You are on page 1of 9

Tribology International 77 (2014) 97105

Contents lists available at ScienceDirect

Tribology International
journal homepage: www.elsevier.com/locate/triboint

Friction and wear performance of laser peen textured surface


under starved lubrication
Kangmei Li a, Zhenqiang Yao a,n, Yongxiang Hu a, Weibin Gu b
a
b

State Key Laboratory of Mechanical System and Vibration, School of Mechanical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
Shanghai Aircraft Design and Research Institute, Shanghai 201210, China

art ic l e i nf o

a b s t r a c t

Article history:
Received 28 January 2014
Received in revised form
3 April 2014
Accepted 12 April 2014
Available online 23 April 2014

Laser peen texturing (LPT), which is a new surface texturing technology, was proposed to fabricate
micro-dimple arrays on copper surface. Pin-on-disk experiments were conducted under different normal
loads and sliding speeds to investigate friction and wear behaviors of untextured and textured samples
in starved lubrication. It was found that friction performance of the surface is improved signicantly
after LPT. Microscope observation and EDS analysis showed that textured surface could reduce both
abrasive and adhesive wear as compared to untextured surface. Results also indicated that an optimum
texture density might exist at which surface shows the best friction and wear behavior.
& 2014 Elsevier Ltd. All rights reserved.

Keywords:
Surface texturing
Laser processing
Friction test
Starved lubrication

1. Introduction
Surface texturing as an approach for controlling the friction and
wear behavior of tribological components is well known for many
years. As early as 1966, the effect of surface texturing on lubrication was studied by Hamilton et al. [1]. Now surface texturing has
been widely recognized as a viable means to reduce friction
coefcient, improve load carrying capacity and wear resistance
of tribological systems. Micro-dimple array is a typical kind of
texture and plays an important part in improving tribological
performance. In starved lubrication, dimples act as reservoirs of
lubricant, which supply back-up lubricant when lubricant lm is
broken down [2]. In hydrodynamic lubrication, dimples are
utilized as micro-hydrodynamic bearings to help improve load
carrying capacity of the lubricant lm [3]. Dimples could also serve
as traps for wear debris in both lubricated and dry sliding [4].
Many studies on various forms, sizes and shapes of microdimples on sliding surfaces for tribological applications have been
reported over the last two decades [2,511]. Ryk et al. [2] studied
the effect of surface texturing under starved lubrication conditions, and found that with optimum dimple depth and low
lubricant viscosity the texturing was benecial over the entire
range of tested ow rates. However, with the deepest dimples or

n
Correspondence to: Room 726, Building A, 800 Dongchuan Road, Minhang
District, Shanghai 200240, China. Tel./fax: 86 21 34206315.
E-mail addresses: lkm718@gmail.com (K. Li), zqyaosjtu@gmail.com (Z. Yao),
huyx@sjtu.edu.cn (Y. Hu), guweibin@comac.cc (W. Gu).

http://dx.doi.org/10.1016/j.triboint.2014.04.017
0301-679X/& 2014 Elsevier Ltd. All rights reserved.

with high lubricant viscosity the surface texturing may be detrimental under certain operating conditions. Galda et al. [8] studied
the effect of surface texturing on lubrication regime transitions
from mixed to hydrodynamic and found that the shape and
distribution of oil pockets are the main factors affecting the
lubrication kinds. Higuera Garrido [10] studied tribological behavior of laser-textured NiCrBSi coatings. Results demonstrated the
strong correlation among texture density, dimple diameter and
contact area for reduction of the friction coefcient. Tang et al. [11]
fabricated multi-dimples using a miniature engraving and studied
the effect of surface texturing on friction and wear under hydrodynamic lubrication, they concluded that a 5% optimal dimple area

Fig. 1. Schematic of laser peen texturing.

98

K. Li et al. / Tribology International 77 (2014) 97105

Table 1
Texture parameters used in this study.
Specimen no.

Texture density, t (%)


Dimple depth, h (m)
Dimple diameter, d (m)

5
10
1100

13
10
1100

35
10
1100

Fig. 2. Schematic of micro-dimple array.

fraction can generate the greatest hydrodynamic pressure compared with other fractions and can reduce friction and wear up to
38% and 72%, respectively.
Nowadays, methods to fabricate surface textures could be
generally classied into mechanical [1216], ion beam texturing
[17,18], etching [1922] and energy beam techniques [2329].
Among these methods, laser surface texturing (LST) is considered
as the most promising texturing technology. The main reason is
that textures fabricated by LST could be precisely controlled and
this process is friendly to the environment [7]. In last few years,
many studies have focused on LST, and it has been successfully
applied on piston rings [24,25], mechanical seals [26,27], hydrostatic gas seals [28], journal bearings [29], thrust bearings [30],
and soft elasto-hydrodynamic lubrication [23,31,32]. However,
laser peen texturing (LPT), which is another energy beam technology for surface texturing, has been given little attention.
Compared with LST, besides the merits of LST mentioned
above, LPT has more advantages: Firstly, LPT utilizes laser-shockinduced mechanical effect rather than thermal effect, so it could
effectively avoid the negative inuence of ablation in material
surface integrity. However, during LST, micro-dimples are fabricated by a laser ablation mechanism. Tensile residual stress may
occur in material due to the high temperature during ablation [33].
This change in surface integrity will shorten the fatigue life of the
material [34]; Secondly, after LPT, there is no apparent pile-up
around the micro-dimple which is generally detrimental to the
tribological performance, thus post-texturing lapping process for
the purpose of removing pile-up could be omitted. Nevertheless,
the lapping process is usually needed for LST according to the
studies of Etsion [8], Yu et al. [27] and Kovalchenko et al. [35].
Thirdly, the great shock pressure of LPT can induce deep compressive residual stress and hardened layer in material surface and
subsurface, which can improve fatigue life of the material dramatically [36,37]. The disadvantage of LPT compared with LST may be
that the equipments and the process procedures of LPT are more
complex.

Fig. 3. Surfaces of the specimens.

K. Li et al. / Tribology International 77 (2014) 97105

In this study, LPT was adopted to fabricate micro-dimples on


copper surfaces, and the main purpose of this paper is to
investigate the friction behavior and wear performance of laser
peen textured surfaces under different normal loads and rotational
speeds in starved lubrication.

2. Experimental details
2.1. Specimen preparation
All textured specimens in this study were processed by LPT.
The schematic of LPT is shown in Fig. 1. The surface to be textured
is rst coated with an opaque coating (such as black insulation
tape) and then covered by a transparent overlay (such as owing
water). When the laser with high power density (typically GW/
cm2) irradiates on the specimen surface, opaque coating melts and
vaporizes, and then forms the steam particles. Steam particles
continue absorbing the laser energy, and then ionize to hightemperature and high-density plasma. Plasma expands rapidly in
the direction away from the specimen and is trapped between the
sample and the transparent overlay, creating an extremely high
pressure which propagates into the material as a shockwave [38].
When the pressure of the shockwave which is on the order of GPa
exceeds the dynamic yield strength of the material, a micro-

99

dimple is fabricated. From the above process, it can be seen that


LPT is a purely mechanical process, which can avoid heat effect of
the laser.
LPT experiments were performed using a Q-switched Nd:YAG
laser with a wavelength of 532 nm and the Full Width at Half
Maximum (FWHM) of the pulse was about 10 ns. The frequency of
the laser was 10 Hz. All the textured specimens were processed
under the same LPT parameters: The laser spot diameter was
1.3 mm, the repeated shock number for every micro-dimple was
1 and the laser power density was 6.76 GW/cm2.
Specimens were prepared from Oxygen-Free High Conductivity
(OFHC) copper plate with dimensions of 50 mm  50 mm  2 mm
(length  width  thickness). The hardness of the OFHC copper
plate was 97 HV (with 25 g load and 20 s hold time) and the
average roughness of the surface was Ra 0.14 m. Table 1 lists the
texture parameters used in this research. Specimens from no.1 to
no. 3 were textured with different texture densities, while specimen of no. 0 was an untextured surface used for comparison with
textured ones.
Under the assumption that d is the diameter of the dimple and
l is the side length of the micro-dimple unit (see Fig. 2), then the
texture density t can be expressed as

Sd
d
 100% 2  100%
St
4l
2

where Sd is the area of the dimple, Sd d =4, and St is the area of


2
the micro-dimple unit, Sd l .
As shown in Fig. 3, the geometric and dimensional characteristics of micro-dimples show good repeatability, indicating that
LPT is an effective method to fabricate controlled micro-dimple
array. The 3D topography and 2D prole of typical micro-dimple
observed by 3D optical surface prolometer (KS-1100, produced by
Japanese Keyence company) are shown in Fig. 4.
2

2.2. Friction and wear tests


Friction and wear tests were both conducted using a CETR UMT-3
tribometer and the pin-on-disk test mode was chosen. Fig. 5(a) and
(b) shows the tribometer and the schematic diagram of the contact
conguration, respectively. The pin was a 10 mm diameter GCr15
steel with the hardness of 940 HV (with 25 g load and 20 s hold
Table 2
Properties of the lubricant used in this study.

Fig. 4. Topography and prole of single micro-dimple.

Parameters

Units

Values

ISO viscosity grade


Flash point
Pour point
Viscosity in 40 1C
Viscosity in 100 1C
Density in 40 1C

1C
1C
mm2/s
mm2/s
kg/m3

32
208
 27
32
5.3
873

Fig. 5. Setup used for friction and wear tests. (a) UMT-3 tribometer and (b) schematic diagram of contact conguration.

100

K. Li et al. / Tribology International 77 (2014) 97105

Table 3
Parameters of friction and wear tests.
Parameters

Rotational speed
Sliding speed
Normal load
Contact pressure
Loading rate
Temperature

Units

rpm
m/s
N
MPa
N/s
1C

Values
Friction test 1

Friction test 2

Wear test

150
0.08
20400 (linearly increased)
0.265.10
1.27, 2.54
207 2

50900 (step increased)


0.030.47
50
0.65

207 2

150
0.08
20280 (linearly increased)
0.263.64
1.3
207 2

Fig. 6. Typical friction coefcient curve under linearly increased normal load.

Fig. 8. Effect of normal load on friction coefcient with loading rate of 2.54 N/s.

Fig. 7. Effect of normal load on friction coefcient with loading rate of 1.27 N/s.

time). The stationary pin was pressed at the required load and slided
against the rotating disk textured by LPT. The rotational radius of the
pin was set as 5 mm. Before the tribilogical test, the sliding surface of
the pin was polished to mirror smoothness with the average roughness of 0.05 m. After polishing, the pin and the disk were cleaned in
acetone solution by ultrasonic cleaning to remove the metal fragments and other attachments on metal surfaces.
Lubricant used in all tests was Mobil Vacuoline 1405 lubricant.
This lubricant is widely used on guide and chute of machine tools
because of its good performance in eliminating stick-slip and chatter
during sliding. The properties of the lubricant are listed in Table 2.
In this research, two sets of friction tests were designed. The rst
set was performed with constant rotational speed and linearly
increased normal load under two different loading rates. The second
set was carried out with constant normal load and step increased

rotational speed. After friction tests, wear tests were conducted to


study the wear resistance of the specimens. The detailed experimental parameters used in this study are summarized in Table 3.
Before each friction and wear test, sufcient lubricant was dropped
on the contact area, and excessive lubricant was removed from the
specimen by using a rubber blade. After the removal of lubricant,
only a thin layer of lubricant appeared on the surface of untextured
specimen. For textured specimens, due to the existence of microdimples, the major proportion of lubricant was stored in the dimples
and a minor proportion remained on the untextured areas. Due to
the difference of texture density, lubricant stored in dimples was
different while the thickness of lubricant above the untextured area
was approximately the same. As a result, it may be considered that
the lubrication condition for each specimen was approximately the
same except the lubricant volume stored in dimples. In this way, the
different friction performance among different specimens could
reect the effect of texture density. Each friction test was conducted
three times to minimize data scattering.

3. Results and discussion


3.1. Friction behavior under different normal loads
During friction test 1, the normal load increased linearly from
20 N to 400 N with the loading rate of 1.27 N/s and 2.54 N/s, while
the rotational speed kept constant of 150 rpm. For all cases in this
test, the trends of the friction coefcient and normal load curves
are very similar. Typical curves of friction coefcient and normal
load are shown in Fig. 6. It is found that the coefcient of friction
keeps above 0.1 during the whole test. For this reason, and due to

K. Li et al. / Tribology International 77 (2014) 97105

the low rotational speed as well as the starved lubrication, it can


be assumed that the sliding contact in this test was boundary
lubricated. Moreover, the variation of friction coefcient could be
divided into two stages. During the rst stage, the friction
coefcient seems insensitive to the increase of normal load, thus
this stage can be called stable stage. When normal load reaches a
certain value, however, friction coefcient increases rapidly with
obvious noise and vibration, indicating that the lubricant lm
collapses and the lubrication fails, therefore, this stage is called
failure stage. During the friction test, the time at which the friction
coefcient increases sharply is dened as the failure time, and the
corresponding normal load at the failure time is called the
maximum normal load (as marked in Fig. 6) which can reect
the load carrying capacity of the surface.
The variations of friction coefcient under loading rate of
1.27 N/s are shown in Fig. 7. There are two points worth noting:

101

transverse force for the relative sliding movement and consequently causes high friction coefcient. Moreover, too high
texture density indicates limited contact area and great contact pressure, therefore leads to large deformation and small
gap between contact surfaces, which makes the lubricant lm
thin and easy to collapse.

(1) With the linear increase of normal load, the untextured surface failed earlier than the textured ones, and the friction
coefcient of the untextured surface in stable stage is relatively high. The result demonstrates that surface texturing
helps to improve friction performance of boundary-lubricated
surface. It probably benets from the so-called secondary
lubrication effect. That is, lubricant reserved in the dimples
acts as a secondary source of lubricant and could be drawn
into the interface during the relative sliding movement.
(2) Surface with texture density of 13% (abbreviated as surface
tex13%) exhibits the best friction performance as compared to
surfaces with texture densities of 5% and 35% (abbreviated as
surface tex5% and tex35%, respectively). In other words, surface tex13% has the lowest friction coefcient and the longest
failure time. This result suggests that although the surface
with micro-dimples is helpful for the improvement of tribological performance compared to untextured, it does not mean
that larger texture density leads to the better tribological
performance. There might be an optimum texture density,
which leads to the best tribological performance under
boundary lubrication. Compared to the surface with appropriate texture density, the surface with too low or too high
texture density may lead to relatively high friction coefcient
and short failure time. Too low texture density means small
number of dimples with very limited lubricant and long
distance among dimples. Therefore, the lubricant volume
might be inadequate to cover the whole surface, leading to
high coefcient of friction and short failure time. On the other
hand, the surface with too high texture density requires great

Fig. 9. Effect of loading rate on friction coefcient.

Fig. 10. Effect of loading rate on friction coefcient, failure time and maximum
normal load. (a) Friction coefcient, (b) failure time and (c) maximum normal load.

102

K. Li et al. / Tribology International 77 (2014) 97105

Fig. 11. Typical friction coefcient curve under step increased rotational speed.

Fig. 8 shows the effect of normal load with loading rate of


2.54 N/s on friction coefcient. It is found that surface tex13% and
untextured surface show the best and worst friction performance,
respectively. Furthermore, it could be seen from Fig. 8 that the
friction process of surface tex13% is much more stable than those
of other surfaces.
In order to compare the effect of loading rate on friction
performance, the variation of the friction coefcient under two
loading rates are shown in one gure as shown in Fig. 9. It is found
that the general trend of the friction coefcient variations under
loading rate of 2.54 N/s and 1.27 N/s are similar. For a given texture
density, high loading rate leads to large friction coefcient and
short failure time. In addition, since high loading rate is more
likely to result in an unstable friction process, the oscillation of
friction coefcient under loading rate of 2.54 N/s is stronger than
that under loading rate of 1.27 N/s especially for the surface tex5%,
tex35% and the untextured surface.
Quantitative comparisons of friction coefcients, failure times
as well as maximum normal loads for different texture densities
under different loading rates are shown in Fig. 10. The data
in Fig. 10(a) are obtained by averaging friction coefcients in
stable stage. It is seen that for surface tex13% with loading rate
of 1.27 N/s, the friction coefcient is reduced most signicantly by
more than 1/2 compared with untextured surface. Fig. 10(b) shows
that the failure time of surface tex13% is about 3 times longer than
that of untextured surface under loading rate of 1.27 N/s, and
when the loading rate increases to 2.54 N/s, this multiple becomes
as high as 5 times. Fig. 10(c) presents the maximum normal loads
gained from the tting curves of raw data. It is found that load
carrying capacity of the surface tex13% is improved by nearly
4 times than that of the untextured one when the loading rate is
2.54 N/s. Moreover, it can also be seen from Fig. 10(c) that for a
certain texture density, the maximum load carrying capacity
was smaller at bigger loading rate. For this reason, as shown in
Fig. 10(b), the failure time of each surface is shortened several times
when the loading rate increases twice (from 1.27 N/s to 2.54 N/s).
3.2. Friction behavior under different rotational speeds
In friction test 2, the rotational speed was step increased from
50 rpm to 900 rpm and the normal load was kept constant at 50 N.
Friction coefcient curves of all specimens in this test have similar

Fig. 12. Schematic of starved lubrication. (a) Under low sliding speed and (b) under
high sliding speed.

trend. So only a typical variation of friction coefcient with step


increased rotational speed is shown in Fig. 11. At the beginning of
the test, the friction coefcient is about 0.1 when the rotational
speed is 50 rpm, as the rotational speed increases to 900 rpm, the
friction coefcient reduced to about 0.05. According to the range of
the friction coefcient, it could be deduced that the lubrication
regime remains boundary and mixed lubrication (that is starved
lubrication), rather than transferring to hydrodynamic lubrication
[39]. From the trend of the curve, it is seen that the coefcient of
friction becomes smaller and the oscillating of the curve becomes
weaker progressively with the increase of rotational speed. This
trend is in good correlation with the typical Stribeck curve.
For starved lubrication, as shown in Fig. 12(a), the lubricant lm
is too thin to separate the two solid surfaces completely, so some
opposing asperities touch each other. The friction coefcient now
depends on both the solidliquid friction and the solidsolid
friction [39]. By assuming that A represents the total real contact

K. Li et al. / Tribology International 77 (2014) 97105

area between two sliding surfaces, friction force F could be


given by
F Aw s 1  w L 

where w is the ratio of the solid contact area Am to the total real
contact area A, w Am =A. s is the shear strength of solid surface,
and L is the shear strength of uid surface. Then the friction
coefcient of starved lubrication f SL could be obtained
f SL

F Aw s 1  w L  Aw s  L L 

N
N
N

where N represents the normal load which is a constant in this


test. From Eq. (3), it can be seen that A, s , L and N are all
constants here, so there exists proportional relationship between
fSL and w .

103

With the increase of rotational speed, as shown in Fig. 12(b),


the lubricant lm becomes thick (from h1 to h2) and less asperities
touch each other, resulting in small area of solidsolid contact and
small value of w [39,40]. As a result, the friction coefcient of
starved lubrication fSL shows a gradual decreasing trend in Fig. 11.
Effect of texture density on the friction coefcient under step
increased rotational speed and constant normal load is shown in
Fig. 13. It is seen that for all specimens tested, friction coefcients
decrease with the increase of rotational speed, but the friction
coefcient of the untextured surface is higher than that of textured
surface. Moreover, the smallest friction coefcient is also obtained
by the surface tex13%, followed by the surface tex35% and tex5%.
These results are consistent with those under the constant rotational speed and linearly increased normal load in friction test 1.
Therefore, it is conrmed that texture density of 13% is the best
among the three texture densities in this study.

3.3. Analysis of worn surfaces

Fig. 13. Effect of texture density on friction coefcient under step increased
rotational speed.

Micrographs in Fig. 14(a)(d) shows the worn surfaces of the


untextured surface, surface tex5%, surface tex13% and surface
tex35%, respectively. The micrographs were taken by a highresolution true color confocal microscope (Axio CSM 700, produced by German Carl Zeiss company). The worn surfaces were
obtained by the wear tests with the parameters listed in Table 3.
Fig. 14(a) shows that the original surface of the untextured specimen almost worn out. Deep scratches and material loss could be
clearly observed, indicating that mixed wear consisting of abrasive
wear and adhesive wear occurred during the test. For untextured
surface, it is difcult to reserve debris and lubricant. On one hand, the
debris generated by friction can cause micro-cutting on the untextured
surface and lead to abrasive wear. On the other hand, the lubrication
without back-up lubricant fails easily and the contact area of two
mating surfaces increases signicantly, resulting in a lot of frictional
heat that is the major cause of adhesive wear. In addition, the lost or

Fig. 14. Optical microscope photographs of worn surfaces with different texture densities. (a) Untextured, (b) tex5%, (c) tex13% and (d) tex35%.

104

K. Li et al. / Tribology International 77 (2014) 97105

transferred material produced by adhesive wear will scratch the


surface further. Thus, the scratches on the surface become much
wider and deeper.
As shown in Fig. 14(b), abrasive and adhesive wear can also be
observed from the topography of worn surface tex5%, which is similar
to that of the untextured surface, while the width and depth of
scratches as well as the area of adhesive wear seem smaller than those
on the untextured surface. These results suggest that micro-dimples
inuenced on wear resistance, but the effect was insignicant due to
the limited number of micro-dimples.
For surface tex13%, the increase of texture density improves the
ability to reserve debris and lubricant. As a result, only several
shallow scratches were observed on the worn surface tex13%, and
some original surface topography can still be observed in Fig. 14(c).
As the texture density increases to 35%, there are more dimples
to reserve lubricant that help cool the friction surfaces and
contribute to the reduction of friction heat generation, this is
probably the main reason for the decrease of adhesive wear for
surface tex35% when compared with surface tex5%, as shown in
Fig. 14(d) and (b). With the increase of texture density, however,
contact area of mating surfaces is reduced simultaneously, giving
rise to the contact pressure that causes more severe abrasive and
adhesive wear. So when compared with surface tex13%, there are
more deeper scratches on the surface tex35%.

Table 4
Composition analysis results of worn surfaces.
Element

C
O
Fe
Cu

Untextured surface

Textured surface

Weight percent

Atomic percent

Weight percent

Atomic percent

23.62
2.72
0.56
73.10

59.66
5.15
0.30
34.89

27.06
1.47

71.47

64.93
2.64

32.43

In addition, it is worth noticing that the diameters of the microdimples on three textured surfaces (see Fig. 14(b)(d)) are different which can also reect the level of wear. Since the original
diameter of the dimples before wear test are almost the same, the
larger diameter of the dimple after wear test means relatively
slighter wear. This phenomenon conrms that surface tex13% has
the best wear resistance in the present study.
In order to further investigate the difference of chemical
compositions of specimen surfaces after wear tests, the energy
dispersive spectrometer (EDS) analysis was performed on wear
tracks of worn surfaces by utilizing the EDS system of scanning
electron microscope (SEM). The EDS results showed that spectra of
textured surfaces with three different texture densities were
similar. Typical spectra of untextured surface and textured surfaces
are shown in Fig. 15(a) and (b), respectively. The weight percent
and atomic percent of the elements are listed in Table 4.
From Fig. 15 and Table 4, it is seen that elements of C, O and Cu are
detected on both untextured and textured surfaces. The element with
the highest weight is Cu, which comes from the substrate material of
OFHC copper. The existence of elements C and O is mainly due to the
residue of lubricant oil on worn surfaces. In addition to this, a small
amount of element Fe is also detected on the worn untextured surface
as shown in Fig. 15(a), while no other elements are found on the
textured surface as shown in Fig. 15(b). Since the hardness of the pin is
much greater than that of OFHC copper, the material transfer from the
pin to OFHC copper surface is difcult to occur. As a result, the transfer
of the Fe element indicates that wear of the untextured surface is
more severe than the textured surfaces.

4. Conclusions
Micro-dimple arrays with different texture densities were
fabricated on OFHC copper plates by laser peen texturing (LPT).
The friction and wear performances of textured surfaces and
untextured surfaces were studied in starved lubrication under
the condition of linearly increased normal load and step increased
rotational speed. The following conclusions could be drawn:

Fig. 15. EDS results of worn surfaces. (a) Untextured surface and (b) textured
surface.

(1) It is veried that LPT is an effective technology to fabricate


controlled micro-dimple array on specimen surface.
(2) Surfaces textured by LPT have better friction performance and
help to reduce both abrasive and adhesive wear as compared
to untextured surface within the considered range of normal
loads and rotational speeds in this research.
(3) Texture density has strong effect on the friction and wear
behavior. It is found that friction coefcient, failure time, load
carrying capacity and wear resistance do not vary monotonically with the texture density. There might exist an optimal
texture density at which the textured surface exhibits the best
tribological performance.
(4) Surface with texture density of 13% shows the best friction and
wear performance in this study. Compared with the untextured
surface, it could reduce friction coefcient up to about 1/2
under the test conditions in this research.

K. Li et al. / Tribology International 77 (2014) 97105

Our future work will focus on the qualitative and quantitative


comparison between LPT and LST and further investigate the
advantages and disadvantages of LPT in the application of surface
texturing.

Acknowledgments
The authors would like to thank the National Natural Science
Foundation of China (Grant nos. 51075271 and 51375305), Foundation for Innovative Research Groups of the National Natural
Science Foundation of China (Grant no. 51121063) and Innovation
Fund of the National Commercial Aircraft Manufacturing Engineering Technology Research Center (Grant no. SAMC12-JS-15-025).
References
[1] Hamilton DB, Walowit JA, Allen CM. A theory of lubrication by microirregularities. J Basic Eng 1966;88:17784.
[2] Ryk G, Kligerman Y, Etsion I. Experimental investigation of laser surface texturing
for reciprocating automotive components. Tribol Trans 2002;45:4449.
[3] Etsion I, Kligerman Y, Halperin G. Analytical and experimental investigation of
laser-textured mechanical seal faces. Tribol Trans 1999;42:5116.
[4] Etsion I. State of the art in laser surface texturing. Tribol-T ASME 2005;125:24853.
[5] Wakuda M, Yamauchi Y, Kanzaki S, Yasuda Y. Effect of surface texturing on
friction reduction between ceramic and steel materials under lubricated
sliding contact. Wear 2003;254:35663.
[6] Andersson P, Koskinen J, Varjus S, Gerbig Y, Haefke H, Georgiou S, et al.
Microlubrication effect by laser-textured steel surfaces. Wear 2007;262:36979.
[7] Etsion I. Improving tribological performance of mechanical components by
laser surface texturing. Tribol Lett 2004;17:7337.
[8] Galda L, Pawlus P, Sep J. Dimples shape and distribution effect on characteristics of Stribeck curve. Tribol Int 2009;42:150512.
[9] Vrbka M, amnek O, perka P, Nvrat T, Kupka I, Hartl M. Effect of surface
texturing on rolling contact fatigue within mixed lubricated non-conformal
rolling/sliding contacts. Tribol Int 2010;43:145765.
[10] Garrido AH, Gonzlez R, Cadenas M, Battez AH. Tribological behavior of lasertextured NiCrBSi coatings. Wear 2011;271:92533.
[11] Tang W, Zhou Y, Zhu H, Yang H. The effect of surface texturing on reducing the
friction and wear of steel under lubricated sliding contact. Appl Surf Sci
2013;273:199204.
[12] Nakatsuji T, Mori A. The tribological effect of mechanically produced microdents by a microdiamond pyramid on medium carbon steel surfaces in rolling
sliding contact. Mechanica 2002;66:66374.
[13] Pettersson U, Jacobson S. Friction and wear properties of microtextured DLC
coated surfaces in boundary lubricated sliding. Tribol Lett 2004;17:5539.
[14] Bulatov VP, Krasny VA, Schneider YG. Basics of machining methods to yield
wear- and fretting-resistive surfaces, having regular roughness patterns. Wear
1997;208:1327.
[15] Zhang YD, Lin JQ, Fu QL, Hu HY. Measuring and controlling of arbor
displacement in low frequency vibration machining surface for micropits.
JZUS-A 2008;42:14104.

105

[16] Friedrich CR. Micromechanical machining of high aspect ratio prototypes.


Microsyst Technol 2002;8:3437.
[17] Zhou L, Kato K, Umehara N, Miyake Y. Nanometre scale island-type texture
with controllable height and area ratio formed by ion-beam etching on harddisk head sliders. Nanotechnology 1999;10:363.
[18] Cui FZ, Luo ZS. Biomaterials modication by ion-beam processing. Surf Coat
Technol 1999;112:27885.
[19] Wang XL, Kato K. Improving the anti-seizure ability of SiC seal in water with
RIE texturing. Tribol Lett 2003;14:27580.
[20] Wang XL, Adachi K, Otsuka K, Kato K. Optimization of the surface texture for
silicon carbide sliding in water. Appl Surf Sci 2006;253:12826.
[21] Stephens LS, Siripuram R, Hyden M, McCartt B. Deterministic micro asperities
on bearings and seals using a modied LIGA process. J Eng Gas Turbines Power
(Trans ASME) 2004;126:14754.
[22] Pettersson U, Jacobson S. Inuence of surface texture on boundary lubricated
sliding contacts. Tribol Int 2003;36:85764.
[23] Shinkarenko A, Kligerman Y, Etsion I. The effect of surface texturing in soft
elasto-hydrodynamic lubrication. Tribol Int 2009;42:28492.
[24] Ryk G, Etsion I. Testing piston rings with partial laser surface texturing for
friction reduction. Wear 2006;261:7926.
[25] Etsion I, Halperin G, Becker E. The effect of various surface treatments on
piston pin scufng resistance. Wear 2006;261:78591.
[26] Etsion I, Burstein L. A model for mechanical seals with regular micro-surface
structure. Tribol Trans 1996;39:67783.
[27] Yu XQ, He S, Cai RL. Frictional characteristics of mechanical seals with a lasertextured seal surface. J Mater Process Technol 2002;129:4636.
[28] Feldman Y, Kligerman Y, Etsion I. A hydrostatic laser surface textured gas seal.
Tribol Lett 2006;22:218.
[29] Brizmer V, Kligerman Y. A laser surface textured journal bearing. J Tribol
2012;134:031702.
[30] Brizmer V, Kligerman Y, Etsion I. A laser surface textured parallel thrust
bearing. Tribol Trans 2003;46:397403.
[31] Shinkarenko A. Kligerman Y. EtsionI I. Partial elastomer texturing in soft elasto
hydrodynamic lubrication, In: Proceedings of the STLE/ASME international
joint tribology conference; 2008. p. 2879.
[32] Shinkarenko A, Kligerman Y, Etsion I. The effect of elastomer surface texturing
in soft elasto-hydrodynamic lubrication. Tribol Lett 2009;36:95103.
[33] Iordanova I, Antonov V, Gurkovsky S. Changes of microstructure and mechanical properties of cold-rolled low carbon-steel due to its surface treatment by
Nd: glass pulsed laser. Surf Coat Technol 2002;153:26775.
[34] Guo YB, Caslaru R. Fabrication and characterization of micro dent arrays
produced by laser shock peening on titanium Ti6Al4V surfaces. J Mater
Process Technol 2011;211:72936.
[35] Kovalchenko A, Ajayi O, Erdemir A, Fenske G. Friction and wear behavior of laser
textured surface under lubricated initial point contact. Wear 2011;271:171925.
[36] Montross CS, Wei T, Ye L, Clark G, Mai YW. Laser shock processing and its
effects on microstructure and properties of metal alloys: a review. Int J Fatigue
2002;24:102136.
[37] Nalla RK, Altenberger I, Noster U, Liu GY, Scholtes B, Ritchie RO. On the
inuence of mechanical surface treatments-deep rolling and laser shock
peeningon the fatigue behavior of Ti6Al4V at ambient and elevated
temperatures. Mater Sci Eng A 2003;355:21630.
[38] Tan Y, Wu G, Yang JM, Pan T. Laser shock peening on fatigue crack growth
behavior of aluminum alloy. Fatigue Fract Eng Mater Struct 2004;27:64956.
[39] Wen SZ, Huang P. Principles of tribology. 4th ed.. Beijing: Tsinghua University
Press; 2012.
[40] Szeri AZ. Fluid lm lubrication: theory and design. 1st ed.. Cambridge:
Cambridge University Press; 2005.

You might also like