Tribological Behavior of Multi-Scaled Patterned Surfaces Machined Through Inclined End Milling and Micro Shot Blasting

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Tribology Letters (2018) 66:132

https://doi.org/10.1007/s11249-018-1086-y

ORIGINAL PAPER

Tribological Behavior of Multi-scaled Patterned Surfaces Machined


Through Inclined End Milling and Micro Shot Blasting
Jesus Resendiz1 · Philip Egberts1   · Simon S. Park1

Received: 22 May 2018 / Accepted: 8 September 2018


© Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
Surface texturing is one mechanism that friction coefficients in both dry and lubricated contacts can be reduced compared
with untextured, flat surfaces. End milling and shot blasting are two processes used to produce surface textures, includ-
ing monolithic textures that have one type of surface feature, as well as multi-scale roughness features when two texturing
processes are sequentially used. In such surfaces, we have observed that surface texturing decreases the measured friction
coefficient under lubricated conditions, and that greater reductions in the friction coefficient are observed for those surfaces
that had been both end milled and shot blasted. Simulations replicating the experiments suggest that the greatest factor con-
tributing to the reduced friction observed for the textured surfaces is a result of increased fluid pressure in the contact region
resulting from cavitation of the lubricant. However, a substantial decrease in the depth of the dimples on worn surfaces was
also observed, suggesting that entrapment of wear particles within the surface texture features may also influence the meas-
ured friction coefficient. Alongside friction measurements, analysis of the wear track depth showed that surface texturing
also has a beneficial influence on the calculated Archard wear coefficient.

Keywords  Multi-scaled micro surface patterns · Diamond dimple machining · Shot blasting · Reciprocating tribometer ·
Surface texturing · Oil lubrication

1 Introduction Recently, it has been shown that end milling may be a lower
cost alternative that can be used to texture surface with a
Engineered surfaces, also referred to as functional surfaces, wider range of surface patterns [6–8]. However, most meth-
consist of repeated micro-scale patterns that can dramati- ods used for texturing surfaces result in a single feature size
cally alter the physical properties of the surface beyond over entire manufactured parts [5].
those related its intrinsic material properties. These surfaces Examination of biological surfaces as inspiration for the
have been the subject of much research [1], with applications design of new engineering materials indicates that hierarchi-
including wear reduction and enhanced lubrication [2], super cal surfaces may be more efficient than monolithic surfaces
nucleated boiling [3], adhesion modification, and optical dif- at reaching their engineering goals [9]. Thus, the objective
fusers [4]. Many methods exist for the fabrication of surface of this research will be to determine the feasibility of both
patterns, being the most popular laser surface texturing [5]. the fabricating hierarchical surfaces and evaluating their per-
formance in engineering systems, specifically as a mecha-
nism to improve lubrication of sliding contacts. Multi-scale
* Philip Egberts structures can be achieved through a number of different
philip.egberts@ucalgary.ca
manufacturing processes. Large-scale surface textures pro-
Jesus Resendiz duced through micro end milling has several advantages over
jdejresen@ucalgary.ca
other techniques used to produce surface textures, including
Simon S. Park laser ablation [10], electrical discharge [11], and focused ion
simon.park@ucalgary.ca
beam milling [12]. Specifically, micro end milling does not
1
Micro Engineering Dynamics and Automation Laboratory require an expensive, purpose-built facility to produce the
(MEDAL), Department of Mechanical & Manufacturing surface textures on metals and other hard materials and can
Engineering, University of Calgary, 2500 University Dr. also be used to produce a number of texture shapes without
NW, Calgary, AB T2N 1N4, Canada

13
Vol.:(0123456789)
132   Page 2 of 13 Tribology Letters (2018) 66:132

significant re-tooling. Subsequent passes of the textured sur- 2 Experiments


faces can result in a hierarchical structure should the later
texturing processes result in surface alterations that have a Aluminum 6061-T6 workpieces with plan view dimensions
significant length-scale advantage over the proceeding tex- of 50 by 50 mm2 were used in all friction experiments. In
turing process or can be controlled to have this length-scale all cases, the lubricant used was mineral oil. Four different
contrast. surfaces were prepared on the workpieces: flat surfaces, dim-
Surface textures produced through all the aforementioned pled surfaces, shot blasted surfaces, and dimpled shot blasted
texturing methods have been used to create surfaces that surfaces. To create these surfaces, a custom-built computer
can exhibit significantly lower friction coefficients, or more numerical control (CNC) micro milling system was used.
specifically, friction coefficients that are greater than 10% In this CNC, an electric spindle (NSK Astro-E 800Z) was
lower than what is possible without surface texturing, in mounted onto a bracket that allowed the spindle to be rotated
both dry sliding [6] and fluid [13–19]-lubricated mechani- at an inclination angle relative to the sample surface and
cal systems. Significant efforts have been expended to gain secured during machining. In every case, the top layer of the
insight into the lubrication mechanisms of textured surfaces workpieces was first flattened using a flat end mill, leaving
using both experiments and simulations [20, 21]. Despite a 2 mm depression on the surface of the workpiece to facili-
these investigations, the exact mechanism by which friction tate lubricants. For this process, an uncoated tungsten carbide
is reduced through surface texturing is still uncertain and (WC) micro end mill (Richards 875-0125) having a diameter
dependent on a number of factors. For oil-lubricated con- of 0.125 mm, and four flutes, was used. This flattening process
tacts, several factors including depth of the surface texture, ensured that surfaces had the exact same surface normal and
the impression of the surface made in the texturing process, initial surface roughness in subsequent friction testing, as well
and the density of impressions, or dimples as they are often as the ability to store enough lubricant during friction testing.
called, can influence the friction coefficient. More recently, The flattening process also ensured that all surfaces had the
other factors including the sliding direction relative to the approximate same roughness before machining dimples or
shape of an asymmetric dimple and contact mechanism can roughening the surface with shot blasting.
affect the efficiency of textured surfaces at reducing friction. The forces exerted on the workpiece during friction testing
Additional parameters have also been investigated, including were measured by a piezo-electric table dynamometer (Kistler
the arrangement/staggering of dimples [21], surface rough- 9256C2), with resolution of 2 mN and frequency bandwidth of
ness outside of the dimpled region [40, 41], aspect ratio of 2 kHz, positioned between the workpiece and the translation
the dimples [13], dimple shapes [6], and the interplay of stages. The response of the table dynamometer was amplified
lubricant structure and surface texture [22–24]. Through all (Kistler 9025B) before recording the signals (National Instru-
these combinations, three mechanisms have been proposed ments 9205) at a sampling rate of 1 kHz in the x, y, and z direc-
to reduce friction through surface texturing: enhancement of tions. Following the surface flattening and milling of the oil
lubricant pressure under elastohydrodynamic sliding; entrap- reservoir depression, the surfaces were polished using a P800
ment of wear particles; and storage and subsequent faster grit sandpaper to remove machining marks and reduce sur-
delivery of the lubricant to the sliding interface. However, no face roughness. After treatment, the roughness on a flat work-
roadmap currently exits to identify under which conditions a piece was measured over a 5 × 5 mm2 area and was always
specific lubrication mechanism occurs, which could be used less than Ra = 0.178 ± 0.02 µm using a contact stylus (Mitu-
to make design predictions when developing a mechanical toyo SJ-201P). The entire system was mounted and secured
system enhanced with surface texturing. on a vibration isolation table. Experiments were conducted in
In this study, surfaces with multi-scale surface texturing laboratory air having a humidity range between 20 and 40%.
will be examined to determine if friction can be reduced The ruby ball was immersed in the oil and was not exposed
beyond monolithically textured surfaces, as well as to gain to the environment between measurements. A series of meas-
further insight into the lubrication mechanisms that occur in urements were acquired within 1 week, where the humidity
oil-lubricated textured contacts. The quality of the surfaces did not vary significantly (less than 5% RH) over the series of
produced through micro end milling and subsequent shot measurements. Thus, the results presented within one graph
blasting the surface will also be evaluated, as these two sur- are not impacted by variations in humidity.
face machining processes are very low cost and can produce
varying length-scale alterations of the surface. Finally, the 2.1 Machining of Dimples
impact of surface texturing on the wear exhibited by the sur-
face under varying loads and sliding speeds will be evaluated Circular-shaped dimples were machined onto the flattened
both to observe changes in the wear rates, as well as to gain surfaces using a single crystal diamond cutter with a diam-
insight into the friction or sliding mechanism occurring in eter of 250 µm. These diamond tools are widely used in
lubricated textured contacts.

13
Tribology Letters (2018) 66:132 Page 3 of 13  132

machining non-ferrous metals because they have a sharper surfaces. In this process, the nozzle-surface distance was
edge, a lower friction coefficient when in contact with alu- 25 mm with an air pressure of 0.45 MPa was used. After
minum, and a higher thermal conductivity than conventional blasting the process, the excess sand was removed with
end mills [25]. These tools also show high hardness, low deionized water. A schematic of sandblasting process is
wear rate, chemical inertness, and low adhesion to most shown in Fig. 2.
workpiece materials, which prevents built-up-edge [26].
Resulting from these properties, diamond cutters produce 2.3 Tribological Test Setup
surface textures with lower variation and more exact toler-
ances than cutters produced from other materials [17]. In To perform friction testing on the surfaces of the flat and tex-
addition, the diamond cutter has only one flute to prevent tured workpieces, the spindle of the mill was again rotated
non-uniformity of cutting depth caused by oscillations (i.e., such that it was parallel with the normal of the workpiece
tool runout) while the micro tool is rotating. Figure 1 shows surface. The end mill was removed, and one capped with a
a schematic of the arrangement of the electric spindle in 6-mm ruby hemisphere was put into the spindle. The spindle
relation to the workpiece during the machining of the micro was then translated downwards, allowing the ruby sphere
dimples. to make contact with the surface of the workpiece. A pho-
Fast translation of the workpiece beneath the tilted tograph of friction measurement apparatus/tribometer is
ball end mill ensures that a small piece of the workpiece shown in Fig. 3.
is removed during every revolution of the end mill. The The workpiece was then translated in a reciprocating
machining of circular dimple patterns was realized by trans- motion beneath the ruby sphere/spindle assembly for 100
lating the workpiece at a feed rate of 5 mm/secs and a con- cycles for a distance of 2 cm in each direction of the recip-
stant depth of 30 µm relative to the flattened surface while rocation cycle. The normal applied load was varied from
the spindle operated at a rotational speed of 1000 revolu- 1 to 25 N and two sliding speeds (0.5 mm/sec and 5 mm/
tions/minute and was placed at an inclination angle of 60° sec) were used. Hertzian contact pressure for these load
relative to the normal of the flattened surface. ranges was determined to be 450–1300 MPa, assuming a
Young’s modulus and Poisson ratio of 68.9 GPa and 0.33
2.2 Shot Blasting Particles for the Al6061-T6 [27] and 345 GPa and 0.28 for the ruby
hemisphere [28], respectively. In comparison to the applied
Shot blasting particle allowed for the development of stresses, the compressive strength for the Al6061-T6 alloy is
surfaces with a smaller scale of roughness than could be 310 MPa [27] and 2-2.9 GPa [28] for the ruby hemisphere,
achieved with the end mill dimple patterning. In this pro- indicating the applied loads are significantly above the yield
cess, a micro-abrasive blaster (COMCO ProCenter) was stress of the aluminum workpiece.
used to blast 10 µm aluminum oxide micro-abrasive particle In all tests, the workpieces were lubricated by mineral
powders (COMCO PD1029-4) at the flattened or dimpled oil (AMRESCO J217), having a viscosity of 14.2–17.0 cS
at 40 °C. A friction coefficient for each reciprocation cycle
was determined by subtracting the lateral forces measured
60°
in the forward sliding direction (+ x) from those measured
in the reverse sliding direction (− x) and then dividing the
Micro tool
tilted 60°

Aluminum 6061
Bedplate Workpiece

Fy
Fx
Table
Dynamometer Fz

Fig. 1  Schematic showing the inclination of the spindle relative to the Fig. 2  Schematic of the shot blasting process. Dimples of 30 µm were
workpiece allowing for the machining of dimple patterns on flattened machined with the end milling set up and then shot blasted with parti-
surfaces of the workpiece cles having an average diameter of 10 µm

13
132   Page 4 of 13 Tribology Letters (2018) 66:132

Ruby Hemisphere

Wear Tracks Slider v

Fluid Micro
Dimple
Table Dynamometer Aluminum
Workpiece

Stage Movement Fig. 4  A schematic of the tribological simulation. A slider moves


with a velocity v over a stationary aluminum workpiece with an
applied normal force F. The surface of the aluminum workpiece
Fig. 3  A photograph of friction measurement apparatus/tribometer. A
designed from the surface topography measured from a real surface,
6-mm ruby ball is attached at the end of an end mill held stationary
but contains at most one dimple in the simulation. A fluid between
while the workpiece, attached to 3-axis table dynamometer, is moved
the slider and the aluminum workpiece matching the physical proper-
in a reciprocating motion
ties of the oil used in the experiments is then placed in the sliding
contact
difference by two, and then dividing it by the average normal
force measured during the sliding period.
the one-dimensional Reynold’s equation for the hydrody-
In this calculation, only the middle 60% of the meas-
namic phenomena by average visco-correction on a multi-
ured forces are considered, removing the influence of turn-
block structured grid. The multigrid solution implemented in
ing effects that occur at the start/stop of each reciprocation
the VTL consists of a mesh of approximately 1 × 105 control
cycle. An average friction coefficient for each test was then
volumes [30]. The code also takes into account the asperity
calculated by taking an average of these friction coefficients
contact models based on the Greenwood and Tipp’s formula-
from cycle 20–80, to remove the influence of run-in. Error
tions for the description of the asperity contact between rough
bars reported in the average friction coefficient represent the
surfaces [31]. Cavitation of the lubricant was predicted adopt-
standard deviation in the calculated mean, as shown in Fig. 6.
ing the complementary Swift–Steiber boundary conditions.
As for the lubricant rheology, the viscosity-pressure depend-
2.4 3D Profilometry ence and viscosity-shear-thinning effects were considered as
well [30].
Characterization of the surface before and after friction test-
To replicate the experimental conditions in the VTL
ing was achieved by 3D optical profilometry (Zeta-20 optical
simulations, 2-D profiles of the machined surfaces acquired
profiler), allowing for the determination of the three-dimen-
from the Zeta-20 optical profiler were input into VTL for
sional shape of the textured surfaces. Following friction test-
the base counter surface. Figure 4 shows an schematic of the
ing, the volume of the workpiece that was removed during
simulation components.
sliding experiments was measured by optical profilometry
The top surface was modeled as a flat slider that moved at
through comparison of the surface before texturing and after
0.05 m/s in relation to the bottom dimpled/textured surface.
texturing.
The mechanical properties of the bottom textured surface
matched those of the AL6061-T6 surface, as stated previ-
2.5 Hydrodynamic Simulations ously, and the top flat surface was modeled to have the same
mechanical properties as the ruby hemisphere. Again, using
In order to understand and predict the tribological behavior
Herzian mechanics (Eqs. 1 and 2).
of lubricated textured surfaces, a series of simulations were
carried out using MAHLE Virtual Tribology Laboratory— (
1 − v21 1 − v22
)−1
VTL 3.8 developed by Tomanik et al. [29, 30]. The VTL ∗
E = + (1)
E1 E2
simulates lubrication conditions by simultaneously solving

13
Tribology Letters (2018) 66:132 Page 5 of 13  132

) 31 not adjusted to determine the impact of aspect ratio of the


3∗N∗R surface texture on the measured friction coefficient in this
(
a= , (2)
4 ∗ E∗ study. The shot blasting process was used to texture both the
where E* is the reduced Young’s modulus, En is the Young’s flat and dimpled surfaces, as mentioned previously. Calibra-
modulus of the n material, νn is the Poisson’s ratio of the n tion of the shot blasting process was done to ensure that
material. While the contact diameter is denoted by a, R is the machined dimples were not completely removed dur-
the radius of the ruby half-sphere, and N is the applied load. ing this stage. An example of a surface that had been both
The contact diameter between the ruby hemisphere and shot blasted and dimpled is shown in Fig. 5c. This optical
the aluminum workpiece was estimated to be 64 µm and image shows a significant change in the surface topography
190 µm, respectively. Given the relative lateral size of the compared to the dimpled surface. Figure 5d shows a line
dimple (150 µm), a flat slider was chosen to increase com- profile through the dimples, indicating that the height of
putational efficiency as the contact can be simplified as a the dimples remained roughly the same as before they were
flat slider at the larger applied loads. This analysis clearly hard shot blasted. The surface roughness was also measured
indicates that the contact is being subject to a high pressure, on the flat surfaces that had been hard shot blasted, and was
resulting in a high wear and deformation of the work piece observed to be 0.33 ± 0.02 µm, an increase of 88% from the
that can be observed directly after testing. flat surfaces.
To summarize the results obtained for the various applied
loads and speeds on the four different surfaces, Fig. 6 pre-
3 Results sents a Stribeck curve, including a flat surface, a shot blast
shot surface, a circular dimpled surface, and a blast shot
An optical image of the surface of the workpiece after dim- dimpled surface, the inside caption shows a zoom of the
ples were cut into it with the end mill is shown in Fig. 5a. highest friction coefficient values for all surfaces. Figure 6
The optical image shows that the dimples are circular and also shows a gradual reduction in friction as the surfaces
approximately 75 µm in radius. Figure 5b shows the line pro- become textured: the shot blasted surface performed better
file taken from red dashed line in the optical image (Fig. 5a) than the flat surface, but not as well as the dimpled sur-
showing that the dimples have been cut 30 µm below the face, and the best performance was detected for the surface
flattened surface. This cutting process resulted in the crea- that had been both dimpled and hard shot blasted. In Fig. 6,
tion of surface textures having an aspect ratio of 0.2, which it also can be observed that surface texturing significantly
is comparable to the aspect ratio that has been shown to impacts the onset of hydrodynamic lubrication, with the
significantly improve friction coefficients in studies spe- combined dimpled and hard shot blasted reducing the onset
cifically examining this property of the surface texture [13, of hydrodynamic lubrication more than the other three sur-
32–36]. In all experiments, this aspect ratio was used and faces. However, the dimpled surface shows a better perfor-
mance in terms of friction coefficient when the applied load

(a) (c)

50 μm 50 μm

(b) (d)

Fig. 5  a Plan view optical microscopy image of the dimpled surface. view optical microscopy image of the dimpled surface after hard shot
Dimples were observed to be circular. b Line profile of the height of blasting. d Line profile of the height of dimples taken through the red
the dimples taken through the red dashed line in (a) indicating that dashed line in (c) indicating that the dimple height remains relatively
the dimples have a similar height of approximately 30  µm. c Plan unchanged as a result of shot blasting. (Color figure online)

13
132   Page 6 of 13 Tribology Letters (2018) 66:132

explain the effect of film thickness on micro textured sur-


faces and its benefits; Křupka et al. [44] concluded that a
combination between film formation and textured surface
increased the tribological properties of machine parts.
The direct influence of the increased hydrodynamic pres-
sure on the measured Stribeck curves is compared in Fig. 9.

Fig. 6  A comparison between friction coefficients versus velocity/


load for a flat surface, a shot blasted surface, dimple surface, and a
shot blasted dimpled surface under lubricated sliding conditions. All
cases were performed for 100 iterations using both a reciprocation
speed of 0.5 mm/sec and 5 mm/sec. The inset in the top right shows
the section of the graph that is enclosed in the dashed gray square

is increased comparing with the flat surface in both studied


cases, 0.5 mm/sec and 5 mm/sec. The sandblasted dimple
surface shows a much better behavior when a high load was Fig. 7  Profiles of surfaces used during simulations. The top black line
applied comparing with both surfaces. In general, as the represents the profile of the flat surface, the top-middle red line repre-
speed increased the friction coefficient decreased indicat- sents the profile of a shot blasted surface, the bottom-middle blue line
represents the profile of a dimpled surface, and the bottom green line
ing that the lubrication regime transitioned from mixed to
represents the profile of a shot blasted dimpled surface. (Color figure
hydrodynamic lubrication at higher values of the lubrication online)
parameter.
Simulations were performed replicating the sliding con-
ditions and surface structure of the aluminum workpiece,
allowing for the generation of Stribeck curves. More spe-
cifically, these surfaces included a flat surface, a shot blast
surface, a dimpled surface with a single dimple of a depth
of 30 µm and diameter of about 150 µm, and a shot blasted
dimple having the same plan view dimensions but a depth
of only 10 µm. In all cases, a fixed flat surface was used for
the slider and a base oil lubricant with similar characteristics
as was used in experiments was chosen. Figure 7 shows the
profiles of every surface used in simulation process, starting
with a flat surface, then a shot blasted surface, after this a
dimpled surface and finally a shot blasted dimpled surface.
Hydrodynamic simulation results of these surfaces using the
VTL software are shown in Fig. 8.
From simulation results shown in Fig. 8, it is possible to
observe the cavitation effect previously observed in Refs.
[37–43], which results in an increase in the hydrodynamic
Fig. 8  Hydrodynamic pressure profile determined from VTL simula-
load capacity around the divergent region of dimples. Thus, tions of the profiles shown in Fig. 7. The pressure profiles for all sur-
the observed increase in the maximum hydrodynamic pres- faces are shown on one graph to make direct comparisons between
sure with changing the surface structure is generating a lift- the four variations, including a flat surface (black line), shot blasted
flat surface (red line), dimpled surface (blue line), and shot blasted
up force on the sliding counter surface. Moreover, the wider
dimpled surface (green line). The sliding speed was 0.05 m/s and an
width of pressure profiles for the shot blasted dimple case applied load of 10N was used in the simulations that generated these
is observed in Fig. 8. Several studies [44–47] have tried to pressure profiles. (Color figure online)

13
Tribology Letters (2018) 66:132 Page 7 of 13  132

The simulation results show enhanced lubrication through observed when the velocity was increased from 0.5 mm/s
improved hydrodynamic forces, and the experimental to 5 mm/s. The results show that roughening the surface has
results, whose enhanced lubrication properties are unknown, a greater impact at the higher sliding speeds than at slower
show that very good overlap in Fig. 9a, c, and d, correspond- sliding speeds, compared to change observed with velocity
ing to flat, dimpled, and shot blasted & dimpled surfaces, for the flat and dimpled surface.
but significant variance is observed between experiment and Optical profilometry was conducted to obtain three-
simulation results in Fig. 9b for the shot blasted surface. This dimensional images of the worn samples after tribological
indicates that either the mechanism responsible for reducing tests. Figure 10a, b, and c show a plan view optical image of
friction on these roughened surfaces is not captured solely a flat surface, dimpled surface, and a shot blasted dimpled
through hydrodynamic effects. This issue will be discussed surface after friction tests. To measure the wear volume on
in greater detail in the discussion section. these surfaces, a cross section from the three-dimensional
The film thickness was calculated for the four different image was taken, as shown in Fig. 11d, e, and f. These line
surfaces tested and at the two sliding speeds examined and profiles show the change in the surface as a result of the fric-
an applied load of 10 N, the results of which are summarized tion measurements. A summation of the area removed in the
in Table 1. Table 1 shows that the film thickness increased line profile along the wear scar to determine a volume lost
significantly for those surfaces that contain dimples, com- allowed for the determination of wear coefficients.
pared to the flat surface and the surface that had been shot Comparing the initial depth of the dimples, which was
blasted only. A small increase in the film thickness was approximately 30 µm as shown in Fig. 5b and d, with the

(a) (b)

(c) (d)

Fig. 9  a Stribeck curve showing acquired experimental data (black line) and simulated (blue dashed line) results of a dimpled surface. d
solid line) and simulated (black dashed line) results of a flat surface. Stribeck curve showing acquired experimental data (green solid line)
b Stribeck curve showing acquired experimental data (red solid line) and simulated (green dashed line) results of a shot blasted dimpled
and simulated (red dashed line) results of a shot blasted flat surface. surface. (Color figure online)
c Stribeck curve showing acquired experimental data (blue solid

13
132   Page 8 of 13 Tribology Letters (2018) 66:132

Table 1  Maximum values of film thickness separating the aluminum corresponding absolute wear volume versus load for each
workpiece and the slider at 0.5 mm/sec and 5 mm/sec and an applied of the surfaces is presented in Fig. 11, for both the dim-
load of 10 N
ple surface and the shot blasting dimpled surface, the vol-
Surface Max. Film thick- Max. Film ume due dimple cavity was not considered as part of wear
ness at 0.5 mm/sec thickness at measurement.
5 mm/sec
A significant decrease in the wear rate of shot blasted
Flat surface 0.27 µm 0.42 µm dimpled surfaces when compared with a flat surface for both
Shot blasted flat surface 0.69 µm 0.84 µm speeds 0.5 mm/sec and 5 mm/sec is shown in Fig. 11. Over-
Dimple surface 68.05 µm 76.03 µm all, Archard’s wear coefficient was higher on the flat surfaces
Shot blasted dimpled surface 68.17 µm 93.7 µm compared with any textured surface regardless the recipro-
cation speed. Comparing with a flat surface; shot blasted
Film thicknesses were determined as the distance from the flat sur-
face outside of the dimpled region to the top of the slider in the VTL surface have a wear reduction of 8.5% and 7% using a speed
simulations of 0.5 mm/sec and 5 mm/sec. Dimpled surface shows a wear
reduction of 23% and 20% with a speed of 0.5 mm/sec and
5 mm/sec., respectively. Finally, shot blasted dimpled sur-
depth measured in Fig. 11e and f, shows that the dimples face has a wear reduction of 30% and 30% when the recip-
have decreased in depth substantially. In fact, their depth rocation speed was 0.5 mm/sec and 5 mm/sec., respectively.
was only approximately 3–17 µm in the worn region. Given
that the wear track itself is only 1–3 µm deep in Fig. 11e and
f, we see that wear particles or other debris have filled the 4 Discussion
dimples significantly over the course of the measurement.
After tribological tests and similar to the procedure In general, most studies have shown that the hydrodynamic
presented by Timothy et al. [48], the area both below and pressure built in the liquid between the two surfaces is
above the zero level was integrated and multiplied by the enhanced in elastohydrodynamic lubrication (EHL), thus
track length to determine the total wear rate considering the typically moving the onset of EHL to higher loads, as well
Archard equation presented by Mandal et al. [49]. as lower velocities and lubricant viscosities [20, 46, 47].
However, other studies have shown that the textured surfaces
K∗l∗P
V= , (3) allow for wear particles to be trapped in the surface textures,
3∗H
reducing friction between the surfaces [34, 50, 51]. The third
where K is the wear coefficient, l is the sliding distance, mechanism that has been observed is that surface texturing
P is the applied load, and H is the material hardness. The acts as a lubricant reservoir, allowing it to reach the sliding

(a) (b) (c)

50 μm 50 μm 50 μm

(d) (e) (f)

Fig. 10  Optical images of the worn surface on a flat surface, b dim- shows the wear profile of d flat surface, e dimpled surface, and f a
pled surface, and c a shot blasting dimpled surface after a tribological shot blasting dimpled surface. The red dotted line indicates the posi-
test. These scratches were taken from samples where a reciprocation tion where the wear profiles were taken. (Color figure online)
speed of 5 mm/sec and applied load of 15 N was used. Also, Fig. 8

13
Tribology Letters (2018) 66:132 Page 9 of 13  132

(a) (b)

(c) (d)

Fig. 11  a Comparison of average wear versus applied load with a reciprocation speed of 5 mm/sec. for the flat surface (black squares),
reciprocation speed of 0.5 mm/sec for the flat surface (black squares), sandblasted flat surface (red triangles), dimple surface (blue inverted
sandblasted flat surface (red triangles), dimple surface (blue inverted triangles), and sandblasted dimple surface (green circles). Dashed
triangles), and sandblasted dimple surface (green circles). Dashed lines in the same colors as the data points represent linear fits to
lines in the same colors as the data points represent linear fits to the wear volume versus normal force data, where the intercept was
the wear volume versus normal force data, where the intercept was forced to 0  N. The slopes of these fits are 1200 ± 100  µm3/Nm (flat
forced to 0  N. The slopes of these fits are 8700 ± 100 µm3/Nm (flat surface), 1120 ± 100  µm3/Nm (shot blasted surface), 970 ± 70  µm3/
surface), 8000 ± 100 µm3/Nm (shot blasted surface), 6700  ± 700 Nm (dimpled surface), and 850 ± 70  µm3/Nm (shot blasted dimpled
µm3/Nm (dimpled surface), and 6000 ± 70 µm3/Nm (shot blasted surface). d Archard wear coefficients calculated from the slopes
dimpled surface). b Archard wear coefficients calculated from the determined in (c), the measured hardness of the aluminum work piece
slopes determined in (a), the measured hardness of the aluminum (141 ± 7 MPa), and the length of the wear track measured (300 µm).
work piece (141 ± 7 MPa), and the length of the wear track measured (Color figure online)
(300 µm). c Comparison of average wear versus applied load with a

interface faster [15, 51, 52]. We believe that we have directly has a hierarchical surface texture to it. While a shift of the
observed two of these three mechanisms in the textured sur- Stribeck curve to the left or right was not observed in Fig. 6,
faces examined in our reciprocating tribometer. a general trend of improved friction coefficients for a given
First, the experimental study of the influence of surface Stribeck coefficient or Hersey number was observed in the
texturing has shown that texturing the surface does indeed mixed and EHL regimes. Post-mortem analysis through the
reduce the friction coefficient compared with the flat sur- optical profilometry images in Fig. 8 shows that most of the
face, consistent with previous studies of textured surfaces surface texture and the roughness induced by shot blasting
[53–55]. Both randomly roughening the surface and pur- were removed in the contact region, so the influence that
posefully texturing the surface individually reduced friction, these two surface texturing strategies have on friction must
as shown in Fig. 6. The combined influence of the two shows occur beyond the contacting region between the workpiece
that improved performance can be made when the surface and the ruby hemisphere. While the observed results indicate

13
132   Page 10 of 13 Tribology Letters (2018) 66:132

a widening of the mixed and elastohydrodynamic lubrication by the surface texturing pattern in reducing friction [10, 18,
regime, the most often reported influence of surface textur- 58, 62].
ing has been to improve hydrodynamic lubrication [16, 42, Secondly, the post-mortem examination of the surfaces
54, 56, 57]. Despite this fact, surface texturing has also been shows that the wear scars are typically on the order of a
shown to improve boundary and mixed lubrication [21, 33, few micrometer deep into the surface. However, while the
55, 58–60]. Thus, the results here indicate the complexity dimples were 30 µm deep before friction testing, they were
of understanding lubrication enhanced by surface texturing, less than 10 µm deep after friction testing. The results sug-
which requires further examination to truly understand. gest that the dimples are being filled in with by-products of
Simulations conducted using the VTL software pack- sliding, suggesting that the capturing of wear particles and
age support the mechanism through enhanced lubrication. other contaminants that may influence the friction coefficient
While the simulation results do not capture the wear damage on the untextured surfaces is also important to the observed
caused through sliding that was observed in the experiments, reduction on the dimpled surfaces. The production and stor-
they do recover the basic structure of the Stribeck curve for age of wear particles or other by-products of sliding were
the various patterned surfaces, and show the same decrease not captured in the simulations. While the simulations did
in measured coefficients with the change in surface topogra- not replicate the experiments in this respect, the good over-
phy through the four samples. This observation suggests that lap between simulations and experiments resulting from the
the primary mechanism influencing the friction coefficient increased hydrodynamic lubrication that resulted from the
measured during lubricated sliding conditions is the buildup surface texturing suggests that the trapping of wear particles
of hydrodynamic pressure. Despite the good reproduction and other debris does not play a significant factor in reducing
of the experimental Stribeck curves at higher values of the the friction coefficient in textured surfaces.
Hersey number, the poor reproduction of the experimental Finally, the calculation of the wear coefficient provides us
results at lower values of the Hersey number is likely a result with some greater insight into the friction and wear perfor-
of the one-dimensional mathematical modeling implemented mance of the surfaces. The reduced friction with surface tex-
in VTL software package, which is insufficient to reproduce turing appears to be linked with reduced wear in every case.
the three-dimensional flows that occur in the experiments. Additionally, wear coefficients approximately one order of
More specifically, the simulations disregard any type of magnitude smaller when the velocity was increased from
influence of the two-dimensional effects on the flow fluid 0.5 mm/s to 5 mm/s, as shown in Fig. 11. Linking Fig. 11
through the micro-cavities present in the contact interface with Fig. 6 and the trends observed with the Hersey number,
[30]. it suggests that EHL lubrication and the associated low wear
To interpret the observed Stribeck curves in more detail, rates that were indeed achieved at the low applied loads and
the simulations show that there is a significant impact on the high sliding speeds, given the reduced friction forces and
film thickness with the change in surface structure. Similar wear coefficients. Thus, the observation of reduced friction
results have been observed previously, where higher sliding coefficients attributed to enhanced fluid pressure resulting
speeds and deeper surface features resulted in a decrease in from the surface texturing is supported by the observation
the measured friction coefficient on textured surfaces [61], of the lowered wear rates also observed with surface textur-
likely a result of improved fluid film thickness. Further- ing. Variation in the dimple spacing or size may result in
more, the observed improvement in the friction coefficient a variation in the relative contribution in the reduction of
observed between surface roughening through shot blasting friction of increased hydrodynamic pressure in the contact
compared with surface texturing, the latter which generates versus wear particle entrapment, but would require further
much deeper surface features, is consistent with the results experimentation to verify this hypothesis.
observed in Ref [61]. In samples where shot blasting was
conducted for longer times, resulting in a significant portion
of the dimples removed (only 5 µm remained), we observed 5 Conclusion
a significant increase in the friction coefficient compared
with those that remained at approximately 30 µm depth. This In this study, multi-scaled micro textured surfaces were cre-
fact also suggests that the dimples on the surface were con- ated by shot blasting and end milling of flat aluminum 6061-
tributing more strongly to the measured reduction in friction T6 surfaces. The performance of these surfaces in reducing
than the surface texture created by shot blasting. Finally, friction was evaluated with a reciprocating tribometer by
the observation of the friction coefficients converging to the pressing a 6-mm ruby hemisphere against the aluminum
same values at high values of the Hersey number for all surfaces and sliding them at 0.5 mm/s or 5 mm/s under
surfaces is likely a result of viscous fluid forces and other mineral oil-lubricated conditions. The performance of the
friction mechanisms overwhelming the contribution of the textured surfaces having a constant aspect ratio of 0.2 was
improved fluid pressure developed in the sliding interface evaluated through their sliding speed and load-dependent

13
Tribology Letters (2018) 66:132 Page 11 of 13  132

friction coefficients, as well as their sliding speed and load- 7. Graham, E., Park, C.I., Park, S.S.: Force modeling and appli-
dependent wear rates. The best performance for this surface cations of inclined ball end milling of micro-dimpled surfaces.
Int. J. Adv. Manuf. Technol. 70, 689–700 (2013). https​://doi.
texture shape, in terms of low friction coefficients and wear org/10.1007/s0017​0-013-5310-5
rates, was those surfaces that had been textured with dim- 8. Chae, J., Park, S.S., Freiheit, T.: Investigation of micro-cutting
ples and subsequently shot blasted, while flat, untextured operations. Int. J. Mach. Tools Manuf. 46, 313–332 (2006). https​
surfaces performed the worse. The combined influence of ://doi.org/10.1016/j.ijmac​htool​s.2005.05.015
9. Greiner, C., Del Campo, A., Arzt, E.: Adhesion of bioinspired
shot blasting and dimple surface textures had a synergistic micropatterned surfaces: effects of pillar radius, aspect ratio,
influence on the friction coefficient, suggesting that surface and preload. Langmuir. 23, 3495–3502 (2007). https​: //doi.
roughness of surface textures is an unrealized mechanism org/10.1021/la063​3987
to further improve the friction and wear performance of 10. Kovalchenko, A., Ajayi, O., Erdemir, A., Fenske, G., Etsion, I.:
The effect of laser surface texturing on transitions in lubrication
textured surfaces. Surface texturing resulted in a lowering regimes during unidirectional sliding contact. Tribol. Int. 38,
of the minimum friction coefficient that could be attained. 219–225 (2005). https​://doi.org/10.1016/j.tribo​int.2004.08.004
This minimum value occurred at approximately the same 11. Aspinwall, D., Wise, M., Stout, K.: Electrical discharge textur-
Hersey number for all surfaces. Surface textures were simu- ing. Int. J. 32, 183–193 (1992). http://www.scien​cedir​ect.com/
scien​ce/artic​le/pii/08906​95592​90077​T. Accessed Dec 5 2014
lated using the VTL software package, which closely repro- 12. Tseng, A.A.: Recent developments in micromilling using
duced the experimental friction results. This good agree- focused ion beam technology. J. Micromech. Microeng. 14,
ment between simulation and experiment suggests that the R15–R34 (2004). https​://doi.org/10.1088/0960-1317/14/4/R01
primary friction reducing mechanism was through enhanced 13. Greiner, C., Merz, T., Braun, D., Codrignani, A., Magag-
nato, F.: Optimum dimple diameter for friction reduction with
fluid pressure at the contact resulting from the presence of laser surface texturing: the effect of velocity gradient, Surf.
the dimples. The increased fluid pressure, and associated Topogr. Metrol. Prop. 3 (2015). https​://doi.org/10.1088/2051-
increasing fluid thickness in the contact, also provides a 672X/3/4/04400​1
mechanism for the reduced wear rates observed on textured 14. Meng, F., Zhou, R., Davis, T., Cao, J., Wang, Q.J., Hua, D., Liu,
J.: Study on effect of dimples on friction of parallel surfaces under
surfaces. Post-mortem analysis of the surfaces showed that different sliding conditions. Appl. Surf. Sci. 256, 2863–2875
significant wear debris was also captured by the dimples, (2010). https​://doi.org/10.1016/j.apsus​c.2009.11.041
indicating a second friction reducing mechanism that has a 15. Galda, L., Pawlus, P., Sep, J.: Dimples shape and distribution
less significant influence on the measured friction coefficient effect on characteristics of Stribeck curve. Tribol. Int. 42, 1505–
1512 (2009). https​://doi.org/10.1016/j.tribo​int.2009.06.001
compared with the increased fluid pressure. 16. Ramesh, A., Akram, W., Mishra, S.P., Cannon, A.H., Polycarpou,
A.A., King, W.P.: Friction characteristics of microtextured sur-
Acknowledgements  We would like to acknowledge to Consejo faces under mixed and hydrodynamic lubrication. Tribol. Int. 57,
Nacional de Ciencia y Tecnologia (CONACYT) México, and the Natu- 170–176 (2013). https​://doi.org/10.1016/j.tribo​int.2012.07.020
ral Sciences and Engineering Research Council (NSERC) of Canada 17. Li, K., Yao, Z., Hu, Y., Gu, W.: Friction and wear performance of
for providing funds to support the study, and Carrie Lin for her help to laser peen textured surface under starved lubrication. Tribol. Int.
develop the wear analysis code. 77, 97–105 (2014). https​://doi.org/10.1016/j.tribo​int.2014.04.017
18. Lu, X., Khonsari, M.M.: An experimental investigation of dimple
effect on the stribeck curve of journal bearings. Tribol. Lett. 27,
169–176 (2007). https​://doi.org/10.1007/s1124​9-007-9217-x
References 19. Yamakiri, H., Sasaki, S., Kurita, T., Kasashima, N.: Effects of
laser surface texturing on friction behavior of silicon nitride
1. Bruzzone, A.A.G., Costa, H.L., Lonardo, P.M., Lucca, D.A.: under lubrication with water. Tribol. Int. 44, 579–584. https​://
Advances in engineered surfaces for functional performance. doi.org/10.1016/j.tribo​int.2010.11.002 (2011)
CIRP Ann. - Manuf. Technol. 57, 750–769 (2008). https​://doi. 20. Gropper, D., Wang, L., Harvey, T.J.: Hydrodynamic lubrication of
org/10.1016/j.cirp.2008.09.003 textured surfaces: a review of modeling techniques and key find-
2. Erdemir, A.: Review of engineered tribological interfaces for ings. Tribol. Int. 94, 509–529 (2016). https​://doi.org/10.1016/j.
improved boundary lubrication. Tribol. Int. 38, 249–256 (2005). tribo​int.2015.10.009
https​://doi.org/10.1016/j.tribo​int.2004.08.008 21. Vladescu, S.C., Olver, A.V., Pegg, I.G., Reddyhoff, T.: The effects
3. Patankar, N.A.: Supernucleating surfaces for nucleate boiling and of surface texture in reciprocating contacts - An experimental
dropwise condensation heat transfer. Soft Matter. 6, 1613 (2010). study. Tribol. Int. 82, 28–42 (2015). https:​ //doi.org/10.1016/j.tribo​
https​://doi.org/10.1039/b9239​67g int.2014.09.015
4. Matsumura, T., Takahashi, S.: Machining of micro dimples in 22. Voevodin, A.A., Zabinski, J.S.: Laser surface texturing for adap-
milling for functional surfaces, In: Proceedings of the 14th Inter- tive solid lubrication. Wear. 261, 1285–1292 (2006). https​://doi.
national ESAFORM Conference on Material Forming, pp. 567– org/10.1016/j.wear.2006.03.013
572 (2011) 23. Wang, X., Kato, K., Adachi, K., Aizawa, K.: Loads carrying
5. Etsion, I.: State of the art in laser surface texturing. J. Tribol. 127, capacity map for the surface texture design of SiC thrust bear-
248 (2005). https​://doi.org/10.1115/1.18280​70 ing sliding in water. Tribol. Int. 36, 189–197 (2003). https​://doi.
6. Resendiz, J., Graham, E., Egberts, P., Park, S.S.: Directional fric- org/10.1016/S0301​-679X(02)00145​-7
tion surfaces through asymmetrically shaped dimpled surfaces 24. Qiu, Y., Khonsari, M.M.: Experimental investigation of tribo-
patterned using inclined flat end milling, Tribol. Int. 91 (2015) logical performance of laser textured stainless steel rings. Tri-
67–73 bol. Int. 44, 635–644 (2011). https​://doi.org/10.1016/j.tribo​
int.2011.01.003

13
132   Page 12 of 13 Tribology Letters (2018) 66:132

25. Jeong-Du Kim and Kang, Y.-H.: High-speed machining of alu- 43. Tang, W., Zhou, Y., Zhu, H., Yang, H.: The effect of surface tex-
minium using diamond end mills. Int. J. Mnch. Tools Manufact. turing on reducing the friction and wear of steel under lubricated
37, 1155–1165 (1997) sliding contact. Appl. Surf. Sci. 273, 199–204 (2013). https​://doi.
26. Heaney, P.J., Sumant, A.V., Torres, C.D., Carpick, R.W., Pfef- org/10.1016/j.apsus​c.2013.02.013
ferkorn, F.E.: Diamond coatings for micro end mills: enabling 44. Křupka, I., Poliščuk, R., Hartl, M.: Behavior of thin viscous
the dry machining of aluminum at the micro-scale. Diam. Relat. boundary films in lubricated contacts between micro-textured sur-
Mater. 17, 223–233 (2008). https​ : //doi.org/10.1016/j.diamo​ faces. Tribol. Int. 42, 535–541 (2009). https​://doi.org/10.1016/j.
nd.2007.12.009 tribo​int.2008.03.013
27. ASM Handbook Volume 2: properties and selection: nonferrous 45. Křupka, I., Hartl, M.: The influence of thin boundary films on real
alloys and special-purpose materials, ASM International, 1990 surface roughness in thin film, mixed EHD contact. Tribol. Int. 40,
28. Dobrovinskaya Elena, R., Leonid, A.: Lytvynov. Valerian Pish- 1553–1560 (2007). https​://doi.org/10.1016/j.tribo​int.2006.10.008
chik, Sapphire: Material, Manufacturing, Applications, Springer 46. Křupka, I., Hartl, M.: The effect of surface texturing on thin EHD
Science & Business Media, (2009). https​://doi.org/10.1007/978- lubrication films. Tribol. Int. 40, 1100–1110 (2007). https​://doi.
0-387-85695​-7 org/10.1016/j.tribo​int.2006.10.007
29. Profito, F.J., Zachariadis, D.C., Tomanik, E.: One dimensional 47. Mourier, L., Lubrecht, A.A., Donnet C. Mazuyer, D. Transient
mixed lubrication regime model for textured piston rings, (2011) increase of film thickness in micro-textured EHL contacts, Tri-
30. Prófito, F.J.: Modelagem Unidimensional Do Regime Misto De bol. Int. 39, 1745–1756 (2006). https​://doi.org/10.1016/j.tribo​
Lubrificação Aplicada a Superfícies Texturizadas, (2010) int.2006.02.037
31. Greenwood, J.A., Tripp, J.H., The contact of two nominally 48. Rupert, T.J., Schuh, C.A.: Sliding wear of nanocrystalline Ni-W:
flat rough surfaces, Proc. Inst. Mech. Eng. (1970). https​://doi. structural evolution and the apparent breakdown of Archard scal-
org/10.1243/PIME ing. Acta Mater. 58, 4137–4148 (2010). https:​ //doi.org/10.1016/j.
32. Greiner, C., Schäfer, M., Popp, U., Gumbsch, P.: Contact split- actam​at.2010.04.005
ting and the effect of dimple depth on static friction of textured 49. Mandal, D., Dutta, B.K., Panigrahi, S.C.: Wear and friction
surfaces. ACS Appl. Mater. Interfaces. 6, 7986–7990 (2014). https​ behavior of stir cast aluminium-base short steel fiber reinforced
://doi.org/10.1021/am500​879m composites. Wear. 257, 654–664 (2004). https:​ //doi.org/10.1016/j.
33. Hu, T., Hu, L., Ding, Q., The effect of laser surface texturing wear.2004.02.006
on the tribological behavior of Ti-6Al-4V, Proc. Inst. Mech. 50. Kim, D.E., Cha, K.H., Sung, I.H., Bryan, J.: Design of surface
Eng. Part J J. Eng. Tribol. 226:854–863 (2012) https​://doi. micro-structures for friction control in micro-systems applica-
org/10.1177/13506​50112​45080​1 tions. CIRP Ann. - Manuf. Technol. 51, 495–498 (2002). https​://
34. Ito, H., Kaneda, K., Yuhta, T., Nishimura, I., Yasuda, K., Mat- doi.org/10.1016/S0007​-8506(07)61569​-8
suno, T.: Reduction of polyethylene wear by concave dimples on 51. Borghi, A., Gualtieri, E., Marchetto, D., Moretti, L., Valeri, S.:
the frictional surface in artificial hip joints. J. Arthroplasty. 15, Tribological effects of surface texturing on nitriding steel for high-
332–338 (2000). https:​ //doi.org/10.1016/S0883-​ 5403(00)90670-​ 3 performance engine applications. Wear. 265, 1046–1051 (2008).
35. Nakano, M., Korenaga, A., Korenaga, A., Miyake, K., Murakami, https​://doi.org/10.1016/j.wear.2008.02.011
T., Ando, Y., Usami, H., Sasaki, S.: Applying micro-texture to 52. Xing, Y., Deng, J., Wu, Z., Cheng, H.: Effect of regular surface
cast iron surfaces to reduce the friction coefficient under lubri- textures generated by laser on tribological behavior of Si3N4/
cated conditions. Tribol. Lett. 28, 131–137 (2007). https​://doi. TiC ceramic. Appl. Surf. Sci. 265, 823–832 (2013). https​://doi.
org/10.1007/s1124​9-007-9257-2 org/10.1016/j.apsus​c.2012.11.127
36. Zhang, B., Huang, W., Wang, J., Wang, X.: Comparison of the 53. Ryk, G., Etsion, I.: Testing piston rings with partial laser surface
effects of surface texture on the surfaces of steel and UHMWPE. texturing for friction reduction. Wear. 261, 792–796 (2006). https​
Tribol. Int. 65, 138–145 (2013). https​://doi.org/10.1016/j.tribo​ ://doi.org/10.1016/j.wear.2006.01.031
int.2013.01.004 54. Kovalchenko, A., Ajayi, O., Erdemir, A., Fenske, G., Etsion,
37. Brizmer, V., Kligerman, Y., Etsion, I.: A laser surface textured I.: The effect of laser texturing of steel surfaces and speed-load
parallel thrust bearing. Tribol. Trans. 46, 397–403 (2003). https​ parameters on the transition of lubrication regime from boundary
://doi.org/10.1080/05698​19049​04260​07 to hydrodynamic. Tribol. Trans. 47, 299–307 (2004). https​://doi.
38. Cupillard, S., Glavatskih, S., Cervantes, M.J., Computational fluid org/10.1080/05698​19049​04409​02
dynamics analysis of a journal bearing with surface texturing, 55. Ryk, G., Kligerman, Y., Etsion, I.: Experimental investigation of
Proc. Inst. Mech. Eng. Part J J. Eng. Tribol. 222:97–107 (2008) laser surface texturing for reciprocating automotive components.
https​://doi.org/10.1243/13506​501JE​T319 Tribol. Trans. 45, 444–449 (2002). https​://doi.org/10.1080/10402​
39. Dobrica, M.B., Fillon, M., Pascovici, M.D., Cicone, T., Opti- 00020​89825​72
mizing surface texture for hydrodynamic lubricated contacts 56. Costa, H.L., Hutchings, I.M.: Hydrodynamic lubrication of tex-
using a mass-conserving numerical approach, Proc. Inst. Mech. tured steel surfaces under reciprocating sliding conditions. Tri-
Eng. Part J J. Eng. Tribol. 224:737–750 (2010) https​://doi. bol. Int. 40, 1227–1238 (2007). https​://doi.org/10.1016/j.tribo​
org/10.1243/13506​501JE​T673 int.2007.01.014
40. Meng, F.M., Yang, T., Preliminary study on mechanism of cavi- 57. Yu, H., Wang, X., Zhou, F.: Geometric shape effects of surface
tation in lubricant of textured sliding bearing, Proc. Inst. Mech. texture on the generation of hydrodynamic pressure between con-
Eng. Part J-Journal Eng. Tribol. 227: 695–708 (2013) https​://doi. formal contacting surfaces. Tribol. Lett. 37, 123–130 (2010). https​
org/10.1177/13506​50112​46856​0 ://doi.org/10.1007/s1124​9-009-9497-4
41. Shinkarenko, A., Kligerman, Y., Etsion, I.: The effect of surface 58. Braun, D., Greiner, C., Schneider, J., Gumbsch, P.: Efficiency
texturing in soft elasto-hydrodynamic lubrication. Tribol. Int. 42, of laser surface texturing in the reduction of friction under
284–292 (2009). https​://doi.org/10.1016/j.tribo​int.2008.06.008 mixed lubrication. Tribol. Int. 77, 142–147 (2014). https​://doi.
42. Zhang, J., Meng, Y.: Direct observation of cavitation phenom- org/10.1016/j.tribo​int.2014.04.012
enon and hydrodynamic lubrication analysis of textured surfaces. 59. Liew, K.W., Kok, C.K., Ervina, M.N., Efzan: Effect of EDM dim-
Tribol. Lett. 46, 147–158 (2012). https​://doi.org/10.1007/s1124​ ple geometry on friction reduction under boundary and mixed
9-012-9935-6 lubrication. Tribol. Int. 101, 1–9 (2016). https:​ //doi.org/10.1016/j.
tribo​int.2016.03.029

13
Tribology Letters (2018) 66:132 Page 13 of 13  132

60. Wakuda, M., Yamauchi, Y., Kanzaki, S., Yasuda, Y.: Effect of behavior of laser textured surfaces. Meccanica. 46, 567–575
surface texturing on friction reduction between ceramic and steel (2011). https​://doi.org/10.1007/s1101​2-010-9316-x
materials under lubricated sliding contact. Wear. 254, 356–363 62. Wang, X., Liu, W., Zhou, F., Zhu, D.: Preliminary investigation of
(2003). https​://doi.org/10.1016/S0043​-1648(03)00004​-8 the effect of dimple size on friction in line contacts. Tribol. Int. 42,
61. Vilhena, L.M., Podgornik, B., Vižintin, J., Možina, J.: Influence 1118–1123 (2009). https​://doi.org/10.1016/j.tribo​int.2009.03.012
of texturing parameters and contact conditions on tribological

13

You might also like