thickness measurement and the new technique are

shown in Fig. 2.9. The sliding contact is constructed with a rotating glass disc and a

stationary slider, in which the inclination angle can be fixed and adjusted using eight

27
adjustment bolts located on the load arm. The surface of the glass disc is partially

coated with a Cr layer of 20 nm in thickness for partial reflection, and a transparent

SiO2 layer of about 200 nm is added on top for protection. The film thickness at the

exit was measured in the tests. The change in film thickness at any location on the

slider surface is equal to the change in film thickness at the exit when a slider with

fixed inclination stops. Lasers of green and red colors (wavelength: 532 and 638 nm,

respectively) were selected as light sources. A three-CCD camera with higher

accuracy than a colored-CCD camera was used to obtain the interference intensity

using a Bayer array. The roughness of the steel slider was 11.8 nm, and its breadth

(in sliding direction) and length (transverse to sliding direction) were 4 and 9 mm,

respectively. PAO400/40 was used as the lubricant, the properties of which are listed

in Table 2.1. The tests were carried out in a controlled environment (ambient

temperature: 21 ±1 oC, humidity: 33% ±1%).

Fig. 2.9 Schematic illustration of the present set up using dichromatic interferometry

measurement

28
Table 2.1 Properties of PAO400/40 used in the tests

Viscosity (mPa s) @21 oC Refractive index

879.8 1.47

To implement the new technique, a dichromatic interference image with known film

thickness variation was needed as a standard. A standard image was captured using a

wedge gap with a fixed inclination. The steel slider was tilted and rested on the glass

disc. The wedge gap was filled with the lubricant PAO400/40. The lower edge of the

slider was in contact with the glass disc surface, i.e. zero film thickness. The selected

standard image is shown in Fig. 2.10. The intensity variation of the two different

color beams (red and green) along a line (as highlighted in Fig. 2.10) was normalized

to their adjacent local maximum and minimum intensities. The difference of the two

normalized intensity curves (𝐼𝑟𝑒𝑑 − 𝐼𝑔𝑟𝑒𝑒𝑛 ) is plotted in Fig. 2.11. The total number

of pixels for the full length of the slider (4 mm) was counted to be 751. The

magnification of the image was thus 5.326 µm/pixel. The inclined angle of the slider

can be readily obtained from the interferogram of the red or green beam. From the

interferogram of the green beam (λ = 532 nm), the cycle was measured to be 50.67

pixel on average (or 269.87 µm in length). Hence, the inclination was calculated to

be 6.705 × 10-4 radians. The envelope of the normalized intensity difference curve

was obtained using Hilbert transform [135], as shown in Fig. 2.11. The envelope

obtained using Hilbert transform was not smooth, but provided the location of the

first and second valleys, which is important for automatizing the measurement

process in subsequent development. The curves of summation and subtraction of the

two intensities, (𝐼1 + 𝐼2 ) and (𝐼1 − 𝐼2 ), are plotted in Fig. 2.12. The two valleys of the

curve of (𝐼1 − 𝐼2 ) were represented by the 227th and 529th pixels. These valleys

29
correspond to the locations of the maximum intensities of (𝐼1 + 𝐼2 ), which are the

bright yellow fringes shown in Fig. 2.10. The bright yellow fringe is the result of the

sum of the constructive interferences of the two primary colors. Hence, the film

thickness of the first and second yellow fringes can be deduced from the known

angle of inclination and are listed in Table 2.2 as a reference. Thus, the cycle of the

envelope was determined as 1078.5 nm.

Fig. 2.10 Interferogram used as the standard (YF: yellow fringe)

3.0

1.5
Intensity difference

0.0

-1.5

-3.0
0 200 400 600
Pixel

Fig. 2.11 Analysis results of the standard interferogram

30
3.0
Summation
Subtraction

Normalized intensity
1.5

0.0

-1.5

-3.0
0 200 400 600
Pixel

Fig. 2.12 Variation of summation (𝐼1 + 𝐼2 ) and subtraction (𝐼1 − 𝐼2 ) of intensities

Table 2.2 Calibrated film thickness of the yellow fringes

Yellow fringe order Film thickness (nm)

1st 810.6

2nd 1889.1

The effectiveness of dichromatic interferometry for measuring lubricating film

thickness was evaluated under static and running conditions, and the results were

compared with those obtained using conventional monochromatic interferometry.

2.4.5 Validation for on-line film thickness measurement

Dichromatic interferometry was applied to measure the inclination of static wedge

gaps and demonstrate the reliability of the current technique. Considering that

monochromatic interferometry can provide accurate results, the inclinations obtained

using the new technique were compared with those obtained using monochromatic

interferometry. Dichromatic interference images with different inclinations under

31
conditions similar to those of the standard image were captured and analyzed, as

shown in Fig. 2.13. The location of the first yellow fringe was detected. Given the

film thickness represented by the first yellow fringe provided in Table 2.2, the angle

of inclination was then calculated. The calculated inclinations were compared with

the results detected using monochromatic interferometry and are shown in Table 2.3.

The table shows that the inclinations obtained through monochromatic

interferometry for cases #1 and #2 are 5.381 × 10-4 and 4.208 × 10-4 radians,

respectively. The inclinations calculated using dichromatic interferometry were

5.378 × 10-4 and 4.205 × 10-4 radians. Differences between the two measured results

are lower than 0.1 %. Thus, the accuracy of dichromatic interferometry fulfills

requirements for film thickness measurement.

Table 2.3 Inclined angle measurement using dichromatic and monochromatic

interferometry

Cases Pixel number of first- Angle of inclination (×10-4 radian)

order yellow fringe Dichromatic Monochromatic

#1 283 5.378 5.381

#2 362 4.205 4.208

32
3.0 3.0
Intensity difference

Intensity difference
1.5 1.5

0.0 0.0

-1.5 -1.5

-3.0 -3.0
0 200 400 600 0 200 400 600
Pixel Pixel

(a) Angle I (b) Angle II

Fig. 2.13 Interferogram and analysis results of the static slider contact with different

inclinations (“Arrow” indicating the yellow fringes)

Dichromatic interferometry was applied to determine the film thickness under a

running lubricating slider contact. During dynamic film thickness measurement, a

constant inclined angle of the slider (1:2064) was adopted and the film thickness was

measured under different speeds. Each measurement was repeated for three times.

Monochromatic measurement was also performed, and the results obtained using the

two techniques were compared. Figure 2.14 shows the dichromatic interference

images formed under different speeds with a constant load of 10 N. Changes in the

position of the second order wide yellow fringe with different speeds are shown.

Figure 2.15 shows a log-log graph of the film thickness against speed. The film

thickness was calculated using the two techniques. Generally, the two curves

coincided, especially under low speeds. All data points which are average of three

33
separate measurements are shown with uncertainties in Table 2.4. The maximum

difference in the film thickness for the specified speed range using the two methods

was approximately 40 nm at a speed of 2 mm/s. This finding indicates that the

maximum relative error (or accuracy) for dichromatic interference is lower than

2.5%, as shown in Table 2.4. The error can be attributed to the fluctuation of the disc

during the tests, particularly at high speeds. The repeatability of the experiments was

very good as reflected by the small uncertainty of each measuring points listed in

Table 2.4. Monochromatic interferometry requires stopping the experiment to

monitor the drop of the slider and measure the film thickness. By contrast,

dichromatic interference requires only one interference image of the lubricant film to

calculate the film thickness. Therefore, dynamic measurements without the need to

stop the disc are realized. The envelope was formed by combining the two

interference fringes to provide an equivalent intensity and film thickness variation in

a larger cycle. Equation (2.3) indicates that the cycle of the envelope is related to the

wavelength of the two beams. The large product of the two wavelengths or their

small differences results in larger envelope cycle. Wyant [129] theoretically obtained

an equivalent wavelength of 28.5 µm using dichromatic light beams with

wavelengths of 496.5 and 488 nm.

34
(a) 0.25 mm/s

(b) 0.5 mm/s

(c) 1.0 mm/s

(d) 1.5 mm/s

(e) 2.0 mm/s

Fig. 2.14 Interference images under different speeds

(Position marked represents the second-order yellow fringe.)

Real-time measurement of the thickness of hydrodynamic lubricating film requires

the two interferograms formed by two beams to be simultaneously obtained.

Therefore, the selected wavelengths must correspond to the response spectrum of the

camera used, which limits the envelope cycle. In the current experiment, lasers with

35
wavelengths of 532 and 638 nm were used. The theoretical envelope cycle was 1.089

µm based on Eqn. (2.3). This result is correlated with the measured cycle of 1.079

µm. However, the lasers can be optimized to obtain an envelope with a larger cycle.

For example, lasers with wavelengths of 550 and 600 nm may be selected according

to the response spectrum of the camera, and the envelope cycle can be theoretically

extended to 2.245 µm under similar working conditions.

2.0

1.5
Film thickness (m)

1.0

Monochromatic
Dichromatic
0.5

0.5 1.0 1.5 2.0 2.5


Speed (mm/s)

Fig. 2.15 Changes in film thickness with speed

Table 2.4 Film thickness measurements using mono- and dichromatic interferometry

Speed Film thickness (nm) Accuracy

(mm/s) Monochromatic Dichromatic (%)

0.25 556.7+23.5
−29.9 558.4+40.6
−33.6 0.30

0.50 897.5+5.7
−8.9 887.9+21.2
−21.9 1.09

1.00 1284.5+21.9
−14.5 1309.2+19.7
−23.9 1.88

1.50 1560.4+5.9
−8.7 1581.2+8.7
−10.5 1.31

2.00 1760.6+2.4
−2.5 1805.8+32.3
−35.7 2.50

36
Dichromatic interferometry is effective for real-time determination of the film

thickness, and its efficiency is higher than that of monochromatic interferometry.

Counting intensity changes using monochromatic light at an arbitrary spot on the

lubricated contact by intermittently interrupting the test is tedious, time consuming,

and impractical for analyzing volatile or water-adsorbing lubricants, which may

suffer from a significant change in viscosity during a lengthy test. For example, in

the experiment described in Fig. 2.15, the duration of the test using dichromatic was

approximately 10 min without interruption. Using monochromatic light, by contrast,

causes the test to last as long as 50 min.

Dichromatic interference is important for tracing the position of the yellow fringe

with a known fringe order. The frame rate of the recording system was 25 fps. If the

changes in the film thickness between two consecutive frames exceeds one cycle of

the yellow fringes, i.e., the magnitude of the film thickness change in 1⁄25 seconds

is longer than 𝜆𝑒 ⁄𝑛, the measurement suffers from wavelength ambiguity.

Measurement errors include glass disc fluctuation and stability of the wavelengths

used. The band widths of the light sources used in the current study are 2 and 10 nm

for green and red lasers, respectively. The glass disc cannot be guaranteed to be

absolutely flat during dynamic measurement. Currently, the fluctuation range of the

running glass disc can be controlled within 100 nm. This uncertainty will lead to

slight errors in measuring the inclination and location of yellow fringes. Furthermore,

the stability of the used wavelengths determines measurement error. Therefore,

lasers with high stability and small band width are recommended in the current

measurement system.

37
2.5 Summary

The existing optical slider test rig was further developed for realizing the

simultaneous measurement of friction force and lubricant film thickness. The design

for friction force measurement which was achieved by using a load cell of high

sensitivity was introduced and the calibration process was also considered. To

implement simultaneous measurement of friction and film thickness, a new optical

method for measuring the film thickness of hydrodynamic lubricated contact using

dichromatic interferometry was developed. Two light beams with different

wavelengths are required for simultaneous projection on the lubricating contact.

Superimposing the two interference patterns results in an envelope of the intensity-

difference and film thickness variation, that is equivalent to an intensity and film

thickness curve with a longer cycle. Using the equivalent intensity curve, the

measurement range of film thickness can be enhanced compared with

monochromatic interferometry. The moving rate of the equivalent intensity curve

corresponding to a rapid change in film thickness was fairly slow. Hence,

dichromatic interferometry can facilitate real-time determination of lubricating film

thickness and is suitable for investigating transient or dynamic lubricating problems.

38
Chapter 3

Boundary Yield Stress Model in Hydrodynamic Lubrication

3.1 Introduction

As discussed in Chapter 1, the existence of boundary slip in fluid mechanics and

lubrication has been proved by different studies. In theoretical work, a number of

slip models have been proposed to describe the boundary slip, such as linear slip

model [11], boundary yield stress model [102, 136, 137] and non-linear slip length

model [100]. The slip length model is applied widely in physics to describe and

probe the related phenomena. However, the boundary yield stress model is more

preferred in engineering. To advance the design of a hydrodynamic lubricated

system from the perspective of boundary slippage, it is necessary to study the

boundary slip model theoretically. From one hand, the existing phenomena can be

explained from the slip model. For the other, the correct slip model can provide great

assistance in bearing design. In this chapter, the main boundary slip models are

introduced, and their advantages and drawbacks are discussed.

The boundary yield stress model is applied in this study for its reasonable description

of the boundary slip in lubrication. Spikes [4] incorporated the boundary yield stress

model into the Reynolds equation and carried out a simple one-dimensional analysis

on the effect of boundary slip on the hydrodynamic properties of a slider. Later, Ma

et al. [111] conducted a similar but two-dimensional analysis on the hydrodynamics

of a finite slider gap with boundary slip by using a multi-linearity method. In this

chapter, the method used by Tauviqirrahman et al. [118] for solving a two-

dimensional modified Reynolds equation is adopted. Its validity is confirmed by

39
comparing with the results calculated by Ma et al. [111]. Based on the calculation

results, the effect of boudnary yield stress on pressure distribution and load capacity

under different working conditions are discussed.

3.2 Slippage models

3.2.1 Slip length model

Navier [11] proposed the slip length model in 1820s, which describes the slip

velocity to be proportional to the shear rate at the solid/liquid interface, as shown in

Eqn. (1.1). The proportional coefficient b, which is called as “slip length”, is the

fictional distance between the solid surface and the point where the velocity linearly

decreases to zero. The model is illustrated in Fig. 1.1(c). Actually, the linear slip

length model includes a true slip model (Fig. 1.1(c)) and an apparent slip length

model (Fig. 1.1(b)). The apparent slip length model is established based on the

hypothesis that there is a stagnant layer with lower viscosity at the solid/liquid

interface [9]. The slip length in the apparent slip length model can be described as

𝑏 = 𝛿(𝜂𝐵 ⁄𝜂𝑠 − 1) (3.1)

where 𝛿 and 𝜂𝑠 is the thickness and viscosity of the stagnant layer, respectively. 𝜂𝐵

is the viscosity of the bulk liquid. Although the physical essence of true slip and

apparent slip is different, equation (1.1) is adopted to describe the slip length in both

slip models. Therefore, there is no need to distinguish the two models in fluid

mechanics analysis. Slip length model is widely used both in physics and

engineering although there is no mathematical derivation for this model.

3.2.2 Boundary yield stress model

40
Many studies proved that the slip length model is inaccurate. Some experiments

proved that the slip length is not a constant and dependent on shear rate [40, 53].

Besides, it is found that there is a limiting yield stress in controlling the onset of

boundary slip [42, 43, 53, 138]. Based on these cases, boundary yield stress model is

accepted and adopted in fluid mechanics. The limiting shear stress model was

proposed initially for describing the rheology behavior under elastohydrodynamic

lubrication (EHL) conditions [136, 137]. In this model, it is assumed that there exists

a limiting shear stress or boundary yield stress at the solid/liquid interface, and slip

occurs if the shear stress at the solid/liquid interface reaches the limiting value.

Based on the boundary yield stress model, Wu and Ma [139] studied the change of

viscous force due to boundary slip when a ball approaches to a plate. The correlation

of their calculated results with experimental data is much better than those calculated

from the slip length model, as shown in Fig. 3.1. Apparently, the slip length model is

only accurate under low shear rates. However, boundary yield stress model is valid

in quite a large range of shear rate.

Fig. 3.1 Comparison of predicted viscous force with experimental data [139]
41
Spikes [4] derived the modified Reynolds equation considering the boundary yield

stress at the stationary slider surface for both one and two-dimensional cases, i.e.

Eqns. (3.2) and (3.3.).

𝜕 ℎ3 𝜕𝑝 𝜕ℎ 3 𝜕 𝜏𝑏 2
( ) = 3𝑢 + ( ℎ ) (3.2)
𝜕𝑥 𝜂 𝜕𝑥 𝜕𝑥 2 𝜕𝑥 𝜂

𝜕 ℎ3 𝜕𝑝 𝜕 ℎ3 𝜕𝑝
( )+ ( )
𝜕𝑥 𝜂 𝜕𝑥 𝜕𝑦 𝜂 𝜕𝑦
(3.3)
𝜕ℎ 3 𝜕 𝜏𝑥𝑏 2 𝜕 𝜏𝑦𝑏
= 3𝑢 + [ ( ℎ ) + ℎ2 ( )]
𝜕𝑥 2 𝜕𝑥 𝜂 𝜕𝑦 𝜂

For a two-dimensional flow, the boundary yield stress is given by,

𝜏𝑏 = ⃗⃗⃗⃗⃗⃗
⃗⃗⃗ 𝜏𝑥𝑏 + 𝜏⃗⃗⃗⃗⃗⃗
𝑦𝑏 (3.4)

where the subscript x and y denote the boundary yield stress component in x- and y-

direction, respectively. Slippage would occur when the resultant shear stress at the

solid/liquid interface (𝜏 = 𝜏⃗⃗⃗𝑥 + ⃗⃗⃗⃗


𝜏𝑦 ) attains the boundary yield stress, 𝜏𝑏 .

3.2.3 Non-linear slip length model

The non-linear slip length model was proposed by Thompson and Troian [100]

based on their molecular simulation results, and the description of this model is,

𝑏⁄𝑏0 = (1 − 𝛾̇ ⁄𝛾̇𝑐 )−1⁄2 (3.5)

where 𝑏0 is the initial slip length, 𝛾̇ is the shear rate at the solid/liquid interface and

𝛾̇𝑐 is the boundary yield shear rate. This model was confirmed later by Priezjev and

Troian [101] who obtained similar results by simulating the shear flow of polymers.

Figure 3.2 demonstrates the relationship between slip length and shear rate. Clearly,

the slip length can be regarded as a constant under low shear rates, which is

consistent with the slip length model. At high shear rates, the slip length increases

42
rapidly with the shear rate and the non-linear slip length model can be treated as

boundary yield stress model.

Fig. 3.2 Correlation between normalized slip length and normalized shear rate [100]

Compared with the slip length model and the boundary yield stress model, it is not

easy to analyze the slip behavior theoretically using a non-linear slip length model

because there are two unknown parameters, i.e. slip length and boundary yield stress,

in controlling the slip behavior.

Based on the above discussion, the boundary yield stress model is adopted here and

incorporated into the Reynolds equation. The modified Reynolds equation is solved

in this chapter.

3.3 Derivation of modified Reynolds equation

Although Spikes [4] provided the modified Reynolds equation considering boundary

yield stress model, more details about the derivation are listed here. Spikes [4] also

proved that the boundary slip is only allowed to appear on the stationary surface as

to guarantee the entrance of fluid into the bearing. Hence, it is assumed that there is

no boundary slip at the moving surface. Other assumptions are listed below:

43
1. The viscosity and density of the fluid or lubricant are constant.

2. The effect of body force and inertia are negligible.

3. There are no pressure gradients across the film.

4. The lubricant can only slip at the stationary solid surface.

5. The lubricant is Newtonian and there is only laminar flow in the contact area.

Fig. 3.3 Equilibrium of an element

Based on the above assumption, the force balance in x-direction of a small fluid

element with sides dx dy dz (Fig. 3.3) can be expressed as,

𝜕𝜏𝑥𝑧 𝜕𝑝
𝑝𝑑𝑦𝑑𝑧 + (𝜏𝑥𝑧 + 𝑑𝑧) 𝑑𝑥𝑑𝑦 = 𝜏𝑥𝑧 𝑑𝑥𝑑𝑦 + (𝑝 + 𝑑𝑥) 𝑑𝑦𝑑𝑧 (3.6)
𝜕𝑧 𝜕𝑥

Eqn. (3.6) can be simplified and rewritten as,

𝜕𝑝 𝜕𝜏𝑥𝑧
= (3.7)
𝜕𝑥 𝜕𝑧

The expression for the shear stress, 𝜏𝑥𝑧 = 𝜂 𝜕𝑢⁄𝜕𝑧, is substituted into Eqn. (3.7) and

it yields,

𝜕𝑝 𝜕 𝜕𝑢
= (𝜂 ) (3.8)
𝜕𝑥 𝜕𝑧 𝜕𝑧

In the current study, it is assumed that the boundary yield stress at the upper

stationary solid surface is 𝜏𝑥𝑏 and 𝜏𝑦𝑏 in x- and y-direction, respectively. Equation

(3.8) can be integrated with the boundary condition 𝜏𝑥𝑧 = 𝜏𝑥𝑏 at 𝑧 = ℎ to get,

44
𝜕𝑝 𝜕𝑢
𝜏𝑥𝑧 = (𝑧 − ℎ) + 𝜏𝑥𝑏 = 𝜂 (3.9)
𝜕𝑥 𝜕𝑧

Equation (3.9) can be integrated with the boundary condition 𝑢 = 𝑈 at 𝑧 = 0 to yield,

1 𝜕𝑝 𝑧 2 𝜏𝑥𝑏
𝑢=𝑈+ ( − ℎ𝑧) + 𝑧 (3.10)
𝜂 𝜕𝑥 2 𝜂

A similar expression for the velocity profile in y-direction can be derived as,

1 𝜕𝑝 𝑧 2 𝜏𝑦𝑏
𝑣= ( − ℎ𝑧) + 𝑧 (3.11)
𝜂 𝜕𝑦 2 𝜂

The relationship between the boundary yield stress 𝜏𝑏 , and 𝜏𝑥𝑏 and 𝜏𝑦𝑏 can be

expressed as,
2 2
𝜏𝑏 2 = 𝜏𝑥𝑏 + 𝜏𝑦𝑏 (3.12)

From a physical perspective, the shear stress applied on the liquid at the solid/liquid

interface always resists the slip against the surface. The sign of 𝜏𝑥𝑏 and 𝜏𝑦𝑏 should

be opposite with the slip velocity. The slip velocity at the upper stationary solid

surface ( 𝑧 = ℎ) in x-direction is,

1 𝜕𝑝 2 𝜏𝑥𝑏
𝑢𝑠 = 𝑈 − ℎ + ℎ (3.13)
2𝜂 𝜕𝑥 𝜂

Based on the analysis of the sign of 𝜏𝑏𝑥 , Eqn. (3.13) can be rewritten as,

1 𝜕𝑝 2 |𝜏𝑥𝑏 |
𝑢𝑠 = 𝑈 − ℎ − sgn(𝑢𝑠 ) ℎ (3.14)
2𝜂 𝜕𝑥 𝜂

where sgn(𝑢𝑠 ) is the sign function of 𝑢𝑠 and equals to 1 when 𝑢𝑠 is positive and -1

when 𝑢𝑠 is negative. Thus the fluid velocity profile in x-direction can be expressed

as,

1 𝜕𝑝 𝑧 2 |𝜏𝑥𝑏 |
𝑢=𝑈+ ( − ℎ𝑧) − 𝑠𝑔𝑛(𝑢𝑠 ) 𝑧 (3.15)
𝜂 𝜕𝑥 2 𝜂

The similar expressions for the fluid flow in y-direction are,

45
1 𝜕𝑝 2 |𝜏𝑦𝑏 |
𝑣𝑠 = − ℎ − sgn(𝑣𝑠 ) ℎ (3.16)
2𝜂 𝜕𝑦 𝜂

1 𝜕𝑝 𝑧 2 |𝜏𝑦𝑏 |
𝑣= ( − ℎ𝑧) − sgn(𝑣𝑠 ) 𝑧 (3.17)
𝜂 𝜕𝑦 2 𝜂

Equations (3.15) and (3.17) are integrated along the film thickness direction to get

the volumetric flow rate in x- and y-direction,


ℎ3 𝜕𝑝 sgn(𝑢𝑠 )|𝜏𝑥𝑏 |ℎ2
𝑞𝑥 = ∫ 𝑢𝑑𝑧 = 𝑈ℎ − − (3.18)
0 3𝜂 𝜕𝑥 2𝜂


ℎ3 𝜕𝑝 sgn(𝑣𝑠 )|𝜏𝑦𝑏 |ℎ2
𝑞𝑦 = ∫ 𝑣𝑑𝑧 = − − (3.19)
0 3𝜂 𝜕𝑦 2𝜂

Considering the continuity of flow and conservation of mass gives,

𝜕𝑞𝑥 𝜕𝑞𝑦
+ =0 (3.20)
𝜕𝑥 𝜕𝑦

Substituting Eqns. (3.18) and (3.19) into Eqn. (3.20) and rearranging gives,

𝜕 ℎ3 𝜕𝑝 𝜕 ℎ3 𝜕𝑝
( )+ ( )
𝜕𝑥 𝜂 𝜕𝑥 𝜕𝑦 𝜂 𝜕𝑦

𝜕ℎ
=3𝑈 (3.21)
𝜕𝑥

3 𝜕 sgn(𝑢𝑠 )|𝜏𝑥𝑏 |ℎ2 𝜕 sgn(𝑣𝑠 )|𝜏𝑦𝑏 |ℎ2


− { [ ]+ [ ]}
2 𝜕𝑥 𝜂 𝜕𝑦 𝜂

In order to get a more general Reynolds equation for describing both the slip and no-

slip bearing, Eqns. (3.15) and (3.17) can be rearranged as,

|𝜏𝑥𝑏 | 1 𝜕𝑝 2
sgn(𝑢𝑠 ) ℎ=𝑈− ℎ − 𝑢𝑠 (3.22)
𝜂 2𝜂 𝜕𝑥

|𝜏𝑦𝑏 | 1 𝜕𝑝 2 (3.23)
sgn(𝑣𝑠 ) ℎ=− ℎ − 𝑣𝑠
𝜂 2𝜂 𝜕𝑦

These two equations can be substituted into Eqn. (3.21) and rearranged to yield,

46
𝜕 ℎ3 𝜕𝑝 𝜕 ℎ3 𝜕𝑝 𝜕ℎ 𝜕 𝜕
( )+ ( )= 6𝑈 +6 (𝑢𝑠 ℎ) + 6 (𝑣 ℎ) (3.24)
𝜕𝑥 𝜂 𝜕𝑥 𝜕𝑦 𝜂 𝜕𝑦 𝜕𝑥 𝜕𝑥 𝜕𝑦 𝑠

Putting 𝑋 = 𝑥⁄𝑙 , 𝑌 = 𝑦⁄𝑏, 𝑃 = 𝑝(ℎ02 ⁄𝜂𝑈𝑙 ), 𝐻 = ℎ⁄ℎ0 , 𝑉 ∗ = 𝑣𝑠 ⁄𝑈 , 𝑈 ∗ = 𝑢𝑠 ⁄𝑈 ,

𝜏𝑏∗ = 𝜏𝑏 (ℎ0 ⁄𝜂𝑈) into Eqn. (3.24) gives its dimensionless form as,

𝜕 𝜕𝑃 𝑙2 𝜕 𝜕𝑃 𝜕𝐻 𝜕(𝑈 ∗ 𝐻) 𝑙 𝜕(𝑉 ∗ 𝐻)
(𝐻 3 ) + 2 (𝐻 3 ) = 6 +6 +6 (3.25)
𝜕𝑋 𝜕𝑋 𝑏 𝜕𝑌 𝜕𝑌 𝜕𝑋 𝜕𝑋 𝑏 𝜕𝑌

3.4 Solution of modified Reynolds equation

3.4.1 Control equation for wall slip

In one-dimensional flow, the slip velocity of fluid flow is either in the same direction

or opposite direction of the velocity of moving solid surface. However, the slip may

appear at any direction. In a two-dimensional lubricant flow, the shear stress applied

to the liquid at the solid/liquid interface can be expressed as,

𝜏 2 = 𝜏𝑥2 + 𝜏𝑦2 (3.26)

where the subscript x and y denotes the shear stress components in x- and y-direction,

just as shown in Fig. 3.4(a). Apparently, the magnitude of the yield stress boundary

is a function of circle. If the magnitude of shear stress 𝜏 is less the the radius of the

circle, there is no slip. On the contrary, slip occurs if |𝜏| equals the radius of the

circle. Therefore, the slip control function can be express as,

𝜏𝑥2 + 𝜏𝑦2 ≤ 𝜏𝑏2 (3.27)

From Fig. 3.4(a), the slip control function can be written as [111],

mod(𝜏) − 𝜏𝑏 ≤ 0 (3.28)

or

𝜏𝑥 cos𝜃 + 𝜏𝑦 sin𝜃 − 𝜏𝑏 ≤ 0 (3.29)

The slip criteria can be expressed in terms of pressure gradient. Based on Eqn. (3.14),

slip occurs in the positive x-direction if [4],

47
𝜕𝑝 2𝜂 2|𝜏𝑥𝑏 |
< 2𝑈− (3.30)
𝜕𝑥 ℎ ℎ

or in the negative x-direction if,

𝜕𝑝 2𝜂 2|𝜏𝑥𝑏 |
> 2𝑈+ (3.31)
𝜕𝑥 ℎ ℎ

Similarly, the expression of the slip criteria in y-direction can be written as,

𝜕𝑝 2|𝜏𝑦𝑏 |
<− (3.32)
𝜕𝑦 ℎ

(Slip occurs in positive y-direction)

and

𝜕𝑝 2|𝜏𝑏𝑦 |
> (3.33)
𝜕𝑦 ℎ

(Slip occurs in negative y-direction).

3.4.2 Multi-linearity algorithm

Although the slip control equation has been derived in Section 3.4.1, there are
2 2
infinite possibilities because the boundary of slip control equation, 𝜏𝑏 2 = 𝜏𝑏𝑥 + 𝜏𝑏𝑦 ,

is a function of a circle. It is impossible to solve Eqn. (3.25) directly. In order to get

a more accurate solution, a multi-linearity algorithm in which the boundary of yield

stress is represented by a polygon was proposed by Ma et al. [111]. The principle of

this method is shown in Fig. 3.4 (b). Considering a circumscribed polygon with N

sides, the slip control equation (3.27) can be replaced approximately by [111],

𝜏𝑥 cos𝛽𝑖 + 𝜏𝑦 sin𝛽𝑖 − 𝜏𝑏 ≤ 0 (3.34)

where 𝛽𝑖 = 2𝜋(𝑖 − 1)/𝑁 and i is the ith side of the polygon. Apparently, the

accuracy of the calculation increases with the number of polygon side. Ma et al. [111]

analyzed a wall slip case of a squeeze film flow of two parallel discs for confirming

the feasibility and reliability. They tried three types of polygons, i.e. circumscribed

48
polygon, amended polygon and inscribed polygon to approach the circle. Figure 3.5

shows the change of dimensionless load support and relative calculation error with

the sides of polygons when slip occurs only at the lower surface. Apparently, both

dimensionless load support and relative converge to fixed values. For the number of

polygon sides is greater than 16, the values keep almost constant and the error of the

predicted load support is within 2% compared with the converged value. Therefore,

they studied a two-dimensional fluid flow case with L/B=1 (B: breadth of the slider

in the sliding direction; L: length of the slider) in which a circumscribed polygon

with 16 sides is used as the slip boundary.

Fig. 3.4 Yield of boundary shear stress (a) circle and (b) polygon approximation

Fig. 3.5 Change of calculated dimensionless load support (a) and (b) relative
calculation error with the number of polygon sides for ℎ0 ⁄𝑅 = 10−4
(Slip occurs only at lower surface, 𝜏𝐵∗ = 𝜏𝑏 𝑅⁄𝜂𝑣 = 108 ) [111]

49
3.4.3 Scalar method

Although the multi-linearity algorithm is a good idea to solve the two-dimensional

slip problem, it is not easy to use it with finite difference method. On the other hand,

there is only Poiseuille flow along the y-direction and the amount of slip can be very

small compared with that in the x-direction. Hence, it is reasonable to consider the

shear stress as a scalar and calculate the boundary slip in x- and y-direction

separately. Actually, this concept has been adopted by Tauviqirrahman et al. [118].

In this chapter, the scalar method is applied for calculating the boundary slip in a

two-dimensional fluid flow. To verify its validity, the comparison between the

calculated results using the scalar method with the numerical solutions of Ma et al.

[111] should be done first.

By employing a discretization scheme, the whole region is evenly divided into 200 ×

200 nodes. Initially, the slip speed at each grid point is set to zero and the pressure

gradient is obtained from the classical Reynolds equation. In each iteration,

Eqns.(3.30-3.33) are applied for determining the slip speed and shear stress at each

grid point. The procedure of the numerical calculation can refer to Appendix C.

Iteration continues until the convergence condition is satisfied. In the present

calculation, the accuracy of pressure is set to 10-5:

∑𝑁 𝑀 𝑘 𝑘−1
𝑖=1 ∑𝑗=1 |𝑝𝑖,𝑗 − 𝑝𝑖,𝑗 |
𝑘 ≤ 10−5 (3.35)
∑𝑁 𝑀
𝑖=1 ∑𝑗=1 𝑝𝑖,𝑗

3.5 Numerical results and analyses – Effects of boundary yield stress

3.5.1 Normalized load support

50
The calculated change of dimensionless load support with boundary yield stress

under different convergence ratio K was compared with the solution calculated by

Ma’s group [111], as shown in Fig. 3.6. Apparently, the present numerical results

coincide quite well with those given by Ma’s group [111], which proved the

reliability of the developed algorithm. When the boundary yield stress is large

enough, the load capacity is equal to that of no slip boundary condition. To relax the

criterion for slippage, i.e. the boundary yield stress with a smaller magnitude,

slippage would first occur at the bearing outlet, wherein the shear rate, i.e. shear

stress, at bearing surfaces is the largest. The load capacity is thus reduced with the

decrease of boundary yield stress due to the appearance of slippage at the outlet. The

load capacity reaches a minimum value before slippage starts to occur at the inlet.

The appearance of slippage at the inlet region reduces the resistance of oil entering

into the bearing, which is certainly beneficial to the bearing effect. Therefore, the

load capacity increases with the reduction in boundary yield stress. The load capacity

approaches to a stable value when the boundary yield stress is small enough, which

is equivalent to the full slip boundary condition (𝜏𝑏∗ = 0,).

51
-1
10
-2
10
-3
10

W*
-4
10
L/B=1
-5
10 K=1 K=0.1
K=0.001 K=0.01
-6
10
-6 -5 -4 -3 -2 -1 0 1
10 10 10 10 10 10 10 10
b
*

(a) (b)

Fig. 3.6 Comparison of the calculated load result with the solution of Ma’s group: (a)

finite element results calculated by Ma’s group [111] and (b) our calculated results

3.5.2 Normalized friction force


0
10
-1
10
-2
10
-3
10
F*

L/B=1
-4
10 K=1
-5
K=0.1
10 K=0.01
-6 K=0.001
10
-6 -5 -4 -3 -2 -1 0 1
10 10 10 10 10 10 10 10
b
*

(a) (b)

Fig. 3.7 Comparison of the calculated friction result with the solution of Ma’s group:

(a) finite element results calculated by Ma’s group [111] and (b) our calculated

results

Figure 3.7 shows the change of non-dimensional friction results with boundary yield

stress. The calculated results used in this chapter almost coincide with that of

52
calculated by multi-linearity algorithm, which proved the reliability of the present

algorithm again. Generally, the friction force keeps rising with increasing 𝜏𝑏∗ until it

reaches the maximum value corresponding to no-slip conditions on the whole

stationary surface. Besides, the friction drag grows faster for lower convergence ratio

when slip occurs at the whole region, which reflects the increasing of Couette

friction as 𝜏𝑏∗ increases for lower convergence ratio under thin film condition. These

phenomena were also found in one-dimensional flow [4].

3.5.3 Pressure distribution

0.20 0.20
essure

0.15 0.15
essure
Normalized pr

Normalized pr

0.10 0.10

0.05 0.05

1.0 1.0
0.00 0.8 0.00 0.8
0.0 0.0
0.6 0.6
0.2 0.4 x/L 0.2 0.4 x/L
y/B 0.2 y/B 0.2
0.4 0.4
0.0 0.0

(a) 𝜏𝑏∗ = 1.5 (b) 𝜏𝑏∗ = 1.0

0.05 0.12

0.04 0.10
essure
essure

0.08
0.03
Normalized pr
Normalized pr

0.06
0.02
0.04
0.01
0.02
0.00 1.0 1.0
0.8 0.00 0.8
0.0 0.0
0.6 0.6
0.2 0.4 x/L 0.2 0.4
y/B y/B x/L
0.2 0.2
0.4 0.4
0.0 0.0

(c) 𝜏𝑏∗ = 0.5 (d) 𝜏𝑏∗ = 0

Fig. 3.8 Change of pressure distribution with normalized critical shear stress (K=1)

53
Figure 3.8 shows the change of pressure distribution under four normalized

boundary yield stresses. Clearly, with the drop of boundary yield stress, the location

of maximum pressure in the contact area moves from the inlet to the outlet of the

sliding contact. It is reasonable because the slip appears firstly at the outlet and thus

reduces the load capacity of this area. When the boundary yield stress is small

enough, for example 𝜏𝑏∗ = 0, the pressure distribution recovers to that of no-slip

condition but with small value.

3.5.4 Lubricant film thickness under different parameters

In order to get the effect of boundary yield stress on lubricant film thickness under

hydrodynamic lubrication, the change of lubricant film thickness with boundary

yield stress under different parameters was calculated and shown in Figs. 3.9 to 3.11.

In the calculation, the slider size and inclination are fixed as 4 mm × 4 mm (B×L)

and 1:1770, respectively. Figures 3.9 to 3.11 demonstrate the calculated results under

different viscosities, sliding speeds and loads. The film thickness shows the similar

trend with load capacity. Generally speaking, the lubricant film thickness increases

with viscosity, sliding speed and decreases with load under both no-slip and full-slip

conditions, which correspond well to the traditional lubrication behavior. Besides,

the minimum lubricant film thickness also increases with viscosity, sliding speed and

decreases with load. However, a strange phenomenon occurs when partial slip

appears (before appearance of slippage at the inlet). Taking the lubricant with 0.2

Pas and 0.3 Pas cases as examples, the film thicknesses of the case coincides when

the boundary yield stress are in the range of 1100 Pa to 1600 Pa. More interestingly,

the lubricant film thickness corresponding to 0.56 Pas exceeds that of lubricant with

viscosity 1 Pas in some special ranges, for example, from 2 kPa to 3 kPa. The similar

54
phenomenon was found for the calculating results under different sliding speeds (Fig.

3.10). The reason for this phenomenon is that the slip area and its size also play an

important role in determining the lubricant film thickness under the partial slippage

condition.

3.0

1 Pas
Film thickness (m)

2.0 0.56 Pas

0.3 Pas

0.2 Pas
1.0
-3 -2 -1 0 1 2 3 4 5
10 10 10 10 10 10 10 10 10
Boundary yield stress (Pa)

Fig. 3.9 Change of film thickness with boundary yield stress under different

viscosity

4.0

3.0 0.025 m/s


Film thickness (m)

0.015 m/s
2.0
0.01 m/s

0.005 m/s

1.0
-3 -2 -1 0 1 2 3 4 5
10 10 10 10 10 10 10 10 10
Boundary yield stress (Pa)

Fig. 3.10 Change of film thickness with boundary yield stress under different speed

55
3.0
2N
Film thickness (m)

2.0 4N

6N

10 N

1.0
-3 -2 -1 0 1 2 3 4 5
10 10 10 10 10 10 10 10 10
Boundary yield stress (Pa)

Fig. 3.11 Change of film thickness with boundary yield stress under different load

3.6 Summary

In this chapter, the generally accepted slip models used in fluid mechanics and

tribology were introduced. Then, the boundary yield stress model was adopted and

incorporated into the traditional Reynolds equation for its reasonable description of

fluid behavior from an engineering viewpoint. After that, the modified Reynolds

equation considering boundary yield stress model was solved by finite difference

method. The boundary yield stress was considered as a scalar and slip behavior was

calculated in x- and y-direction separately. The calculated results were compared

with the numerical solutions of Ma et al. [111]. The good correlation of the two

calculated values validated the scalar assumption. Finally, the change of film

thickness with boundary yield stress was calculated under different working

conditions, including load, viscosity and speed.

56
Chapter 4

Identifying the Best Interfacial Parameter Correlated with

Hydrodynamic Lubrication

4.1 Introduction

Even though the earliest experimental evidence of liquid/solid boundary slip can date

back to the same era when the lubrication theory was first developed, the no-slip

boundary condition is still widely accepted in engineering practice. The magnitude

of boundary slip in a system of common engineering surfaces and engine oils is

small, i.e. it can be neglected. However, the recent advancement in nano/micrometer

scale technologies leads to a scenario that the seemingly negligible slip may be

significant and its effect can be considerable in the nano/micro world of technology.

On the other hand, the concept of no-slip boundary condition is based on the idea

that the adhesive force between solid and liquid molecules is stronger than the

cohesive force among the liquid molecules. However, following the advent of super-

hydrophobic surfaces as a result of the recent progress in surface science research,

the fluid molecular layer sliding on the outer layer of solid molecules has become

plausible. To quantify the strength of the solid/liquid interface, different facial and

interfacial parameters have been proposed. The most popular one is contact angle

which is the angle formed by the tangent at the contact point of the liquid droplet

profile and the solid plane. Conceptually, larger the contact angle means the smaller

the interfacial forces. Nevertheless, independent investigations have come up with

different conclusions on the relation of the contact angle and friction of a lubricated

contact. For example, Bongaerts et al. [109] found that there is no relationship

between wettability and hydrodynamic force in the EHL regime, as shown in Fig.

57
4.1(a). However, Hild et al. [56] detected that the viscous forces are significantly less

with hydrophobic surfaces when compared with hydrophilic surfaces, as shown in

Fig. 4.1(b). Simply speaking, the results of the surface effect on viscous resistance

provided by the two independent studies are contradictory. Hydrophobic surfaces are

in favor of lubrication as reported in [56] (Fig. 4.1 (b)), but not in [110] (Fig. 4.1 (a)).

(a) (b)

Fig. 4.1 Different findings on relation of friction and contact angle (excerpt from

Refs. [109] and [56] respectively.)

To advance the design of a hydrodynamic lubricated system from the perspective of

interfacial properties, it is worthwhile to identify the key interfacial parameter which

correlates the best with hydrodynamic lubrication. This chapter presents an

examination of some common facial and interfacial parameters, including contact

angle, contact angle hysteresis, and a new spreading parameter proposed by Kalin

and Polajnar [140].

4.2 Interfacial parameters

4.2.1 Contact angle

58
As shown in Fig. 4.2, a liquid drop rests on a solid surface and the contact angle 𝜃 is

defined as the angle formed by the tangent of the droplet and the solid surface. The

interface where the three phases, liquid, solid and vapor meet is referred to as ‘three-

phase contact line’. If the contact angle formed by a drop of water and a solid surface

is less than 90°, the solid surface is hydrophilic and wetting of the solid surface is

favorable, which means water tends to spread on the surface. On the other hand, the

solid surface is hydrophobic and water is not easy to spread on the surface if the

contact angle is greater than 90°. Based on the Young’s theory, the contact angle of a

liquid drop and an ideal solid surface is determined by the mechanical equilibrium of

three surface tensions,

𝛾𝑆𝐿 − 𝛾𝑆𝑉
𝑐𝑜𝑠𝜃𝑌 = (4.1)
𝛾𝐿𝑉

where 𝛾𝑆𝐿 is the solid-liquid interfacial tension, 𝛾𝑆𝑉 is the solid-vapor interfacial

tension, 𝛾𝐿𝑉 is the liquid-vapor interfacial tension and 𝜃𝑌 is the Young contact angle

(the contact angle formed by a liquid drop and an ideal solid surface).

Fig. 4.2 Illustration of contact angle formed by sessile drop on a homogenous

smooth surface

The contact angle of a liquid on a solid surface is related to the inter-molecular

attractive force. Some studies demonstrated the connection of the contact angle to

the lubrication effect [107-110]. A ‘sticky’ surface results in a smaller contact angle

59
due to the strong affinity or adhesion between the liquid and the surface. However, it

does not seem always the case. For example, Wang et al. [141] developed a

technique for fabricating a superoleophobic surfaces with switchable adhesion force

with oil drops. With increasing the treatment time of UV, the sliding angle of the oil

drop changes from 0 to 180 degrees although the contact angle keeps almost constant,

as shown in Fig. 4.3.

Fig. 4.3 Example of strong affinity between a liquid drop with a very large contact
angle and a surface (excerpt from Ref. [141].)

4.2.2 Contact angle hysteresis

Lubrication processes are dynamic in nature, whereas the contact angle is only a

static measurement. Yaminsky [142] discussed the importance of liquid-solid

molecular adsorption to wettability and brought up the idea of interfacial parameter,

contact angle hysteresis (CAH), of a surface and a liquid resembling the effect of

static friction. CAH is the difference between two dynamic parameters, namely

advancing and receding contact angles, which are determined at the impending

moment of the liquid droplet on a plane through adding and removing water to and

from the droplet.

4.2.3 Kalin’s spreading parameter

60
Kalin and Polajnar [140] proposed a new spreading parameter, which is defined as

the difference in the work of adhesion and cohesion, to describe the wetting

phenomenon. They proved that this parameter is better to describe the actual wetting

property of DLC coating and steel with oils instead of contact angle. The spreading

parameter is related to the two components (dispersive and polar) of the solid and the

liquid surface tension, and can be expressed as,

𝑆𝑃 = 2 (√𝛾𝑆𝐷 √𝛾𝐿𝐷 + √𝛾𝑆𝑃 √𝛾𝐿𝑃 − 𝛾𝐿 ) (4.2)

where 𝛾𝐿 is the liquid total surface tension, 𝛾𝑆𝐷 and 𝛾𝑆𝑃 represent the dispersive and

polar component of the solid surface tension, 𝛾𝐿𝐷 and 𝛾𝑆𝑃 denotes the dispersive and

polar component of the liquid surface tension.

4.3 Measurement of interfacial parameters

4.3.1 Measurement of contact angle

Fig. 4.4 Contact angle and contact angle hysteresis measurement test rig.

(1-camera, 2-microscope system, 3-base frame, 4-position adjusting part, 5-syringe)

61
Contact angle and contact angle hysteresis are measured by using a commercial

contact angle goniometer. Figure 4.4 presents the main body of the instrument which

is composed of a camera (1), a microscope system (2), a base frame (3), a position

adjusting system (4) and a syringe (5).

Static sessile drop method is applied for the measurement of contact angle. By

capturing the profile of a liquid drop on a solid surface by using goniometer’s optical

microscope system, the angle formed between the liquid/vapor interface and the

liquid/solid interface is the measured contact angle. During the experiments, the

volume of liquid drops is fixed to 3 µL for each measurement, which means the

effect of weight of the liquid drop on the contact angle can be ignored. For some

specimens, the contact angle will change at first until reaching a stable state. Figure

4.5 shows a series of classical contact angle results formed between glycerol and two

different solid surfaces, EGC (an oleophobic coating) and steel surface. The stable

contact angle was selected as reference for hydrodynamic lubrication research.

100

80 EGC
Contact angle (degree)

Steel

60

40

20
0 10 20 30 40 50 60
Time (second)

Fig. 4.5 Change of contact angle with time

62
4.3.2 Measurement of contact angle hysteresis

Contact angle hysteresis can be obtained by two methods, dynamic sessile drop and

tilting plate method. The principle of dynamic sessile drop method is shown in Fig.

4.6. Dynamic sessile drop method is similar to the static sessile drop method but

requires the volume of the liquid drop to be changed. By adding the volume of the

liquid drop, the maximum contact angle can be reached without increasing the

liquid/solid interface area. This measured largest contact angle is the advancing

contact angle. Similarly, the measured minimum contact angle by reducing the

volume of the liquid drop is defined as the receding contact angle. The difference

between the advancing contact angle and the receding contact angle is the measured

contact angle hysteresis. Figure 4.7 presents the principle of tilting wafer method for

contact angle hysteresis measurement. A drop of liquid is dispensed on a solid

surface, and the solid surface is tilted slightly. Immediately before the moment the

droplet moves, the contact angles in the front and rear of the liquid droplet are

captured. These angles are the advancing and receding contact angles, and their

difference is the measured contact angle hysteresis. The tilting angle of the solid

surface is called the sliding angle.

Fig. 4.6 Advancing and receding contact angles, θa and θr, of a sessile liquid droplet

on a solid surface

63
Fig. 4.7 Principle of tilting wafer method for contact angle hysteresis measurement

In the present study, the dynamic drop method is applied for the measurement of

contact angle hysteresis. That is because the adhesion force between lubricants and

normal surfaces is much stronger than that of water and solid surfaces and the oil

drop cannot move even the sliding angle reaches 180 degrees.

4.3.3 Measurement of surface energy and spreading parameter

Based on the definition, the polar and dispersive components of the solid surface

tension and liquid surface tension should be known before calculating the spreading

parameter. The widely used Owens-Wendt-Rabel-Kaelble (OWRK) [143] method

was applied here for solid surface tension measurement. Demineralized water and

hexadecane were selected as model liquids A and B respectively. The details of the

surface tension of the two model liquids are shown in Table 4.1. The contact angle

formed between the two model liquids and the target solid surface was measured

firstly. Based on the measured contact angle, the surface tension of the solid surface

was calculated according to the Owen-Wendt’s theory [143].

Table 4.1 Surface tension and its components for model liquids

Model Total surface Dispersive Polar component


liquids tension (mN/m) component (mN/m) (mN/m)
Water 72.80 21.80 51.00
Hexadecane 27.60 27.60 0

64
The surface tension of lubricants was determined by using the pendant drop method.

The pendant drop is a drop suspended from a needle with a liquid phase and the

shape of the drop is determined by the relationship between the liquid surface

tension and gravity. From the shadow image of the pendant drop, the liquid’s surface

tension can be calculated according to the Young-Laplace equation.

4.4 Specimen surface and lubricant

Five sliders of the same size but different surfaces were used in the tests. The size of

the sliding plane was 4 mm (Breath, B) × 9 mm (Length, L). Surface and bulk

materials of the sliders and their surface roughness are listed in Table 4.2. Besides

the untreated steel slider which provides a steel surface (#1-steel slider), steel sliders

were respectively coated with a kind of oleophobic PTFE coating (#2-referred to as

‘EGC’ slider) and a commercial anti-fingerprint coating (#3- ‘AFC’ slider). The

commercial anti-fingerprint coating is a kind of oleophobic thin film coating with

CF3 bonds. It is a consumer product for the protection of the touch screen of phones

or PCs from oils and water. Glass sliders with thin chromium coating were also used

in the tests. One of them was further coated with a layer of SiO2. Hence, at the end,

there were two other surfaces: chromium (#4-Cr slider) and SiO2 (#5-SiO2 slider).

Surface roughness of the sliders is maintained at a nanometer level, except the one

with the EGC coating, which is relatively rougher and its roughness is about ten

times of the others.

65
Table 4.2 Sliders adopted in the study

Slider Bulk material Surface layer Surface tension (mN/m) Roughness (nm)

#1 Steel Steel 60.61 6

#2 Steel EGC 25.09 49

#3 Steel AFC* 34.01 9

#4 Glass Cr 54.56 2

#5 Glass SiO2 47.22 2


*
Anti-fingerprint coating

Table 4.3 Properties of lubricants used in the experiments

Lubricant Dynamic viscosity Refractive

(22℃, mPas) Index

65% Glycerol 14 1.45

99% Glycerol 704 1.47

PAO400/40 880 1.47

Three lubricants, 65 wt% glycerol, 99 wt% glycerol and PAO400/40, were used in

the tests and the properties of these lubricants are tabulated in Table 4.3. The contact

angle, CAH and the spreading parameter of the sliding surfaces and these lubricants

were measured. As wetting parameters are quite sensitive to solid surface, all the

specimen slider surfaces were cleaned with the exact procedure. Firstly, the sliders

were cleaned in an ultrasonic bath of alcohol for 30 minutes. Then the alcohol left on

the slider surfaces were removed by cleaning tissues. Finally, the sliders were blow-

dried for five minutes. Furthermore, each set of data was tested for at least six times.

The data, as shown in Table 4.4, are listed in groups of the same lubricant and in

ascending order of the contact angle.

66
Table 4.4 Contact angle and contact angle hysteresis of lubricants on slider surfaces

Slider Lubricant Contact angle Contact angle hysteresis Spreading

surface CA (o) CAH (o) parameter

Steel 65% Glycerol 37.9+9.2


−6.9 47.7+3.0
−3.1 10.33

SiO2 65% Glycerol 51.0+9.0


−8.2 30.5+2.1
−3.0 10.27

Cr 65% Glycerol 65.3+4.6


−5.2 29.4+2.3
−1.9 4.23

AFC 65% Glycerol 87.0+3.3


−1.7 49.6+1.2
−0.9 -17.91

EGC 65% Glycerol 105.0+2.2


−1.1 16.2+1.3
−1.4 -45.99

Steel 99% Glycerol 45.0+5.2


−4.2 46.7+1.6
−1.1 1.60

EGC 99% Glycerol 89.2+1.1


−1.7 22.8+1.1
−1.0 -64.67

Steel PAO400/40 28.5+3.5


−7.1 33.6+1.5
−1.2 30.16

AFC PAO400/40 68.6+2.1


−1.3 32.5+0.5
−0.9 NA.

EGC PAO400/40 73.9+1.5


−2.9 23.0+1.1
−1.3 0.72

The experiments were carried out in a controlled environment (ambient temperature:

22±1oC, humidity: 60±2%). Every set of data was measured within twenty minutes

and fresh glycerol was used for each set of experiments.

4.5 Comparison of different parameters

The experiments with 65 wt% glycerol were conducted using different sliders with a

constant load of 5 N and a constant inclination of 1:2036. Figure 4.8 shows the

variations of the film thickness against speed for the test with 65 wt% glycerol

solution. To make good understanding the lubrication process, two theoretical film

thickness-speed curves are also depicted. The theoretical results were obtained with

67
a full two-dimensional finite difference solution of Reynolds equation for the

lubricated contact. The theoretical curve noted with no-slip corresponds to the

classical Reynolds equation as expressed in Eqn. (4.3). The theoretical full-slip

results were calculated by the extended Reynolds equation with full-slip boundary

conditions as described by Eqn. (4.4) [4]. Equation (4.4) was derived based on the

boundary shear stress slip model and taking the boundary shear stress equal to zero.

Comparing the terms on the right-hand side of the two equations, the one with full-

slip boundary conditions (Eqn. (4.4)) is only half of the one with no-slip boundary

conditions in magnitude. As a result, the theoretical film thickness with full-slip

boundary conditions is lower, as shown in Fig. 4.8.

𝜕 𝜕𝑝 𝜕 𝜕𝑝 𝑑ℎ
(ℎ3 ) + (ℎ3 ) = 6𝑢𝜂 (4.3)
𝜕𝑥 𝜕𝑥 𝜕𝑦 𝜕𝑦 𝑑𝑥

𝜕 𝜕𝑝 𝜕 𝜕𝑝 𝑑ℎ
(ℎ3 ) + (ℎ3 ) = 3𝑢𝜂 (4.4)
𝜕𝑥 𝜕𝑥 𝜕𝑦 𝜕𝑦 𝑑𝑥

where h is the local film thickness, p is the pressure, u is the speed and η is the

dynamic viscosity. As shown in Fig. 4.8, the steel and AFC sliders provided the

largest film thickness, which coincided with the classical no-slip hydrodynamic

lubrication theory. The lubricating film thickness formed by the SiO2 and Cr sliders

was almost the same and was slightly less than that generated by the AFC and steel

sliders. The lubricating film thickness generated by the EGC slider was the lowest

and its variation with speed is well correlated by the full-slip hydrodynamic

lubrication theory. It shows that the molecular bonding of the EGC and the lubricant

is weak. The five sets of tests were conducted with the same lubricant under the

exact conditions (same load and inclination in the same speed range). The only

difference is their surfaces. The surface roughness of the EGC slider is about ten

times of the others, as shown Table 4.2, but it is still an order of magnitude smaller

68
than the minimum film thickness measured. Hence, the low film thickness obtained

with the EGC slider must not be due to the relatively large surface roughness.

Otherwise, the larger roughness would have enhanced the hydrodynamic effect

leading to a higher film thickness. Therefore, the difference in film thickness of the

sliders is attributed to the surface or interfacial effect. The correlations of film

thickness with contact angle, CAH, the spreading parameter and surface energy of

solid surfaces are shown in Fig. 4.9. The film thickness decreases largely with the

increase in contact angle, as shown in Fig. 4.9(a), which matches the general concept,

but there is a scattered point from the AFC slider. Its contact angle is the second

largest among all the sliders, but it generates relatively large film thickness. A

similar case was found for the surface energy of solid surfaces (Fig. 4.9(d)).

Although the surface energy between steel and AFC are big enough, the thickness

generated by these two surfaces are almost the same. There is no apparent

relationship between the spreading parameter and lubricant film thickness, as shown

in Fig. 4.9(b). The spreading parameter corresponding to AFC and steel is quite

different although they generate the same glycerol film thickness. The film thickness

generated by SiO2 and chrome is smaller than that of AFC, but their spreading

parameters are larger than the one corresponding to AFC. On the other hand, Fig.

4.9(c) shows that the correlation of film thickness and CAH is much better.

69
1.5
Theoretical (no slip)
Theoretical (full slip)
Antifinger
Steel
Film thickness (m)
1.0 SiO2
Cr
EGC

0.5

0.0
0 10 20 30
Speed (mm/s)

Fig. 4.8 Change of film thickness with speed (65 wt% glycerol, load: 5 N)

1.4
Film thickness (m)

1.2

EGC
1.0 Cr
SiO2
Steel
AFC
0.8
40 60 80 100
Contact angle (degree)

(a)

70
1.4

Film thickness (m)


1.2

1.0

0.8
-50 -40 -30 -20 -10 0 10
Spreading parameter (mN/m)

(b)

1.4
Film thickness (m)

1.2

1.0

0.8
20 30 40 50
Contact angle hysteresis (degree)

(c)

71
1.6

Film thickness (m) 1.4

1.2

1.0

0.8
20 30 40 50 60
Surface energyof solid surface (mN/m)

(d)

Fig. 4.9 Correlation of film thickness and (a) contact angle; (b) spreading parameter;

(c) contact angle hysteresis; (d) surface energy of solid surface

The differences between the steel and EGC sliders, which illustrate the two extremes

in film forming capacity as shown in Fig. 4.8, were further evaluated by repeating

the test with 99 wt% glycerol. The viscosity of the specimen lubricant increases five

times, as shown in Table 4.3, by increasing the concentration of glycerol from 65 to

99 wt%. The chemical properties of the two are expected to be very much the same.

The tests were conducted with two loads, 4 N and 8 N, and the inclination of the

slider was 1:1697. The results of two different loads are depicted in log-log plots

shown in Fig. 4.10(a) and (b), respectively. The film thickness data of the steel slider

are well correlated by the no-slip theoretical curves in the specified speed range for

the two loads. However, the film thickness generated with the EGC slider is quite

small. Due to the instability of interference images of the EGC slider under slow

72
speeds, only film thickness measured at the higher speeds are shown in the figures.

The minimum film thickness shown in Fig. 4.10 is still about five times greater than

the roughness of the EGC slider as listed in Table 4.2. It proves that there was no

direct contact of the two running surfaces. The great reduction in film thickness with

the presence of EGC coating on the slider surface can be attributed to the weak

affinity of the lubricant and the slider surface. The contact angle hysteresis of EGC

surface and steel with 99 wt% glycerol are, respectively, 22.8o and 46.1o as

illustrated in Table 4.4. It shows that the larger the CAH, the higher the film

thickness. Figure 4.10(a) and (b) illustrate a strange phenomenon that the resultant

film thickness achieved by the EGC slider is even lower than the full-slip theoretical

curves.

Theoretical (no slip)


Theoretical (full slip)
5.0 Steel
Film thickness (m)

EGC

1.0

0.5
2.5 5.0 7.5 25.0
Speed (mm/s)

(a)

73
Theoretical (no slip)
5.0 Theoretical (full slip)
Steel
Film thickness (m) EGC

1.0

0.5

0.1
2.5 5.0 7.5 25.0
Speed (mm/s)

(b)

Fig. 4.10 Change of film thickness with speed (99% glycerol, load: (a) 4 N; (b) 8 N)

The experiments were repeated with a polyalphaolefin oil. PAO400/40 was selected

because it has about the same viscosity of 99 wt% glycerol, but different polarity.

PAO400/40 is a non-polar oil while glycerol is polar. Figure 4.11 shows the change

of PAO400/40 film thickness with speed (inclination: 1:1858) for loads of 4 N and

10 N. The steel and AFC sliders were used because they have a large magnitude

difference in contact angle with PAO400/40 but are very much the same in CAH, as

shown in Table 4.4. Figure 4.11 shows that the two film thickness curves of steel and

AFC sliders cannot differentiate themselves and both of them correspond well with

the classical no-slip hydrodynamic lubrication theory. The two slider surfaces have a

large difference in contact angle (by 40o) but about the same CAH. This proves

further that the CAH but not contact angle can reflect the interface-dependent

hydrodynamic lubrication effect generated by the two different slider surfaces.

74
10.0
4N
5.0

Film thickness (m)


10 N

1.0
Theoretical
0.5 AFC
Steel

0.5 2.5 5.0


Speed (mm/s)

Fig. 4.11 Change of film thickness with speed for two loads (PAO400/40)

The EGC slider which was proven having very weak adhesive strength with glycerol

was also evaluated with PAO400/40. EGC has a large contact angle with

PAO400/40 (73.9o), which is very close to that of the AFC surface (68.6o), but it has

a smaller CAH (23.0o) than those of AFC and steel surfaces (32.5o and 33.6o,

respectively). The PAO400/40 film thickness generated by the three sliders: AFC,

steel and EGC at different speeds are shown in Fig. 4.12. The inclination was 1:2064

in these experiments. The results of the steel and AFC sliders are correlated very

well by the classical no-slip hydrodynamic theory for both loads of 2 N and 4 N. The

film thickness generated with the EGC slider, however, is much smaller than that of

steel and AFC sliders. The difference in these film thickness results from different

surfaces corresponds well with their CAH, but not the contact angle. Furthermore,

the results shown in Fig. 4.10 and 4.12 point out a strange fact that the film thickness

generated by the EGC slider with 99 wt% glycerol or PAO400/40 is even lower

those full-slip theoretical values.

75
10.0

5.0
Film thickness (m)

1.0
Theroetical (no slip)
0.5 Theoretical (full slip)
AFC
Steel
EGC

0.1
5.0 10.0 15.0 20.0 25.0
Speed (mm/s)

(a)

10.0

5.0
Film thickness ( m)

Theoretical (no slip)


Theoretical (full slip)
AFC
1.0 Steel
EGC
0.5

0.1
5.0 10.0 15.0 20.0 25.0
Speed (mm/s)

(b)

Fig. 4.12 Change of film thickness with speed (PAO400/40, load: (a) 2 N; (b) 4 N)

76
4.6 Why CAH is the best?

The experimental results obtained using sliders of different surfaces show the surface

effect on hydrodynamic lubrication. The hydrodynamic film forming capability is

related to the adhesive strength between the liquid and the solid surface. The liquid

molecules can move relatively or slide on the solid surface only if the liquid

molecules have gained enough energy to overcome the energy barrier brought along

with the adhesion force between the liquid and solid molecules. The energy barrier is

thus governed by the interfacial properties of the surface and the liquid, which can be

characterized by the two interfacial parameters: contact angle and contact angle

hysteresis. Previous experimental studies [80, 81] and the present results show that

there is no direct relationship between the contact angle and hydrodynamic

lubrication behavior. Whyman et al. [144] derived based on thermodynamic

principles that the potential energy barrier is related with the contact angle and the

contact angle hysteresis, and it can be expressed as,

𝛾𝑅
𝑈𝑝 = (CAH)2 𝑓(𝜃) (4.5)
27/3

(1 + cos 𝜃)1⁄2
𝑓(𝜃) = (4.6)
(1 − cos 𝜃)1⁄6 (2 + cos 𝜃)4⁄3

where 𝛾 is the liquid surface tension, 𝑅 is the radius of the spherical drop on solid

surface and 𝑈𝑝 is the potential energy barrier. The potential energy barrier can be

calculated from contact angle 𝜃 and CAH. However, the potential barrier U does not

show a strong function of contact angle in a broad range, approximately from 20o to

140o , because 𝑓(𝜃) changes very little and keeps almost constant in the middle

range, as shown in Fig. 4.13. Hence, the potential energy barrier is largely dependent

on CAH, especially in the contact angle range of 20o to 140o . In general, a surface

which has a contact angle greater than 90o is considered as hydrophobic. Hence, a

77
small slider bearing having hydrophobic surfaces may not be necessarily inferior in

lubricating film formation than one using hydrophilic surface. In the present

experiments, the contact angles of all specimen lubricants and slider surfaces are in

this insensitive contact angle range. According to Eqn. (4.5), the smaller the CAH,

the smaller the potential energy barrier is. Extrand [145] also pointed out that the

major cause of CAH is the molecular interaction between the contacting solid and

liquid. That is why the measured hydrodynamic lubricant film thickness correlates

well with CAH in the study. Barrat and Bocquet [77] also found from their MDS

results that boundary slip becomes apparent only when the contact angle is greater

than 140o . Similar conclusion was drawn by Huang et al. [78]. Their results on the

variation of slip length with contact angle show the increase becoming more

significant at the contact angle of 140o.

Fig. 4.13 Change of f(θ) with contact angle

4.7 Summary

Three interfacial parameters: contact angle, contact angle hysteresis and spreading

parameter were compared for their correlation with the hydrodynamic lubricating

78
effect of a slider bearing. Five very smooth slider surfaces of different materials and

three liquids, polar and non-polar, which provided contact angles ranging from 40o

to 110o, were used in the bearing tests. Contact angle hysteresis, but not the contact

angle and Kalin’s spreading parameter, was found closely correlate with the

hydrodynamic effect. The finding is also supported by an existing theory derived

based on thermodynamic principles that the adhesive strength between the

contacting solid and liquid molecules is a strong function of contact angle hysteresis

but not the contact angle if its value falls into the range of 20o to 140o.

79
Chapter 5

Correlation of Boundary Yield Stress and Interfacial Potential

Energy Barrier

5.1 Introduction

The idea that fluid might slip on a solid surface was first proposed by Navier in 1823

[11]. As introduced in Chapter 1, many studies have found that boundary slip can

only occur when the shear stress at the liquid/solid interface attains a yield value [40,

53]. Slip would start if the fluid flow achieves a condition that the magnitude of the

shear stress at the solid/liquid interface exceeds the adhesive strength between the

solid and the liquid molecules. Thus, the boundary yield stress and the adhesive

strength at the solid/liquid interface are probably related.

The adhesive strength at the solid/liquid interface is commonly correlated to the

contact angle. Conceptually, larger the contact angle means the smaller the

interfacial forces. However, the results in Chapter 4 have confirmed that thin film

hydrodynamic lubrication is a stronger function of another interfacial parameter,

contact angle hysteresis (CAH), than contact angle, especially in the contact angle

range of 20o - 140o. CAH is the difference between two dynamic parameters, namely

advancing and receding contact angles, which are determined under the condition of

impending motion of the liquid droplet on a horizontal plane by adding more liquid

to and removing from the drop, respectively. The potential energy barrier to the

movement of liquid molecules on the solid surface, E, was derived based on

thermodynamic principles by Whyman et al. [144] as,

80
𝛾𝑅 2
(1 + cos 𝜃)1⁄2
𝐸 = 7/3 (CAH) (5.1)
2 (1 − cos 𝜃)1⁄6 (2 + cos 𝜃)4⁄3

where θ is the contact angle, 𝛾 is the liquid surface tension and 𝑅 is the radius of the

liquid drop before deposition on solid surface. Thus, the movement of liquid

molecules will start after the energy increase reaches the potential barrier.

The aim of the study in this chapter is to examine the correlation of two parameters:

boundary yield stress and potential energy barrier as they have quite similar physical

meaning. It is of interest to consider how a mechanical parameter, boundary yield

stress, might be related to a physical behavior at the interface. Thin film

hydrodynamic lubrication tests were conducted using a steel slider bearing.

Lubricant samples were prepared by adding different concentrations of

perfluoroheptanoic acid (C6F13COOH) in Silicon oil 201-500. Contact angle and

CAH of different oil samples with the slider surface were measured and the potential

barriers were then calculated. Boundary yield stress was derived from the

experimental results by a two-dimensional hydrodynamic lubrication analysis using

a simple boundary yield stress slip model.

5.2 Specimen surface and lubricant

Two steel sliders with different sizes (4 mm × 4 mm and 4 mm × 6 mm, (B × L))

were employed in this study, and the width B is along the sliding direction. Surface

roughness, Ra, of the sliders is approximately 9 nm.

The control variables in this study are contact angle and CAH. They can be prepared

by modifying the surface with coating of different materials, or adding different

amount of additive into the base oil. Actually, the means of adding additives into

81
fluids has long been used in the study of wettability effect on flow behaviors. Pit et

al. [31] used a fluorescence recovery technique to assess the movement of liquid

close to a solid surface and obtained different shear rates with different

concentrations of stearic acid solution in hexadecane for under the same operating

conditions. The result was attributed to the change of boundary condition. The

stearic acid adsorbed on the solid surface through its polar head and formed a

monolayer with its methyl group towards the liquid [146, 147]. Henry et al. [81]

found that the drainage force between a colloid probe and a mica surface can be

changed if a cationic surfactant is added to pure water. Shafrin and Zisman [148]

studied the wettability of low energy surface, and found that solid surfaces of very

low surface energy could be accomplished by preparing adsorbed monolayer from

CF3 solutions on solids. Thus, fluorine-containing organic acids are expected very

effective as additives in changing the wettability. The idea was later verified by Fuks

and Berlin [149]. They performed lubrication tests with a wide spectrum of oil

samples which were prepared with different base oils and four perfluorinated acids

as additives. Results confirmed that these additives could effectively prevent the

spreading of the specimen oils and the magnitude of spreading could be controlled

by the concentration of these additives. Obviously, perfluorinated acids are quite

ideal additives in changing boundary conditions at the solid/fluid interface and hence

regulating the lubrication behavior.

82
750

700

Viscosity (mPas)
650
201-500+0.1%
201-500+0.2%
201-500+0.3%
600

550

500
0 100 200 300 400
Time (hour)

Fig. 5.1 Temporal change in viscosity of specimen oils with different concentrations

of C6F13COOH

This study employed Silicon oil 201-500 as the base oil and specimen oils were

prepared with five different concentrations of perfluorohepanoic acid (C6F13COOH).

It is shown in Fig. 5.1 that the viscosity of silicon oil increases with time when

perfluoroheptanoic acid is added. By controlling the preparing time, the specimen

lubricants with the same viscosity were obtained. Moreover, each experiment was

completed within half an hour, which allowed only little change in viscosity during

the experiment. It is postulated that C6F13COOH may act as a catalyst to link the

polymer chains of the silicon oil. Therefore, the lubricant viscosity is larger with

higher concentrations of C6F13COOH due to longer molecular chains. Figure 5.2

shows the infrared spectra of silicon oil samples with and without the addition of

C6F13COOH. The two spectra are well overlapping. Thus, there is no chemical

change after adding C6F13COOH in the silicon oil. The properties of these five

83
lubricants were listed in Table 5.1. In this study, the contact angle and contact angle

hysteresis formed between the lubricants and the steel slider surfaces were measured

before the slider tests. Each data point is the average of six independent

measurements taken at different spots of the surface. Given that wetting related

parameters are quite sensitive to surface conditions, all slider surfaces used in this

study were cleaned with the same procedure. The sliders were initially washed with

alcohol in an ultrasonic bath for 30 minutes. The remaining alcohol on the surface

was then removed by cleaning tissues. They were finally dried by blowing air for

five minutes. The dynamic sessile drop method was used to get the CAH.

1.4

1.2
Silicon oil 201-500
Silicon oil 201-500+0.3% additive
1.0
Amplitude

0.8

0.6

0.4

0.2

0.0
4000 3500 3000 2500 2000 1500 1000 500
Infrared spectrum

Fig. 5.2 Comparing of infrared spectra between silicon oils with and without

C6F13COOH.

84
Table 5.1. Properties of specimen oils

C7HF13O2, wt% Dynamic viscosity Refractive

(21oC, mPas) index

0.05% 559.9 1.40

0.10% 559.9 1.40

0.15% 561.9 1.40

0.20% 560.9 1.40

0.25% 560.9 1.40

All the experiments were conducted in a controlled temperature and humidity

environment in order to eliminate the environmental effect (ambient temperature:

21±0.5℃, humidity: 70±2%).

5.3 Results and discussion

The measured contact angle and CAH of all specimen oils are shown in Fig. 5.3. The

contact angle increases while CAH decreases with the increase in the additive

concentration. The potential energy barrier for all oil samples was calculated based

on their contact angle and CAH using Eqn. (5.1), and their normalized values are

also plotted in Fig. 5.3. The normalized potential energy barrier is expressed as,

∗ 7/3
𝐸 (CAH)2 (1 + cos 𝜃)1⁄2
𝐸 =2 = (5.2)
𝛾𝑅 (1 − cos 𝜃)1⁄6 (2 + cos 𝜃)4⁄3

Apparently, a higher concentration of perfluoroheptanoic acid leads to a larger

contact angle and a smaller CAH. Figure 5.3 shows that the magnitude of the

potential energy barrier also decreases with the additive concentration, which is

attributed to the fact that perfluoroheptanoic acid forms a chemical layer on the steel

85
surface. The boundary layer has a weak bonding with CF3 [146-148], as

schematically illustrated in Fig. 5.4, which gives a small apparent surface energy and

leads to the change in contact angle and CAH. The increase in the concentration of

perfluoroheptanoic acid results in the reduction in CAH as well as the potential

energy barrier (Fig. 5.3). Hence, CAH and the potential energy barrier between a

lubricant and a solid surface are correlated.

The lubricant film thickness between the slider and the rotational disc was measured

at different speeds and loads with two sliders of different sizes. The inclination angle

of the slider, which was fixed for each set of experiment, was monitored through the

number of fringes in the interferogram of the lubricating contact during a test to

ensure that it has no change at different speeds. Furthermore, all interference images

of the bearing contact showed no elastic deformation, which indicates that the

experiments were under the hydrodynamic lubrication regime.

80 2.0
Contact angle
Normalized potential energy barrier

70 Contact angle hysteresis


Normalized potential energy barrier
1.5
CA and CAH (degree)

60

50
1.0
40

30
0.5
20

10 0.0
0.05 0.10 0.15 0.20 0.25
Concentration (%)

Fig. 5.3 Contact angle, contact angle hysteresis and potential energy barrier as

functions of C6F13COOH concentrations

86
Fig. 5.4 The connection between perfluoroheptanoic acid and steel surface

The experiments with two specimen oils were conducted using the slider of 4 mm ×

6 mm. The slider has a constant inclination of 1:1770 and two different loads of 4 N

and 6 N. Figure 5.5 presents the change of film thickness with speed for the tests

with two silicon oils containing 0.05% and 0.20% perfluoroheptanoic acid. The

theoretical curve with no slip was gained from the conventional two-dimensional

Reynolds equation as shown in Eqn. (4.3). It can be seen that the measured film

thickness increases with speed and decreases with load, which agrees well with the

classical hydrodynamic lubrication theory. The specimen oil containing a higher

concentration of perfluoroheptanoic acid provided lower film thickness. The

differences in film thickness are more apparent at high speeds. The drop of film

thickness for the silicon oils containing 0.05% and 0.20% perfluoroheptanoic acid

from the no-slip theoretical value reaches 3% and 15%, respectively, at the

maximum speed for the load of 6 N. It indicates that the molecular bonding of the

lubricant with 0.20% perfluoroheptanoic acid is weaker than that with 0.05%.

87
5
Theory
0.05%
4 0.20% w=4 N
Film thickness (m)

w=6 N

5 10 15 20 25 30
Speed (mm/s)

Fig. 5.5 Change of film thickness with speed (slider size: 4 mm × 6 mm)

Similar results were obtained with the other slider of 4 mm × 4 mm, which is smaller

in width. The measured film thickness at different speeds for three different loads of

2 N, 4 N and 6 N, and a fixed inclination of 1:1770 is plotted in Fig. 5.6. Two

theoretical curves which were obtained with no-slip and full-slip boundary

conditions are also illustrated in Fig. 5.6. The no-slip theoretical curve was

calculated using the Reynolds equation (Eqn.(4.3)). The full-slip curve was from a

modified Reynolds equation which is derived with the yield shear stress model and

by taking the yield stress at the slider/lubricant interface equal to zero. Overall

speaking, the film thickness decreases with increasing the concentration of

perfluoroheptanoic acid. The silicon oil with 0.05% C6F13COOH presented the

highest film thickness among all lubricant samples and was only slightly less than

the no-slip theoretical curve (calculated with the no-slip Reynolds equation,

Eqn.(4.3)). On the other hand, the sample oil containing 0.25% C6F13COOH

generated the lowest film thickness, but it is still larger than the full-slip theoretical

88
curve (from the full-slip Reynolds equation, Eqn.(4.4)). Furthermore, even the

measured minimum film thickness is much larger than the surface roughness of the

sliders, the effect of roughness on film thickness can thus be ignored in this study.

Therefore, the reduction in film thickness with increasing concentration can only be

attributed to the decrease in the affinity between the lubricant and the slider surface.

4
Film thickness (m)

Theory (no-slip)
2
Theory (full slip)
0.15%
0.20%
0.25%

1
5 10 15 20 25 30
Speed (mm/s)

(a) Load: 2 N

89
4

3
Film thickness (m)

2
Theory (no-slip)
Theory (full slip)
0.05%
0.10%
0.15%
1 0.20%
0.25%
5 10 15 20 25 30
Speed (mm/s)

(b) Load: 4 N

3
Film thickness (m)

Theory (no-slip)
Theory (full slip)
1 0.05%
0.15%
0.20%
0.25%

5 10 15 20 25 30
Speed (mm/s)

(c) Load: 6 N

Fig. 5.6 Change of film thickness with speed for different loads

(slider size: 4 mm × 4 mm)

In order to describe the relationship between boundary yield stress and potential

90
energy barrier, the boundary yield stresses under the current experimental conditions

were calculated. The method described in Chapter 3 was adopted here. The change

in film thickness with boundary yield stress was calculated with experimental

conditions of Fig. 5.6(b) and is shown in Fig. 5.7. Based on the measured film

thickness, the boundary yield stresses under different speeds for these five specimen

oils were obtained from Fig. 5.7. Figure 5.8 depicts the curves of boundary yield

stress against normalized potential energy barrier for four different speeds. The

boundary yield stress increases with the potential energy barrier. The boundary yield

stress and potential energy barrier are monotonously related since both terms are

critical values for the movement of liquid molecules from the solid surface.

3
21.3 mm/s
16.6 mm/s
Film thickness (m)

13.0 mm/s
2

6.45 mm/s

-2 -1 0 1 2 3 4 5
10 10 10 10 10 10 10 10
2
Boundary yield stress (N/m )

Fig. 5.7 Change of film thickness with boundary yield stress

(under condition in Fig. 5.6 (b).)

91
4.5

Boundary yield stress (kPa)


4.0 21.3 mm/s

3.5 16.6 mm/s

13.0 mm/s
3.0

2.5
6.45 mm/s
2.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Normalized potential energy barrier

Fig. 5.8 Correlation between boundary yield stress and potential energy barrier

(under condition in Fig. 5.6(b).)

5.4 Summary

The monotonous relationship between boundary yield stress and potential energy

barrier was identified under thin film hydrodynamic lubrication conditions. For

surfaces of high potential energy barrier which is more difficult to generate

movement of the liquid molecules on the solid surface, the magnitude of boundary

yield shear stress is higher.

92
Chapter 6

Sliding Speed and Viscosity Effect on Boundary Slippage

6.1 Introduction

In Chapter 4, experiments have proved that the contact angle hysteresis CAH is the

key parameter in reflecting the strength of the adhesive force and determining the

boundary yield stress for a given liquid and different solid surfaces. However, the

experimental results show some interesting phenomena for lubricants with different

viscosities. For examples, the 65% glycerol film thickness corresponding to EGC

coating is almost equal to the theoretical value under full slip condition (Fig.

4.8),while the 99% glycerol film thickness is, however, much smaller than the

theoretical full-slip value (Fig. 4.7). These two similar oils, 65% and 95% glycerol,

have different viscosities and different CAH with the EGC slider. Therefore, it is

reasonable to expect that the boundary yield stress is viscosity dependent. Actually,

some researchers have focused on correlation between slip phenomenon and

viscosity. For example, Craig et al. [53] measured hydrodynamic drainage force with

AFM and found that slip length increases with viscosity for Newtonian liquids under

higher shear rate. However, Cho et al. [82] showed that there is no dependence of

slip length on liquid viscosity by using the same experimental method as Craig et al.

[53]. They also noted that the viscosity of solutions used by Craig et al. [53] is ten

times larger than theirs. Cottin-Bizonne et al. [45] drew the same conclusion as Craig

et al. [53] about the relation between viscosity and boundary slip with a dynamic

surface force apparatus. The traditional lubrication theory is based on viscous

lubricant flow. It is thus meaningful to clarify the relation of viscosity and boundary

slippage.

93
Furthermore, the boundary yield stress is found sliding speed dependent (Fig. 5.8).

The boundary yield stress increases with sliding speed. The sliding speed of the glass

disc affects the shear rate and probably the slip velocity at the solid/liquid interface.

Many experiments concluded that boundary slip is a function of shear rate. Craig et

al. [53] carried out experiments with Newtonian liquids using AFM and obtained

that the slip length increases with shear rate (noted with driving rate). Horn et al. [39]

studied the boundary slip phenomenon of non-Newtonian liquids by using SFA and

came up with the same conclusion as Craig et al. [53]. Besides, polymer physicists

have argued that shear rate might play an important role in determining the slip

length when polymers are adsorbed to the solid surfaces [150, 151]. However, the

opposite phenomenon was also found in some experimental studies. By using the

Near Field Laser Velocimetry technique, Hervet and Leger [88] studied the

boundary slip phenomenon and found that the slip length is independent of shear rate.

Vinogradova and Yakubov [52] investigated the water flow on hydrophobic surfaces

using an AFM and found that the slip length does not depend on shear rate. Cottin-

Bizonne et al. [45] measured the viscous force with a dynamic surface force

apparatus and found that the slip length is independent of shear rate, which is in

agreement with the result of Vinogradova and Yakubov [52].

Spikes and Granick [152] proposed a slip model in which the slip velocity or shear

rate and the liquid’s viscosity are considered. In this new slip model, it is believed

that the boundary slip occurs only when the shear stress at the solid/liquid boundary

exceeds a critical value, and the magnitude of boundary shear stress would increase

94
with the slip velocity and the liquid’s viscosity. The shear stress when slip occurs is

given by

𝑢s
𝜏𝑏 = 𝜏𝑐 + 𝜂 (6.1)
𝑏

where 𝜏𝑏 is the shear stress at the solid/liquid interface under slip condition, 𝜏𝑐 is the

critical shear stress and 𝜂 is the viscosity of the fluid. This model combines both

features of the linear slip length model and boundary yield stress model. It is

meaningful to verify this model for it providing a quantitative guideline for boundary

slip design.

In this chapter, the effect of shear rate and the viscosity of liquid on boundary yield

stress was investigated.

6.2 Effect of sliding speed on boundary yield stress

Because the shear rate and slip velocity at the slider/lubricant interface varies on the

slider surface and the slip velocity cannot be directly obtained from the present

experiments, sliding speed of the glass disc was used here as the studying parameter

with the assumption that the slip velocity is directly related with the sliding speed of

the glass disc. Figure 5.8 describes the correlation between boundary yield stress and

potential energy barrier. However, it also shows that the slip shear stress increases

with the sliding speed. The speed-dependency indicates that the calculated boundary

yield stress based on the experimental film thickness measurements is not a system

constant, but depends on the operating conditions of the experiments. Figure 6.1

reproduces the data shown in Fig. 5.8. Clearly, the change of boundary yield stress is

linearly related with the sliding speed. Based on the previous assumption, it can be

concluded that the boundary yield stress increases monotonously with slip velocity at

95
the solid/liquid interface, which is correspond well with the slip model of Spikes and

Granick [152].

4000.0
Boundary yield stress (N/m )
2

3000.0
0.05%
0.15%
0.2%
0.25%
2000.0

5 10 15 20
Speed (mm/s)

Fig. 6.1 Change of boundary yield stress with sliding speed

6.3 Effect of viscosity on boundary yield stress

6.3.1 Specimen surface and lubricant

The study in Chapter 4 has proved that EGC coating can significantly change the

interfacial property of the steel slider with some lubricants. Therefore, the steel slider

with EGC coating was applied in this study. The size of the sliding surface was 4

mm (Breath, B) × 9 mm (Length, L). The roughness, Ra, of the EGC slider was 20

nm.

The research objective in this study is the viscosity effect on boundary slippage. Two

types of lubricants were prepared. The first type is glycerol solutions. Solutions with

five different concentrations of glycerol provide a series of viscosity. Three different

96
PAO oils were also used here, PAO4, PAO10 and the combination of half PAO4 and

half PAO10 (by weight). The viscosities of all applied lubricants are listed in Table

6.1. The contact angle and contact angle hysteresis formed between the lubricants

and EGC surface were also measured because they affect the lubrication behavior

significantly, especially the contact angle hysteresis. The measured contact angle and

contact angle hysteresis are listed in Table 6.2. Apparently, the five glycerol

solutions correspond to almost the same contact angle and contact angle hysteresis. It

is reasonable because they have almost the same chemical component. It is similar

for these three PAO oils.

Table 6.1 Properties of the used lubricants

Lubricants Viscosity, mPas, @22 oC Refractive Index

80% Glycerol solution 41.8 1.45

85% Glycerol solution 75.0 1.45

90% Glycerol solution 157.5 1.455

92% Glycerol solution 214.2 1.46

95% Glycerol solution 381.9 1.46

PAO4 28.1 1.45

PAO4/10 (1:1) 58.7 1.45

PAO10 132 1.45

97
Table 6.2 Contact angle and contact angle hysteresis formed with EGC

Lubricant Contact angle Contact angle hysteresis

80% Glycerol solution 107.98 15.37

85% Glycerol solution 106.74 15.97

90% Glycerol solution 106.21 15.81

92% Glycerol solution 106.53 15.79

95% Glycerol solution 107.49 15.02

PAO4 75.58 23.0

PAO4/10 (1:1) 74.72 22.1

PAO10 76.71 24.1

The experiments were carried out in a controlled environment (temperature: 22±1oC,

humidity: 70±2%). Besides, each experiment was finished within twenty minutes,

and fresh glycerol solution was used for each set of experiments.

6.3.2 Results and discussion

Figure 6.2 shows the change of lubricant film thickness with speed by using different

glycerol solutions. In this experiment, the load and inclination are fixed as 2 N and

1:1745, respectively. Apparently, all the film thicknesses corresponding to these five

glycerol solutions are lower than the theoretical film thickness with full-slip

condition. Based the calculated results in Chapter 3, the measured results are located

in the range 3 or 4, which is marked in Fig. 6.3. With the increase of lubricant

viscosity, the measured film thickness of glycerol solution approaches the full-slip

theoretical value (dotted line) gradually, especially under high speed. This change

cannot be attributed to the interface affinity for the same contact angle hysteresis

98
provided by these five lubricants with EGC. The most possible parameter affecting

this phenomenon is viscosity. However, there is no measured film thickness

exceeding the full-slip theoretical curve, leading to no way to distinct the viscosity

effect on boundary yield stress. The reason is that with the boundary yield stress

increasing in region 3 and decreasing in region 4, the lubricant film thicknesses both

approach to the theoretical value under full-slip conditions. In order to find out more

evidences, a new experiment was designed. The results are illustrated in Fig. 6.4.
Film thickness (m)

1.0

Theory (no slip)


Theory (full slip)
80% Glycerol, 2 N

0.1
5 10 15 20
Speed (mm/s)

(a)

99
3.0

Film thickness (m) 2.0

1.0
Theory (no slip)
Theory (full slip)
85% Glycerol, 2 N

5 10 15 20
Speed (mm/s)

(b)

5.0 Theory (no slip)


Theory (full slip)
4.0 90% Glycerol, 2 N
Film thickness (m)

3.0

2.0

1.0
5 10 15 20
Speed (mm/s)

(c)

100
Theory (no slip)
5.0 Theory (full slip)
92% Glycerol, 2 N

Film thickness (m)


4.0

3.0

2.0

5 10 15 20
Speed (mm/s)

(d)

8.0
Theory (no slip)
Theory (full slip)
6.0 95% Glycerol, 2 N
Film thickness (m)

4.0

2.0
5 10 15 20
Speed (mm/s)

(e)

Fig. 6.2 Change of lubricant film thickness with speed by different glycerol solutions

Figure 6.4 shows the change of film thickness of PAO oils with speed. The film

thickness of all experiments was adjusted to about the same through different loads

101
with the same speed, which realized about the same apparent shear rate. In this set of

experiments, three PAO oils were used (Table 6-1). The inclination corresponding to

the three PAO oils was same and fixed as 1:1745. In rheology, the boundary yield

stress can be expressed as [136]

𝜏𝑏 = 𝜏𝑏0 + 𝑘𝑝 (6.2)

where 𝜏𝑏0 is the boundary yield stress at atmospheric pressure. Besides, Briscoe and

Evans [153] studied the correlation between shear stress and contact pressure for

alipathic carboxylic acids monolayers (a special boundary slip) and got the same

conclusions. Many experiments [153-156] have proved that the value of k is in the

range of 0.007 to 0.15. Therefore, the load effect on boundary yield stress in the

present study can be ignored for the quite small value of k. These results present the

same trends that the lubricant film thickness increases with the lubricant viscosity.

Specially, the film thickness corresponding to PAO 10 exceeds the full-slip

theoretical curve in high speeds, although it is still less than the traditional

theoretical value. That means the film thickness reaches the region 2 for PAO10

under the last two measured speeds in the present study. Apparently, there is no

chance for the film thickness jumping from the region 4 to region 2 directly.

Therefore, the conclusion can be confirmed that the boundary yield stress increases

from region 3 to region 2 marked in Fig. 6.3 with increasing the viscosity of PAO

oils.

To evaluate the change of boundary yield stress with viscosity quantitatively, the

boundary yield stress was calculated from the data shown in Fig. 6.4. The results are

shown in Fig. 6.5. Figure 6.5 shows that the boundary yield stress increases linearly

102
with the viscosity of lubricant under all sliding speeds. These results correspond well

with the slip model proposed by Spikes and Granick [152].

Film thickness 1

5
3
4

Boundary yield stress (Pa)

Fig. 6.3 Correlation between lubricant film thickness and boundary yield stress

3
Theory (no slip)
Theory (full slip)
2 pao4, 1 N
Film thickness (m)

5 10 15 20
Speed (mm/s)

(a)

103
3 Theory (no slip)
Theory (full slip)
pao4+pao10 (1:1) 2 N
2
Film thickness (m)

5 10 15 20
Speed (mm/s)

(b)

3.0 Theory (no slip)


Theory (full slip)
pao10, 4 N
Film thickness (m)

2.0

1.0

5 10 15 20
Speed (mm/s)

(c)

Fig. 6.4 Change of PAO oils film thickness with speed

104
0.9 ud = 21.3 mm/s

Boundary yield stress (kPa)


ud = 13.0 mm/s
ud = 7.9 mm/s
0.6

0.3

0.0
25 50 75 100 125
Viscosity (mPas)

Fig. 6.5 Correlation between boundary yield stress and viscosity for PAO lubricants
used in the study

6.4 Summary

The effect of sliding speed and viscosity on boundary yield stress was investigated in

this chapter. Experimental results show that the boundary yield stress increases with

sliding speed and lubricant’s viscosity. These phenomena confirmed the slip model

proposed by Spikes and Granick [152].

105
Chapter 7

Conclusions and Suggestions for Future Work

7.1 Summary of main conclusions

This thesis focuses on the boundary slip in thin film hydrodynamic lubrication. The

main conclusions are summarized as follows.

7.1.1 Development of test rig for thin film hydrodynamic lubrication

An existing test rig (optical fixed-incline slider) was further developed for

facilitating on-line and simultaneous measurements of friction force and lubricant

film thickness in Chapter 2. The design for friction measurement was realized by

using a load cell of high sensitivity with careful calibration of the internal friction of

load arm assembly. The on-line film thickness measurement was achieved by a

newly developed dichromatic interferometry technique.

7.1.2 New optical method for lubricating film thickness measurement

In Chapter 2, a new optical method for measuring the film thickness of

hydrodynamic lubricated contact using dichromatic interferometry was developed.

Two light beams with different wavelengths are required for simultaneous projection

on the lubricating contact. Superimposing the two interference patterns results in an

envelope of intensity difference and film thickness variation equivalent to an

intensity and film thickness curve with a longer cycle. Using the equivalent intensity

curve, the measurement range of film thickness can be enhanced compared with

monochromatic interferometry. The moving rate of the equivalent intensity curve

corresponding to a rapid change in film thickness was fairly slow. Hence,

106
dichromatic interferometry can facilitate real-time determination of lubricating film

thickness and is suitable for investigating transient or dynamic lubricating problems.

7.1.3 Solution of modified Reynolds equation

The solution of two-dimensional modified Reynolds equation considering the

boundary yield stress slip model on the stationary solid surface was not easy to

obtain for the unknown magnitude and direction of the boundary slip. In Chapter 3,

the boundary yield stress was thus taken as a scalar quantity to simplify the solution

process. The results were validated by comparing with existing solutions of other

research groups. The change of lubricant film thickness with boundary yield stress

was calculated under different working conditions, such as load, speed and viscosity.

The results show that the lubricant film thickness can be much thinner than the

theoretical value corresponding to the full-slip condition.

7.1.4 The best interfacial parameter correlated with hydrodynamic lubrication

In Chapter 4, the correlation of various interfacial parameters, including contact

angle, contact angle hysteresis, a newly proposed spreading parameter and surface

energy of solid surface, with lubricant film thickness in hydrodynamic lubrication

was studied experimentally. The experimental results proved that contact angle

hysteresis is the most significant parameter in determining the boundary slippage and

lubrication behavior within the working conditions of the present study. The

conclusion was consistent with the theory proposed by Whyman et al. [144] who

described the molecular interaction between a liquid and a solid surface from a

physical point of view.

107
7.1.5 Correlation between boundary yield stress and potential energy barrier

In Chapter 5, the study successfully correlated a mechanical parameter, boundary

yield stress, with a physical parameter, potential energy barrier between a liquid and

a solid surface. The two parameters carry similar physical meanings which describe

the minimum energy or force in separating the liquid from the solid surface. They

were found to be monotonously related, based on experimental results obtained

using the combinations of different surfaces and a same type of oil or a single slider

surface with various specimen oils which were prepared by a silicon oil with

different amount of additive, perfluorohepanoic acid.

7.1.6 Sliding speed and viscosity effect on boundary yield stress

The effect of sliding speed and viscosity on boundary yield stress was studied in

Chapter 6. It was found that the boundary yield stress increases linearly with sliding

speed and the lubricant’s viscosity. These two conclusions are consistent with the

slip model proposed by Spikes and Granick [152].

7.2 Suggestions for future work

The boundary slip was studied through lubricant film thickness measurement.

However, it is equally meaningful to analyze this correlation through friction data.

Although the calibration of the test rig was conducted before taking the friction force

measurement, it was found that the internal force in the bearing connecting load arm

and the body of the test rig is a dynamic value and load dependent. More attention

should be paid to the friction force measurement, such as on-line calibration or

through the small deformation from the holder of slider.

108
The monotonous relationship between sliding speed and the lubricant’s viscosity

with boundary yield stress was found. It is thus worthwhile to evaluate the slip

model of Spikes and Granick [152] with the present experimental results. The model,

as shown in Eqn. (6.1), contains two parameters, which makes it difficult to

complete a full theoretical analysis. In fact, the two parameters of the model may be

able to extract from experimental data obtained under a wider range of conditions.

109
Reference

[1] Hamrock BJ. Fundamentals of Fluid Film Lubrication. New York: McGraw-
Hill Inc.; 1994.
[2] Gohar R, Rahnejat H. Fundamentals of Tribology. London: Imperial College
Press; 2008.
[3] Tzanakis I, Hadfield M, Thomas B, Noya S, Henshaw I, Austen S. Future
perspectives on sustainable tribology. Renewable and Sustainable Energy
Reviews. 2012;16:4126-40.
[4] Spikes HA. The half-wetted bearing. Part 1: extended Reynolds equation.
Proceedings of the Institution of Mechanical Engineers, Part J: Journal of
Engineering Tribology. 2003;217:1-14.
[5] Spikes HA. The half-wetted bearing. Part 2: potential application in low load
contacts. Proceedings of the Institution of Mechanical Engineers, Part J:
Journal of Engineering Tribology. 2003;217:15-26.
[6] Bernoulli D. Hydrodynamica. 1738;59. Cited in Neto C, Evans DR,
Bonaccurso E, Butt HJ, Craig VSJ. Boundary slip in Newtonian liquids: a
review of experimental studies. Reports on Progress in Physics.
2005;68:2859-97.
[7] Du Buat PLG. Principes d’Hydraulique. 1786;92-3. Cited in Neto C, Evans
DR, Bonaccurso E, Butt HJ, Craig VSJ. Boundary slip in Newtonian liquids:
a review of experimental studies. Reports on Progress in Physics.
2005;68:2859-97.
[8] Coulomb CA. Mémoires de l’Institut National des Sciences et des Arts:
Sciences Mathématiques et Physiques. 1801;3. Cited in Neto C, Evans DR,
Bonaccurso E, Butt HJ, Craig VSJ. Boundary slip in Newtonian liquids: a
review of experimental studies. Reports on Progress in Physics.
2005;68:2859-97.
[9] Girard PS. Mémoires de la Classe des Sciences Mathématiques et Physiques
de l’Institut de France. 1815;14:329. Cited in Neto C, Evans DR, Bonaccurso
E, Butt H-J, Craig VSJ. Boundary slip in Newtonian liquids: a review of
experimental studies. Reports on Progress in Physics. 2005;68:2859-97.

110
[10] Riche de Prony GCFM. Recherches Physico-Mathématiques sur la Théorie
des Eaux Courantes Paris. 1804. Cited in Neto C, Evans DR, Bonaccurso E,
Butt H-J, Craig VSJ. Boundary slip in Newtonian liquids: a review of
experimental studies. Reports on Progress in Physics. 2005;68:2859-97.
[11] Navier CLMH. Mémoire sur les lois du mouvement des fluids. Mem Acad
Sci Inst France. 1823;6:389-416. Cited in Neto C, Evans DR, Bonaccurso E,
Butt H-J, Craig VSJ. Boundary slip in Newtonian liquids: a review of
experimental studies. Reports on Progress in Physics. 2005;68:2859-97.
[12] Bingham EC. Fluidity and plasticity: McGraw-Hill Book Compny,
Incorporated; 1922.
[13] Stokes GG. Trans Cambridge Phil Soc. 1845;3:287. Cited in Neto C, Evans
DR, Bonaccurso E, Butt H-J, Craig VSJ. Boundary slip in Newtonian liquids:
a review of experimental studies. Reports on Progress in Physics.
2005;68:2859-97.
[14] Poiseuille J. Recherches expérimentales sur le mouvement des liquides dans
les tubes de trés-petits diamétres C R Acad Sci. 1841. Cited in Neto C, Evans
DR, Bonaccurso E, Butt H-J, Craig VSJ. Boundary slip in Newtonian liquids:
a review of experimental studies. Reports on Progress in Physics.
2005;68:2859-97.
[15] Darcy H. Recherches expérimentales relatives au mouvement de léau dans
les tuyaux. Paris: Mallet-Bachelier. 1857. Cited in Neto C, Evans DR,
Bonaccurso E, Butt H-J, Craig VSJ. Boundary slip in Newtonian liquids: a
review of experimental studies. Reports on Progress in Physics.
2005;68:2859-97.
[16] von Helmholtz H, Piotrowski C. ¨UberReibung tropfbarer Fl¨ussigkeiten
Sitzungsberichte der Kaiserlich Akademie der Wissenschaften (Wien). 1860
40:607. Cited in Neto C, Evans DR, Bonaccurso E, Butt H-J, Craig VSJ.
Boundary slip in Newtonian liquids: a review of experimental studies.
Reports on Progress in Physics. 2005;68:2859-97.
[17] Maxwell JC. On the viscosity or internal friction of air and other gases. The
Scientific Papers of James Clerk Maxwell (Cambridge: Cambridge
University Press) 1890;2:1-25. Cited in Neto C, Evans DR, Bonaccurso E,
Butt H-J, Craig VSJ. Boundary slip in Newtonian liquids: a review of
experimental studies. Reports on Progress in Physics. 2005;68:2859-97.
111
[18] Whetham WCD. Phil Trans R Soc Lond Ser A. 1890;181:559. Cited in Neto
C, Evans DR, Bonaccurso E, Butt H-J, Craig VSJ. Boundary slip in
Newtonian liquids: a review of experimental studies. Reports on Progress in
Physics. 2005;68:2859-97.
[19] Couette M. études sur le frottement des liquides. Ann Chim Phys.
1890;21:433-510. Cited in Neto C, Evans DR, Bonaccurso E, Butt H-J, Craig
VSJ. Boundary slip in Newtonian liquids: a review of experimental studies.
Reports on Progress in Physics. 2005;68:2859-97.
[20] Ladenburg R. ¨Uber der Einfluss von W¨anden auf die Bewegung einer
Kugel in einer reibenden Fl¨ussigkeit. Ann Phys. 1907;4. Cited in Neto C,
Evans DR, Bonaccurso E, Butt H-J, Craig VSJ. Boundary slip in Newtonian
liquids: a review of experimental studies. Reports on Progress in Physics.
2005;68:2859-97.
[21] Tretheway DC, Meinhart CD. Apparent fluid slip at hydrophobic
microchannel walls. Physics of Fluids. 2002;14:L9-L12.
[22] Joseph P, Tabeling P. Direct measurement of the apparent slip length.
Physical Review E. 2005;71:035303.
[23] Degre G, Joseph P, Tabeling P, Lerouge S, Cloitre M, Ajdari A. Rheology of
complex fluids by particle image velocimetry in microchannels. Applied
Physics Letters. 2006;89:024104.
[24] Nghe P, Degre G, Tabeling P, Ajdari A. High shear rheology of shear
banding fluids in microchannels. Applied Physics Letters. 2008;93:204102.
[25] Bouzigues CI, Tabeling P, Bocquet L. Nanofluidics in the debye layer at
hydrophilic and hydrophobic surfaces. Physical Review Letters.
2008;101:114503.
[26] Buscall R. Wall slip in dispersion rheometry. Journal of Rheology.
2010;54:1177-83.
[27] Medhi BJ, Kumar AA, Singh A. Apparent wall slip velocity measurements in
free surface flow of concentrated suspensions. International Journal of
Multiphase Flow. 2011;37:609-19.
[28] Figueredo-Cardero A, Chico E, Castilho L, Medronho RD. Particle image
velocimetry (PIV) study of rotating cylindrical filters for animal cell
perfusion processes. Biotechnol Progress. 2012;28:1491-8.

112
[29] Kikuchi K, Mochizuki O. Micro PIV measurement of slip flow on a hydrogel
surface. Measurement Science and Technology. 2014;25:065702.
[30] Terekhov VI, Smulsky YI, Sharov KA, Zolotukhin AV. Boundary-layer
structure in the flow around the cellular surface in a flat channel.
Thermophysics Aeromechanics. 2014;21:701-6.
[31] Pit R, Hervet H, Leger L. Friction and slip of a simple liquid at a solid
surface. Tribology Letters. 1999;7:147-52.
[32] Pit R, Hervet H, Leger L. Direct experimental evidence of slip in hexadecane:
Solid interfaces. Physical Review Letters. 2000;85:980-3.
[33] Zettner CM, Yoda M. Particle velocity field measurements in a near-wall
flow using evanescent wave illumination. Experiments in Fluids.
2003;34:115-21.
[34] Jin S, Huang P, Park J, Yoo JY, Breuer KS. Near-surface velocimetry using
evanescent wave illumination. Experiments Fluids. 2004;37:825-33.
[35] Kazoe Y, Iseki K, Mawatari K, Kitamori T. Evanescent wave-based particle
tracking velocimetry for nanochannel flows. Analytical Chemistry.
2013;85:10780-6.
[36] Li ZZ, D'eramo L, Monti F, Vayssade AL, Chollet B, Bresson B, et al. Slip
length measurementsusing uPIV and TIRF-based velocimetry. Israel Journal
of Chemistry. 2014;54:1589-601
[37] Li ZZ, D’eramo L, Lee C, Monti F, Yonger M, Tabeling P, et al. Near-wall
nanovelocimetry based on Total Internal Reflection Fluorescence with
continuous tracking. Journal of Fluid Mechanics. 2015;766:147-71.
[38] Campbell SE, Luengo G, Srdanov VI, Wudl F, Israelachvili JN. Very low
viscosity at the solid–liquid interface induced by adsorbed C60 monolayers.
Nature. 1996;382:520-2.
[39] Horn RG, Vinogradova OI, Mackay ME, Phan-Thien N. Hydrodynamic
slippage inferred from thin film drainage measurements in a solution of
nonadsorbing polymer. Journal of Chemical Physics. 2000;112:6424-33.
[40] Zhu YX, Granick S. Rate-dependent slip of Newtonian liquid at smooth
surfaces. Physical Review Letters. 2001;87:096105.
[41] Baudry J, Charlaix E, Tonck A, Mazuyer D. Experimental evidence for a
large slip effect at a nonwetting fluid-solid interface. Langmuir.
2001;17:5232-6.
113
[42] Zhu YX, Granick S. No-slip boundary condition switches to partial slip when
fluid contains surfactant. Langmuir. 2002;18:10058-63.
[43] Zhu YX, Granick S. Limits of the hydrodynamic no-slip boundary condition.
Physical Review Letters. 2002;88:106102.
[44] Persson B, Mugele F. Squeeze-out and wear: fundamental principles and
applications. Journal of Physics: Condensed Matter. 2004;16:R295.
[45] Cottin-Bizonne C, Steinberger A, Cross B, Raccurt O, Charlaix E.
Nanohydrodynamics: The intrinsic flow boundary condition on smooth
surfaces. Langmuir. 2008;24:1165-72.
[46] Pottier B, Fretigny C, Talini L. Boundary condition in liquid thin films
revealed through the thermal fluctuations of their free surfaces. Physical
Review Letters. 2015;114:227801.
[47] Xue YH, Wu Y, Pei XW, Duan HL, Xue QJ, Zhou F. How solid-liquid
adhesive property regulates liquid slippage on solid surfaces? Langmuir.
2015;31:226-32.
[48] Vinogradova OI. Drainage of thin liquid film confined bet hydrophoic
surfaces. Langmuir. 1995;11:2213-20.
[49] Vinogradova OI. Coagulation of hydrophobic and hydrophilic solids under
dynamic conditions. Journal of Colloid Interface Science. 1995;169:306-12.
[50] Vinogradova OI. Hydrodynamic interaction of curved bodies allowing slip
on their surfaces. Langmuir. 1996;12:5963-8.
[51] Vinogradova OI, Butt HJ, Yakubov GE, Feuillebois F. Dynamic effects on
force measurements. I. Viscous drag on the atomic force microscope
cantilever. Review of Scientific Instruments. 2001;72:2330-9.
[52] Vinogradova OI, Yakubov GE. Dynamic effects on force measurements. 2.
Lubrication and the atomic force microscope. Langmuir. 2003;19:1227-34.
[53] Craig VS, Neto C, Williams DR. Shear-dependent boundary slip in an
aqueous Newtonian liquid. Physical Review Letters. 2001;87:054504.
[54] Bonaccurso E, Kappl M, Butt HJ. Hydrodynamic force measurements:
Boundary slip of water on hydrophilic surfaces and electrokinetic effects.
Physical Review Letters. 2002;88:076103.
[55] Neto C, Craig VS, Williams DR. Evidence of shear-dependent boundary slip
in Newtonian liquids. The European Physical Journal E. 2003;12:71-4.

114
[56] Hild W, Opitz A, Schaefer JA, Scherge M. The effect of wetting on the
microhydrodynamics of surfaces lubricated with water and oil. Wear.
2003;254:871-5.
[57] McBride SP, Law BM. Improved in situ spring constant calibration for
colloidal probe atomic force microscopy. Review of Scientific Instruments.
2010;81.
[58] Rodrigues TS, Butt HJ, Bonaccurso E. Influence of the spring constant of
cantilevers on hydrodynamic force measurements by the colloidal probe
technique. Colloid Surface A. 2010;354:72-80.
[59] Bowles AP, Honig CDF, Ducker WA. No-slip boundary condition for weak
solid-liquid interactions. Journal of Physical Chemistry C. 2011;115:8613-21.
[60] Zhu LW, Attard P, Neto C. Reliable measurements of interfacial slip by
colloid probe atomic force microscopy. II. hydrodynamic force
measurements. Langmuir. 2011;27:6712-9.
[61] Zhu LW, Attard P, Neto C. Reconciling slip measurements in symmetric and
asymmetric systems. Langmuir. 2012;28:7768-74.
[62] Pan Y, Li D, Zhao X. An improved method for measuring boundary slip on
hydrophobic surface with atomic force microscope. 2013 13th Ieee
Conference on Nanotechnology (Ieee-Nano). 2013:422-5.
[63] Bhushan B, Pan YL, Daniels S. AFM characterization of nanobubble
formation and slip condition in oxygenated and electrokinetically altered
fluids. Journal of Colloid and Interface Science. 2013;392:105-16.
[64] Jing DL, Bhushan B. Quantification of surface charge density and its effect
on boundary slip. Langmuir. 2013;29:6953-63.
[65] Ahmad K, Zhao XZ, Pan YL, Wang WJ, Huang YD. Atomic force
microscopy measurement of slip on smooth hydrophobic surfaces and
possible artifacts. Journal of Physical Chemistry C. 2015;119:12531-7.
[66] Kiseleva OA, Sobolev VD, Churaev NV. Slippage of the aqueous solutions
of cetyltrimethylammonium bromide during flow in thin quartz capillaries.
Colloid Journal. 1999;61:263-4.
[67] Watanabe K, Udagawa Y, Udagawa H. Drag reduction of Newtonian fluid in
a circular pipe with a highly water-repellent wall. Journal of Fluid Mechanics.
1999;381:225-38.

115
[68] Mala GM, Li D. Flow characteristics of water in microtubes. International
Journal of Heat and Fluid Flow. 1999;20:142-8.
[69] Cheng JT, Giordano N. Fluid flow through nanometer-scale channels.
Physical Review E. 2002;65:031206.
[70] Kim J, Kim CJC. Nanostructured surfaces for dramatic reduction of flow
resistance in droplet-based microfluidics. Proceedings, IEEE Micro Electro
Mechanical Systems. 2002:479-82.
[71] White J, Ma H, Lang J, Slocum A. An instrument to control parallel plate
separation for nanoscale flow control. Review of Scientific Instruments.
2003;74:4869-75.
[72] Choi CH, Westin KJA, Breuer KS. Apparent slip flows in hydrophilic and
hydrophobic microchannels. Physics of Fluids. 2003;15:2897-902.
[73] Wu H, Cheng P. Friction factors in smooth trapezoidal silicon microchannels
with different aspect ratios. International Journal of Heat and Mass Transfer.
2003;46:2519-25.
[74] Ou J, Perot B, Rothstein JP. Laminar drag reduction in microchannels using
ultrahydrophobic surfaces. Physics of Fluids. 2004;16:4635-43.
[75] Cheikh C, Koper G. Stick-slip transition at the nanometer scale. Physical
review Letters. 2003;91:156102.
[76] Cui HH, Silber-Li ZH, Zhu SN. Flow characteristics of liquids in microtubes
driven by a high pressure. Physics of Fluids. 2004;16:1803-10.
[77] Barrat JL, Bocquet L. Large slip effect at a nonwetting fluid-solid interface.
Physical Review Letters. 1999;82:4671-4.
[78] Huang DM, Sendner C, Horinek D, Netz RR, Bocquet L. Water slippage
versus contact angle: a quasiuniversal relationship. Physical Review Letters.
2008;101:226101.
[79] Bonaccurso E, Butt HJ, Craig VS. Surface roughness and hydrodynamic
boundary slip of a Newtonian fluid in a completely wetting system. Physical
Review Letters. 2003;90:144501.
[80] Joseph P, Tabeling P. Direct measurement of the apparent slip length.
Physical Review E. 2005;71:035303.
[81] Henry CL, Neto C, Evans DR, Biggs S, Craig VSJ. The effect of surfactant
adsorption on liquid boundary slippage. Physica A: Statistical Mechanics and
Its Applications. 2004;339:60-5.
116
[82] Cho JH, Law B, Rieutord F. Dipole-dependent slip of Newtonian liquids at
smooth solid hydrophobic surfaces. Physical Review Letters.
2004;92:166102.
[83] Granick S, Zhu YX, Lee H. Slippery questions about complex fluids flowing
past solids. Nature Materials. 2003;2:221-7.
[84] Hao PF, Yao ZH, He F, Zhu KQ. Experimental investigation of water flow in
smooth and rough silicon microchannels. Journal of Micromechanics and
Microengineering. 2006;16:1397-402.
[85] Hetsroni G, Mosyak A, Pogrebnyak E, Yarin LP. Fluid flow in micro-
channels. International Journal of Heat and Mass Transfer. 2005;48:1982-98.
[86] Kandlikar SG, Joshi S, Tian SR. Effect of surface roughness on heat transfer
and fluid flow characteristics at low reynolds numbers in small diameter
tubes. Heat Transfer Engineering. 2003;24:4-16.
[87] McHale G, Newton MI. Surface roughness and interfacial slip boundary
condition for quartz crystal microbalances. Journal of Applied Physics.
2004;95:373-80.
[88] Hervet H, Leger L. Flow with slip at the wall: from simple to complex fluids.
Comptes Rendus Physique. 2003;4:241-9.
[89] Léger L. Friction mechanisms and interfacial slip at fluid–solid interfaces.
Journal of Physics: Condensed Matter. 2003;15:S19.
[90] Koplik J, Banavar JR, Willemsen JF. Molecular dynamics of fluid flow at
solid surfaces. Physics of Fluids A: Fluid Dynamics (1989-1993).
1989;1:781-94.
[91] Jabbarzadeh A, Atkinson J, Tanner R. Effect of the wall roughness on slip
and rheological properties of hexadecane in molecular dynamics simulation
of Couette shear flow between two sinusoidal walls. Physical Review E.
2000;61:690.
[92] Cottin-Bizonne C, Barentin C, Charlaix E, Bocquet L, Barrat JL. Dynamics
of simple liquids at heterogeneous surfaces: molecular-dynamics simulations
and hydrodynamic description. The European Physical Journal E.
2004;15:427-38.
[93] Cottin-Bizonne C, Barrat JL, Bocquet L, Charlaix E. Low-friction flows of
liquid at nanopatterned interfaces. Nature materials. 2003;2:237-40.

117
[94] Ligrani P, Blanchard D, Gale B. Slip due to surface roughness for a
Newtonian liquid in a viscous microscale disk pump. Physics of Fluids.
2010;22:052002.
[95] Schmatko T, Hervet H, Leger L. Effect of nanometric-scale roughness on
slip at the wall of simple fluids. Langmuir. 2006;22:6843-50.
[96] Galea TM, Attard P. Molecular dynamics study of the effect of atomic
roughness on the slip length at the fluid-solid boundary during shear flow.
Langmuir. 2004;20:3477-82.
[97] Priezjev N, Darhuber A, Troian S. Slip behavior in liquid films on surfaces
of patterned wettability: Comparison between continuum and molecular
dynamics simulations. Physical Review E. 2005;71.
[98] Choi CH, Kim CJ. Large slip of aqueous liquid flow over a nanoengineered
superhydrophobic surface. Physical Review Letters. 2006;96:066001.
[99] Cottin-Bizonne C, Cross B, Steinberger A, Charlaix E. Boundary slip on
smooth hydrophobic surfaces: Intrinsic effects and possible artifacts.
Physical Review Letters. 2005;94:056102.
[100] Thompson PA, Troian SM. A general boundary condition for liquid flow at
solid surfaces. Nature. 1997;389:360-2.
[101] Priezjev NV, Troian SM. Molecular origin and dynamic behavior of slip in
sheared polymer films. Physical Review Letters. 2004;92:018302.
[102] Smith FW. Lubricant behaviour in concentrated contact-some rheological
problems. ASLE Tribology Transactions. 1960;3(1):18-25.
[103] Johnson KL, Tevaarwerk JL. Shear behavior of elastohydrodynamic oil films.
Proceedings of the Royal Society of Lond Series A. 1977;356:215-36.
[104] Hersey MD. Theory of Lubrication. New York: J Wiley and Sinc Inc; 1936.
[105] Spikes HA. Film-forming additives: direct and indirect ways to reduce
friction. Lubrication Science. 2002;14(2):147-67.
[106] Chappuis J. Lubrication by a new principle-the use of non-wetting liquids.
Wear. 1982;77:303-13.
[107] Choo JH, Glovnea RP, Forrest AK, Spikes HA. A low friction bearing based
on liquid slip at the wall. Journal of Tribology-Transactions of the ASME.
2007;129:611-20.

118
[108] Choo JH, Spikes HA, Ratoi M, Glovnea R, Forrest A. Friction reduction in
low-load hydrodynamic lubrication with a hydrophobic surface. Tribology
International. 2007;40:154-9.
[109] Bongaerts JHH, Fourtouni K, Stokes JR, Jin ZM. Soft-tribology: lubrication
in a compliant pdms-pdms contact. Tribology International. 2007;40:1531-42.
[110] Guo F, Yang SY, Ma C, Wong PL. Experimental study on lubrication film
thickness under different interface wettabilities. Tribology Letters.
2014;54:81-8.
[111] Ma GJ, Wu CW, Zhou P. Multi-linearity algorithm for wall slip in two-
dimensional gap flow. International Journal for Numerical Methods in
Engineering. 2007;69:2469-84.
[112] Guo F, Wong PL. Theoretical prediction of hydrodynamic effect by tailored
boundary slippage. Proceedings of the Institution of Mechanical Engineers,
Part J: Journal of Engineering Tribology. 2006;220:43-8.
[113] Fortier AE, Salant RF. Numerical analysis of a journal bearing with a
heterogeneous slip/no-slip surface. Journal of Tribology-Transactions of the
ASME. 2005;127:820-5.
[114] Wu CW, Ma GJ, Zhou P, Wu CD. Low friction and high load support
capacity of slider bearing with a mixed slip surface. Journal of tribology.
2006;128:904-7.
[115] Guo F, Wong PL, Fu Z, Ma C. Interferometry measurement of lubricating
films in slider-on-disc contacts. Tribology Letters. 2010;39:71-9.
[116] Nagayama G, Cheng P. Effects of interface wettability on microscale flow by
molecular dynamics simulation. International Journal of Heat and Mass
Transfer. 2004;47:501-13.
[117] Olgun U, Kalyon DM. Use of molecular dynamics to investigate polymer
melt–metal wall interactions. Polymer. 2005;46:9423-33.
[118] Tauviqirrahman M, Ismail R, Jamari, Schipper DJ. Optimization of the
complex slip surface and its effect on the hydrodynamic performance of two-
dimensional lubricated contacts. Computers & Fluids. 2013;79:27-43.
[119] Gohar R, Cameron A. Optical measurement of oil film thickness under
elastohydrodynamic lubrication. Nature. 1963;200:458-9.

119
[120] Cameron A, Gohar R. Theoretical and experimental studies of the oil film in
lubricated point contact. Proceedings of the Royal Society A: Mathematical,
Physical and Engineering Sciences. 1966;291:520-36.
[121] Foord CA, Wedeven LD, Westlake FJ, Cameron. Optical
elastohydrodynamics. Proceedings of the Institution of Mechanical Engineers.
1969-70;184:487-505.
[122] Roberts AD, Tabor D. Extrusion of liquids between highly elastic solids.
Proceedings of the Royal Society of Lond Series A. 1971;325:323.
[123] Luo JB, Wen SZ, Huang P. Thin film lubrication. Part I: study on the
transition between EHL and thin film lubrication using a relative optical
interference intensity technique. Wear. 1996;194:107-15.
[124] Guo F, Wong PL. A multi-beam intensity-based approach for lubricant film
measurement in non-conformal contacts. Proceedings of the Institution of
Mechanical Engineers, Part J: Journal of Engineering Tribology.
2002;216:281-91.
[125] Guo F, Wong PL. A wide range measuring system for thin lubricating film:
From nano to micro thickness. Tribology Letters. 2004;17:521-31.
[126] Hildebrand B, Haines K. Multiple-wavelength and multiple-source
holography applied to contour generation. Journal of the Optical Society of
America. 1967;57:155-7.
[127] Zelenka JS, Varner JR. A new method for generating depth contours
holographically. Applied Optics. 1968;7:2107-10.
[128] Zelenka JS, Varner JR. Multiple-index holographic contouring. Applied
Optics. 1969;8:1431-4.
[129] Wyant J. Testing aspherics using two-wavelength holography. Applied
Optics. 1971;10:2113-8.
[130] Polhemus C. Two-wavelength interferometry. Applied Optics.
1973;12:2071-4.
[131] Cheng YY, Wyant JC. Two-wavelength phase shifting interferometry.
Applied Optics. 1984;23:4539-43.
[132] de Groot PJ. Extending the unambiguous range of two-color interferometers.
Applied Optics. 1994;33:5948-53.

120
[133] van Brug H, Klaver RG. On the effective wavelength in two-wavelength
interferometry. Pure and Applied Optics: Journal of the European Optical
Society Part A. 1998;7:1465.
[134] Houairi K, Cassaing F. Two-wavelength interferometry: extended range and
accurate optical path difference analytical estimator. Journal of the Optical
Society of America. 2009;26:2503-11.
[135] Feilat EA. Detection of voltage envelope using prony analysis-Hilbert
transform method. IEEE Transactions on Power Delivery. 2006;21:2091-3.
[136] Bair S, Winer W. Shear strength measurements of lubricants at high pressure.
Journal of Tribology. 1979;101:251-7.
[137] Bair S, Winer W. The high pressure high shear stress rheology of liquid
lubricants. Journal of tribology. 1992;114:1-9.
[138] Sanchez-Reyes J, Archer LA. Interfacial slip violations in polymer solutions:
Role of microscale surface roughness. Langmuir. 2003;19:3304-12.
[139] Wu CW, Ma GJ. On the boundary slip of fluid flow. Science in China Series
G-Physics Mechanics & Astronomy. 2005;48:178-87.
[140] Kalin M, Polajnar M. The correlation between the surface energy, the contact
angle and the spreading parameter, and their relevance for the wetting
behaviour of DLC with lubricating oils. Tribology International.
2013;66:225-33.
[141] Wang DA, Wang XL, Liu XJE, Zhou F. Engineering a titanium surface with
controllable oleophobicity and switchable oil adhesion. Journal of Physical
Chemistry C. 2010;114:9938-44.
[142] Yaminsky VV. Molecular mechanisms of hydrophobic transitions. Journal of
Adhesion Science and Technology. 2000;14:187-233.
[143] Owens DK, Wendt R. Estimation of the surface free energy of polymers.
Journal of Applied Polymer Science. 1969;13:1741-7.
[144] Whyman G, Bormashenko E, Stein T. The rigorous derivation of Young,
Cassie–Baxter and Wenzel equations and the analysis of the contact angle
hysteresis phenomenon. Chemical Physics Letters. 2008;450:355-9.
[145] Extrand CW. Contact angles and their hysteresis as a measure of liquid-solid
adhesion. Langmuir. 2004;20:4017-21.
[146] Maoz R, Sagiv J. On the formation and structure of self-assembling
monolayers .1. a comparative atr-wetability study of langmuir-blodgett and
121
adsorbed films on flat substrates and glass microbeads. Journal of Colloid
and Interface Science. 1984;100:465-96.
[147] Thompson WR, Pemberton JE. Characterization of octadecylsilane and
stearic-acid layers on Al2O3 surfaces by Raman-Spectroscopy. Langmuir.
1995;11:1720-5.
[148] Shafrin EG, Zisman WA. Constitutive relation in the wetting of low energy
surfaces and the theory of the retraction method of preparing monolayers.
The Journal of Physical Chemistry. 1960;64:519–24.
[149] Fuks GI, Berlin LI. Perfluorinated aliphatic acids as additives regulating the
wetting of metals by oils. Chemistry and Technology of Fuels and Oils.
1971;7:305-8.
[150] Brochard F, Degennes PG. Shear-dependent slippage at a polymer solid
interface. Langmuir. 1992;8:3033-7.
[151] Ajdari A, Brochardwyart F, Degennes PG, Leibler L, Viovy JL, Rubinstein
M. Slippage of an entangled polymer melt on a grafted surface. Physica A:
Statistical Mechanics and its Applications. 1994;204:17-39.
[152] Spikes HA, Granick S. Equation for slip of simple liquids at smooth solid
surfaces. Langmuir. 2003;19:5065-71.
[153] Briscoe BJ, Evans DCB. The shear properties of Langmuir-Blodgett layers.
Proceedings of the Royal Society of London A: Mathematical, Physical and
Engineering Sciences. 1982;380:389-407.
[154] Wilson W, Huang X. Viscoplastic behavior of a silicone oil in a
metalforming inlet zone. Journal of Tribology. 1989;111:585-90.
[155] Kato K, Iwasaki T, Kato M, Inoue K. Evaluation of limiting shear stress of
lubricants by roller test. JSME International Journal, Series C. 1993;36:515-
22.
[156] Ostensen J, Wikstrom V, Hoglund E. Interaction effects between temperature,
pressure and type of base oil on lubricant shear strength coefficient.
Tribologia Finnish Journal of Tribology. 1993;11:123-32.

122
Appendix A: Envelope cycle of dichromatic interference

Suppose the two normalized intensities 𝐼1 (ℎ) and 𝐼2 (ℎ) be described by the

following equations,

4𝑛𝜋
𝐼1 (ℎ) = cos( ℎ + Φ1 ) (A1)
𝜆1

4𝑛𝜋
𝐼2 (ℎ) = cos( ℎ + Φ2 ) (A2)
𝜆2

Combining the two equations,

2𝑛 2𝑛 Φ1 + Φ2
𝐼1 (ℎ) − 𝐼2 (ℎ) = 𝐴(ℎ) × sin[ 𝜋( + )ℎ + ] (A3)
𝜆1 𝜆2 2

where

2𝑛 2𝑛 Φ1 − Φ2
𝐴(ℎ) = −2 sin[ 𝜋 ( − ) ℎ + ] (A4)
𝜆1 𝜆2 2

𝐴(ℎ) is the expression of the envelope, as shown in Fig. A.1. The cycle of the

envelope,

𝜆1 𝜆2 (A5)
𝑇=
𝑛|𝜆1 − 𝜆2 |

If only the positive amplitude is considered, the cycle can be expressed as:

𝑇 𝜆1 𝜆2
𝑇1 = = (A6)
2 2 2𝑛|𝜆1 − 𝜆2 |

Fig. A.1 Illustration of envelope expression

123
Appendix B: Ratio of changing rate

Suppose the change of the film thickness is ∆ℎ in ∆𝑡 seconds. Change in the fringe

order of monochromatic interference,

∆ℎ 2∆ℎ𝑛
∆𝑇𝜆1 = =
𝜆1 𝜆1 (B1)
2𝑛

The change of fringe order of the envelope is,

∆ℎ 2∆ℎ𝑛|𝜆1 − 𝜆2 |
∆𝑇𝜆𝑒 = =
𝜆1 𝜆2 𝜆1 𝜆2 (B2)
2𝑛|𝜆1 − 𝜆2 |

The ratio between the two changing rate is,

∆𝑇𝜆𝑒
𝑣𝑒𝑛𝑣𝑒𝑙𝑜𝑝𝑒 |𝜆1 − 𝜆2 |
= ∆𝑡 = (B3)
𝑣𝜆1 ∆𝑇𝜆1 𝜆2
∆𝑡

124
Appendix C: Flow chart of numerical analysis

Fig. C.1 Flow chart of numerical analysis

125
Appendix D: List of my publications

[1] Guo L., Guo F. and Wong P.L., “Boundary yield stress and interfacial
potential energy barrier in thin film hydrodynamic lubrication”, Submitted to
Tribology Letters, August 2015.

[2] Guo L., Wong P.L. and Guo F., “Correlation of contact angle hysteresis and
hydrodynamic lubrication”, Tribology Letters, Vol. 58, 45, pp. 1-9, June 2015.

[3] Liu H.C., Guo F., Guo L. and Wong P.L., “A dichromatic interference
intensity modulation approach to measurement of lubricating film thickness”,
Tribology Letters, Vol. 58:15, pp. 1-11, April 2015.

[4] Wong P.L., Guo L., Hiu H.C. and Guo F., “Parallel plane sliding bearing based
on surface heterogeneity by bovine serum albumin (BSA) aqueous solution”,
presented at the 2014 International Conference on Engineering Tribology
Technology, Sun Moon Lake, Nantou, Taiwan, Nov. 21-23, 2014. (Received
the best paper award)

[5] Guo L., Wong P.L., Guo F. and Liu H.C., “Determination of thin
hydrodynamic lubricating film thickness using dichromatic interferometry”,
Applied Optics, 53, 26, Sep. 2014.

[6] Guo F., Zang S., Guo L. and Wong P.L., "Can a liquid drop lubricate?", 7th
China International Symposium on Tribology, Xuzhou, PRC, 27-30 April 2014,
pp 127-129.

[7] Guo L., Wong P.L. and Guo F., "Contact angle hysteresis effect on
hydrodynamic lubrication", 7th China International Symposium on Tribology,
Xuzhou, PRC, 27-30 April 2014, pp 138-140.

[8] Guo L, Wong P.L. and Guo F., "Boundary effect on hydrodynamic lubricated
contact", World Tribology Congress 2013, Torino, Italy, 8-13 September 2013.

[9] Guo L., Yip L.K., Wong P.L. and Guo F., “Effects of wettability on
hydrodynamic lubrication”, 2013 National Youth Conference on Tribology,
Qingdao, China, E017, pp.556-559, 2-4 June 2013 (In Chinese).

126

You might also like