Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

Journal Pre-proof

Wear mechanisms of a sintered tribofilm in boundary lubrication regime

Steven J. Thrush, Allen S. Comfort, James S. Dusenbury, Xue Han, Gary C. Barber,
Xia Wang, Hongwei Qu

PII: S0043-1648(21)00321-5
DOI: https://doi.org/10.1016/j.wear.2021.203932
Reference: WEA 203932

To appear in: Wear

Received Date: 15 January 2021


Revised Date: 9 April 2021
Accepted Date: 24 April 2021

Please cite this article as: S.J. Thrush, A.S. Comfort, J.S. Dusenbury, X. Han, G.C. Barber, X. Wang, H.
Qu, Wear mechanisms of a sintered tribofilm in boundary lubrication regime, Wear (2021), doi: https://
doi.org/10.1016/j.wear.2021.203932.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2021 Published by Elsevier B.V.


Steven J Thrush: Conceptualization, Methodology, Formal analysis, Investigation, Writing –
Original Draft, Writing – Review and Editing Allen S Comfort: Conceptualization,
Methodology, Writing – Original Draft James S Dusenbury: Conceptualization, Writing –
Original Draft, Writing – Review and Editing, Supervision Xue Han: Methodology, Writing –
Original Draft Xia Wang: Formal analysis, Writing – Original Draft Hongwei Qu: Resources,
Writing – Original Draft Gary C Barber: Conceptualization, Resources, Writing – Original
Draft, Supervision

of
ro
-p
re
lP
na
ur
Jo
WEAR MECHANISMS OF A SINTERED TRIBOFILM IN BOUNDARY

LUBRICATION REGIME

Steven J Thrush1,2

Allen S Comfort1

James S Dusenbury1

Xue Han2

of
ro
Gary C Barber2

-p
Xia Wang2
re
Hongwei Qu3
lP

1
US Army DEVCOM Ground Vehicle Systems Center, 6501 East 11 Mile Rd, Warren,
Michigan, 48397;
na

2
Automotive Tribology Center, Oakland University, 2200 N Squirrel Rd, Rochester, Michigan,
ur

48309;
3
Advanced Microanalysis Lab, Oakland University, 2200 N Squirrel Rd, Rochester, Michigan,
Jo

48309;

*Email: Steven.J.Thrush.civ@mail.mil (Steven J Thrush)

1
ABSTRACT
A research effort was conducted to investigate wear mechanisms of a sintered tribofilm formed

by a zirconia nanoparticle antiwear additive. Spherical five nanometer diameter zirconium oxide

(ZrO2) nanoparticles were dispersed in polyalphaolefin (PAO) synthetic base oil and used to

generate approximately 120 nm thick tribofilms on AISI 52100 steel counterfaces in a ball-on-

disk tribometer under boundary lubrication conditions. The sintered tribofilms were subsequently

worn using the same conditions without the nanoparticle additive. Semi-in situ measurements of

of
tribofilm thickness were conducted with an optical interference-measuring device to observe

ro
tribofilm growth and subsequent tribofilm wear. A rapid initial wear process was observed,
-p
however, a thinner tribofilm provided enduring protection of the substrate steel for the 120 min
re
test duration. Without nanoparticles to replenish the tribofilm, the wear was deconvoluted from
lP

growth. Comparatively, lack of a tribofilm would result in scuffing within 10 min when testing

the unadditized PAO base oil. Scanning electron microscopy and profilometry suggest the
na

tribofilms were smoothened by the wear process. Evidence of crack formation between pores of
ur

the tribofilm illustrate a wear mechanism for the tribofilms. The cracks were concentrated where
Jo

the contact pressure was the largest. Mild abrasive wear appeared as minor furrows in the sliding

direction throughout the full width of the tribofilm.

KEYWORDS

Antiwear Additives, Boundary Lubrication, Wear, Lubricants.

1) INTRODUCTION

Nanoparticles dispersed in fluids have shown beneficial tribological performance [1-5]; however,

the mechanisms are not well understood. Some of the originally proposed mechanisms included

2
rolling, protecting, mending and polishing [6-7]. As the name implies, the rolling mechanism

refers to the nanoparticles acting as ball bearings, encouraging rolling contact as opposed to

sliding. The protecting mechanism introduced the potential of a nanoparticle film being formed.

Mending and polishing fill furrows and remove peaks, respectively, to create smoother surfaces.

The proposed mechanisms were later extended to include shearing lamellar nanostructures such

as platelets and fullerenes to reduce the interfacial shear inside the tribocontact [8-9]. These

mechanisms have since been supported by several sources [10-21], suggesting the dominate

of
nanoparticle mechanism is the response to a combination of operating conditions, material

ro
properties, and surface topography. More recently, tribosintering has gained attention as an
-p
additional mechanism for forming tribofilms and could be interpreted as a subtype of the above-
re
mentioned protective and mending mechanisms [1-3, 22-28]. Tribosintering can occur under the
lP

high pressure and temperature in a tribological contact to permit sintering mechanisms resulting
na

in tribofilm growth. Thus, similar to conventional sintering, tribosintering involves particle-to-

particle diffusion processes to create a larger formation, but, in this case, in the form of a
ur

tribofilm. Regardless, it is worth noting that all of the aforementioned mechanisms for
Jo

nanoparticle friction and wear reduction are based on mechanical means. This is directly contrary

to the manner in which conventional antiwear additives work such as ZDDP, which form

tribofilms through complex chemical and physical processes [29]. Although the details have not

been completely resolved, ZDDP requires a significant activation energy through either high

temperature or high shear to form chemical reaction films on the surface [29-32]. ZDDP

tribofilms been documented to have exceptional resistance to wear except in the presence of a

dispersant or soot [30-31]. However, the performance of ZDDP does not always translate well to

3
the non-ferrous materials and coatings that are becoming more common in the automotive and

other industries [33-34].

Although tribosintering may be a more recent concept, conventional sintering has been studied

for decades. Some industrial sintering processes include thermal, resistive, pressure assisted

thermal, plasma assisted and dynamic compaction [35-42]. In pressure-assisted thermal sintering

for example, the application of high pressure reduces the amount of thermal energy required for

of
sintering by encouraging particle-to-particle diffusion and densification. Similarly, in dynamic

ro
compaction, gigapascal scale impact pressures occurring on the microsecond time scale result in

-p
large strain rates and compressive stresses reducing the need for high temperature [43-45]. The

fundamental mechanisms of these processes are believed to be similar to that of the ZrO2
re
nanoparticles outlined in this study. It can be envisioned that nanoparticles entrained into the
lP

contact can be trapped between converging asperities. The resulting local stress concentrations at
na

these interfaces would exceed gigapascal pressures and each event would transpire over a very
ur

limited time. Such a progression would lend to numerous nanoscale sintering events culminating

in a sintered tribofilm. Thus, this pressure driven process would require temperatures far below
Jo

the 1400 °C needed for thermal sintering of ZrO2 [46].

In a prior investigation, functionalized ZrO2 nanoparticles dispersed in a PAO base oil were

studied to understand the role of nanoparticle concentration on tribofilm growth in boundary

lubrication [1]. Higher nanoparticle concentration was found to increase the tribofilm growth

rate, however, further improvement was not observed above a nanoparticle concentration of 0.1

%wt, which was identified as the ‘critical concentration’ for the operating conditions. Further

study revealed additional applied load improved tribofilm growth [2]. These findings supported

the tribosintering formation mechanism through fundamental sintering processes such as

4
increased particle-to-particle diffusion. While the tribosintering mechanism has proven to be

dependent on nanoparticle concentration, temperature, and load, the antiwear behavior of the

sintered tribofilms have not yet been documented. Thus, sintered tribofilms will be grown in

accordance with previous study, which will be followed by a novel approach to examine the

wear, wear mechanisms and overall durability of these films.

2) MATERIALS AND METHODS

of
Hypoid gear contact is predominately rolling motion with an element of sliding. Therefore,

ro
possible wear mechanisms can include those of both motion types. These contacts can be heavily

-p
loaded involving hardened steel, which can create gigapascal contact pressures in the boundary

lubrication regime [47]. Thus, the performance of boundary lubrication additives (i.e. extreme
re
pressure and antiwear additives) are paramount to ensure hardware longevity. Commercially
lP

available, ligand-capped ZrO2 nanoparticles (PixClear PC14-10-L01, Pixelligent Technologies


na

LLC) dispersed in a nominal 4 cSt at 100 °C PAO (i.e., PAO4) were selected because of their
ur

demonstrated ability to generate robust tribofilms [1-3]. However, it is unclear how robust the

generated tribofilms are when the free suspended nanoparticles are depleted. The ligand capping
Jo

agent successfully stabilized the ZrO2 nanoparticles in the PAO, reducing particle-to-particle

adhesion (e.g., van der Waals) and preventing agglomeration.

A PCS Instruments MTM2 ball-on-disk tribometer was selected due to its ability to control the

slide-to-roll ratio (SRR) of the test under high pressure boundary lubrication conditions

mimicking that of hypoid gear application. The apparatus was instrumented with 3D spacer layer

imaging allowing for optical interference images of the ball to be collected over time. The ball

would return to an indexed position allowing for analysis of the same location for all images,

ignoring variation around the track. During optical measurements, the ball is stopped and

5
pressed, with a 50 N load, against a glass lens that has a semi-reflective chromium layer. White

light is directed through a microscope to the contact and an RGB camera is used to collect data.

A portion of the light will reflect off the chromium layer, and a portion will travel through the

lens and the tribofilm and reflect off the substrate steel material. This difference in path length

will cause a phase shift in the imaging and is the mechanism used to create interference images.

The interference image is then correlated to tribofilm thickness using the refractive index of the

tribofilm. While zirconia can have a refractive index of 2 [48], this is the case of a pristine

of
zirconia film. The observed porosity of the zirconia tribofilm and integration of oil should

ro
reduce the effective refractive index. Without a true measurement of the zirconia tribofilm, the
-p
value used for the correlation was 1.4, that of a typical oil and ZDDP tribofilm. Mechanical
re
stylus measurements of the film thickness corroborated the optical measurements and the validity
lP

of the used refractive index. A larger refractive index would slightly reduce the tribofilm
na

thickness of the optical measurements.


ur

A wear study was conducted as follows. Initially, a 2 hour test is performed with PAO4 additized

with 0.1%wt ZrO2 nanoparticles under conditions selected to mimic hypoid gear application
Jo

listed in Table 1 to sinter a tribofilm, followed by 30 min of zero applied load. The 30 min pause

allowed for the PAO4 additized with ZrO2 nanoparticles to be drained, and three subsequent

rinses of the cup with 50 mL of neat PAO4 to remove any residual free particles. The test was

continued with 50 mL of neat PAO4 for 2 additional hours. The purpose was to disallow

additional tribofilm growth to occur and quantify the tribofilm wear rate for 2 hours. Isolation of

the tribofilm removal rate will help assist the deconvolution of competition between tribofilm

growth and removal processes. Previous work has concluded that the addition of 0.1%wt ZrO2

nanoparticles exhibited negligible effect on the viscosity of PAO4 over a temperature range of 23

6
to 100°C, thus the fluid film thickness and fluid entrainment should remain constant [1]. The

lambda ratio was calculated using Equation 1.

= (1)

Where is the lambda ratio, hmin is the minimum film thickness, Rq1 and Rq2 are the root mean square

roughness of the ball and disk. The minimum film thickness was calculated using the Hamrock and

Dowson equation. If the lambda ratio is less than one, the minimum film thickness is less than the

of
composite roughness of the surfaces. Thus, the calculated lambda value of 0.265 presented in Table 1

ro
verified the boundary lubrication regime in agreement with hypoid gear application.

-p
re
Table 1: Ball-on-Disk Tribological Test Conditions
lP

Temperature 100 °C
Ball load 50 N
Test duration 120 min growth +
120 min wear
na

Ball diameter 19 mm
Disk diameter 46 mm
ur

Hertzian contact 1.12 GPa


pressure (max)
Hertzian contact 0.75 GPa
Jo

pressure (mean)
Slide to roll ratio 50%
Entrainment speed 150 mm/s
Ball speed 112 mm/s
Disk speed 187 mm/s
λ 0.265
Optical interference 05, 10, 20, 30, 40, 50, 60, 70, 80, 90,
imaging times 100, 110, 120 min. Repeated for
wear.

The ball test specimen used was AISI 52100 steel with a surface finish of 10-20 nm Ra and a

reported hardness of 800-920 Hv. The counterface flat disk was also AISI 52100 steel with a

surface finish of 10-20 nm Ra and a reported hardness of 720-780 Hv. Both ball and disk test

7
specimens had an elastic modulus of 200-210 GPa. Prior to tribological testing, ball and discs

were cleaned by toluene rinse in glass beakers. The test specimens were then immersed in n-

hexane and placed into an ultrasonic water bath (70W, 40KHz output) for 10 min. Finally, the

specimens were rinsed with acetone and dried with a lint-free wipe. All tests were conducted in

laboratory ambient conditions between 20-50% relative humidity, and 20-23°C. The tribometer

cup and ball holder were cleaned via triplicate rinsing of 50/50 isooctane and isopropanol mix

followed by compressed air purge.

of
ro
After testing, the test specimens were placed in glass beakers, cleaned with an n-hexane rinse

-p
followed by submersion in n-hexane and placed in an ultrasonic bath for 5 min to remove any

residual oil. Verification of the presence of a tribofilm and chemical constituents were conducted
re
on a Hitachi S-3700 scanning electron microscope (SEM) paired with Oxford Instruments
lP

Energy Dispersive Spectroscopy (EDS). A working distance of 10.8 mm and accelerating


na

voltage of 20 kV were used for these measurements.


ur

A more detailed investigation of tribofilm morphology and porosity were then characterized on
Jo

the ball specimens using a JEOL JSM-6510 series SEM. Secondary electron images were

gathered at a working distance of 5 mm and an accelerating voltage of 10 kV. Examination of the

complete width of the tribofilm was gathered at low magnification to observe macroscale

tribofilm morphology, as well as, examinations of smaller sections at higher magnification to

investigate tribofilm porosity and microstructure. Examination of the tribofilm edge was

conducted to analyze the transition between test specimen and tribofilm, and to inspect for the

presence of any tribofilm delamination and debris. Images after the wear process were

juxtaposed with unworn tribofilms grown under the same tribological conditions for comparative

purposes.

8
A mechanical stylus was used to quantify tribofilm surface roughness on the disk specimens. The

stylus was a cantilever system with a 2 µm radius diamond tip, using a 3 mg applied force so no

damage to the tribofilm would occur. Specimens were prepared in the same fashion as prior to

SEM analysis. Disk specimens were analyzed by 4 sets of 3 measurements. Each set was

separated by 90 degrees along the track controlled by rotation of the sample stage and each

measurement of the set was spaced 250 µm apart. 1 mm scan lengths were taken perpendicular to

the track. The instrument software was subsequently used to identify the edges of the tribofilm,

of
calculate arithmetic mean (Ra) and measure width from edge-to-edge. Due to lack of tribofilm

ro
thickness data on the disk, a mechanical stylus was used to quantify average final tribofilm
-p
thickness as well. Standard deviation was represented with error bars for all mechanical stylus
re
data.
lP

3) RESULTS
na

First, characterization of a tribofilm grown under the conditions of Table 1 without the 120 min
ur

wear step was conducted to verify chemical constituents. The results from SEM imaging and
Jo

EDS spectra are shown in Figures 1A and 1B respectively. The SEM image clearly displays the

evidence of a tribofilm on the surface of the ball. EDS chemical maps confirmed a zirconium-

rich tribofilm. EDS spectra revealed a highly resolved L-α zirconium peak inside the tribofilm.

No indication of zirconium was found outside the tribofilm, negating the possibility of thermal

growth without the pressure of the tribocontact at 100 °C.

9
of
ro
-p
re
lP
na
ur
Jo

Figure 1. A) SEM micrographs and EDS maps of Fe and Zr demonstrating a sintered tribofilms
generated on the steel ball from the ball-on-disk tribometer. Tribofilm track designated with red
lines. B) EDS spectra outside and inside the tribofilm. Secondary electron beam at 80 times
magnification, 20 kV accelerating voltage, and 10.8 mm working distance.

Optical interference images were collected at times specified in Table 1 using the previously

described 3D spacer layer imaging approach. These images were correlated to tribofilm

10
thickness for the sintered tribofilm growth and subsequent wear shown in Figure 2. The tribofilm

growth was rapid during the first 20 min of test, reaching an equilibrium tribofilm thickness from

20 to 120 min. This equilibrium state can be related to a balance of growth and wear processes of

the tribosystem. The vertical dashed line at the 120 min step represents the 30 min pause in the

test where the nanofluid was changed to unadditized PAO4 (i.e. removal of the free

nanoparticles), effectively disrupting the balance between growth and wear processes for the

remainder of the test. This disruption is observed through the rapid decrease in tribofilm

of
thickness. The rapid decrease in tribofilm thickness can be attributed to the isolation of the

ro
tribofilm wear process. While trial 3 maintained a thick tribofilm until roughly 170 min, the
-p
slope of the rapid wear process thereafter was similar to trial 1 and 2. A maximum wear rate
re
approximation was calculated for the 20 minutes of most wear. These values ranged from 3.2 to
lP

3.8 with an average of roughly 3.5 nm/min. For all three trials, an enduring tribofilm remained at
na

roughly a thickness of 30-60 nm. The enduring tribofilm appeared to have a slight increasing

thickness for the remainder of the test.


ur
Jo

11
of
ro
-p
re
lP

Figure 2. Tribofilm thickness over time for ZrO2 nanofluids. The vertical dashed line represents
the oil change to neat PAO4.
na

Optical interference images used to correlate to tribofilm thickness values shown on Figure 2 are
ur

shown in Figure 3. The first image at time 00 represents a clean ball specimen, as the light blue
Jo

is a result of the semi-reflective chromium layer of the optical system. Already by the 05 min

image, tribofilms grew substantially ranging from orange color and dark blue, representing

moderate and high thickness regions, respectively. Generation of a steady state tribofilm

occurred within the first 20 min. Verification of the 30 min pause not affecting the tribofilm can

be shown by comparing the image of the 120 min growth to 00 min of wear. As the images

appear to be identical, the 30 min pause and oil change was of no consequence to the tribofilm.

As the test progressed, the wear process occurred from the edge of the tribofilm, which

progressed to the center as the yellow regions representing a relatively thin tribofilm expanded to

12
the entire track width. The equilibrium state of the thinner film observed from 60 to 120 min was

subjected to further scrutiny employing SEM to be discussed later.

of
ro
-p
Figure 3. Optical interference imaging over time of the tribofilm growth and wear procedure on
re
the ball surface. Lines drawn on the 05 minute growth image to highlight the location of the
tribofilm.
lP
na

No in-situ measurements could be made on the disk regarding tribofilm thickness over time,

however, final tribofilm thickness measurements were conducted with the mechanical stylus. An
ur

average tribofilm thickness value was calculated from 12 stylus measurements per trial, where
Jo

each measurements were an average step height from tribofilm edge-to-edge perpendicular to the

track. These values were compared to the reduction in tribofilm thickness of the ball via the

optical interference imaging and are presented in Figure 4. Standard deviation of the data is

represented as error bars. Unworn sintered tribofilms were generated for 120 min under the same

tribological conditions, but were separate tests for comparison purposes, as the worn tribofilm

specimens were not removed from the tribometer during the 30 min pause to enable mechanical

stylus measurements. However, the good test repeatability provided assurance that it is

representative of the disk tribofilm before undergoing the wear process. The worn thickness was

13
reduced from 149.2 to 104.0 nm, a 30% reduction. Comparatively, the ball tribofilm thickness

was reduced from 137.2 to 42.9 nm, a 69% reduction.

180
Tribofilm Thickness (nm)

160
140
120
100

of
Disk
80

ro
Ball
60
40 -p
20
re
0
lP

Unworn Worn
na

Figure 4. Average final tribofilm thickness on the disk and ball, unworn vs worn. Twelve
measurements per disk for three trials. Four measurements per ball for three trials. Average and
standard deviation reported.
ur
Jo

Figure 5 provides the surface roughness from the mechanical stylus measurements of the disk

tribofilm. As all measurements were taken in the transverse orientation, they will capture peaks

and furrows along the sliding direction. The wear process reduced the average roughness from

54.0 to 40.5 nm. Thus, the wear process effectively created a smoother surface by roughly 25%.

However, the attribution for the smoothening wear process would have to be discovered through

subsequent SEM analysis.

14
70

60

50

40
Ra (nm)

30

of
20

ro
10

0
-p
re
Unworn Worn

Figure 5. Average tribofilm surface roughness on the disk, unworn vs worn. Twelve
lP

measurements per disk for three trials. Average and standard deviation reported.
na

SEM images are assembled to characterize tribofilm wear at several magnifications and locations
ur

shown in Figures 6 and 7. The unworn tribofilm is imaged in A, C, and E, which is compared to
Jo

the worn tribofilm of B, D, and F at magnifications of 200, 1000, and 3000, respectively, in

Figure 6. All images for Figure 6 were gathered centered to the tribofilm width and the sliding

direction was horizontally maintained. At 200 magnification, mild abrasive furrows can be

observed for both unworn and worn films. Thus, even though the nomenclature “unworn” is

used, the evidence of mild abrasive wear suggests minor wear processes concurrently occurring,

but dominated by tribofilm growth. The worn micrograph of Figure 6B reveals a high

concentration of elongated pitting towards the center of the tribofilm, where the Hertzian

pressure is greater in magnitude. At 1000 magnification, micropores can be observed for unworn

15
and worn tribofilms. The worn surface appears to highlight these pores as the peaks of the

tribofilm are removed. Further magnification to 3000 enables pore size to be quantified. The

unworn tribofilm contains pores of 1 μm or less, whereas the worn tribofilm shows a wider range

of pores from sub- to a few microns in size. Evidence of pores being connected by crack

formation along the sliding direction and pores coalescing with neighboring pores are also noted

for the worn tribofilms. Very little evidence of abrasive wear can be found at the 1000 and 3000

magnifications, as the wear process seems to be governed by adhesive and pitting wear

of
processes.

ro
-p
Figure 7 focuses on the edge of the tribofilm for the unworn tribofilm in A and C, and worn

tribofilm in B and D, at 600 and 2000 magnifications, respectively. This elucidates the transition
re
from machining marks of the hardware at the bottom of the images to tribofilm. At 600
lP

magnification, the mild furrows along the sliding direction of the unworn tribofilm are
na

effectively polished away during the wear process. Two thousand times magnification allows for
ur

the identification of tribofilm islands in the unworn tribofilm. It is believed the transition can be

attributed to the low Hertzian pressure not permitting adequate tribosintering. The wear process
Jo

appears to remove these tribofilm islands in the transition zone.

16
of
ro
-p
re
lP
na
ur
Jo

Figure 6. SEM micrographs of worn and unworn sintered tribofilms generated on the ball from
the ball-on-disk tribometer. Unworn tribofilm at A) 200, C) 1000, and E) 3000 magnification.
Worn tribofilm at B) 200, D) 1000, and F) 3000 magnification. Secondary electron beam, 10 kV
accelerating voltage, and 4-5 mm working distance.

17
of
ro
-p
re
lP
na

Figure 7. SEM micrographs of worn and unworn sintered tribofilms generated on the ball from
ur

the ball-on-disk tribometer at the edge of the tribofilm. Unworn tribofilm at A) 600, and C) 2000
magnification. Worn tribofilm at B) 600, and D) 2000 magnification. Secondary electron beam,
Jo

10 kV accelerating voltage, and 4-5 mm working distance.

Unexpectedly, Figure 6B of the worn tribofilm also captured a deeper micron-sized scratch on

either side of the tribofilm. It is believed that the scratch is most likely an anomaly from the

polishing process of the steel balls, or damage that occurred before testing during shipping or

handling. The tribofilm covers the scratch, which endures after the wear test. The upper and

lower transition areas between scratch and tribofilm are further magnified and presented in

Figures 8A and 8B respectively. At 2000 magnification, it is evident in Figure 8A that the

micron wide scratch is covered where thick tribofilm is generated. In the transition zone between

18
thick tribofilm and machining marks of the hardware, the thinner tribofilm is not sufficient to fill

the scratch, but particulate is observed along the right edge of the scratch. With the sliding

direction from right to left, there is an element of tribofilm movement under the high pressures of

the test, perhaps during the wear process, which attempted to extend across the valley. This

particulate, without support underneath from substrate steel, could result in sintered

agglomerates delaminating and re-entering the fluid. In Figure 8B, similar to 8A, particulate

could be observed on the right edge of the valley in agreement with the sliding direction of the

of
test. The scratch slightly extended to a pore of the thicker tribofilm area by roughly a micron

ro
before being hidden by overlaying tribofilm.
-p
re
lP
na
ur
Jo

19
of
ro
-p
re
lP
na
ur
Jo

Figure 8. SEM micrographs of a micron wide scratch being covered by sintered tribofilm on the
ball at the A) upper and B) lower areas of interest. Secondary electron beam, 10 kV accelerating
voltage, 5 mm working distance and magnifications of 2000.

20
Figure 9 attempts to resolve a worn tribofilm generation of a microcrack at 10,000 magnification.

Pores appear to coalesce to almost form an 8 μm crack along the sliding direction. Over a dozen

micropores contribute to this crack formation. Area 1 shows clear evidence of material plastic

deformation and subsequent fracture resulting in a coalesced pore. However, some material still

bridges the microcrack with evidence of material necking shown in areas 2, 3 and 4. Area 5, on

the right side of the crack, highlights a triangular shaped structure on the lower edge, which

appears to have been torn from the upper side as there is a triangular shaped void of material of

of
similar geometry and size.

ro
-p
re
lP
na
ur
Jo

Figure 9. Microcrack formation SEM image of worn tribofilm on ball. Secondary electron beam,
10 kV accelerating voltage, 5 mm working distance and magnifications of 2000. Areas of interest
circled and labelled 1 through 5.

21
The friction coefficient data is presented in Figure 10. During rapid tribofilm growth, the friction

coefficient is repeatable with an initial increase during the first 20 min followed by gradual

decrease over time. After the free particle removal and testing in neat PAO4, a slight depression

in friction coefficient is noted before a steady-state friction. Trial 3, the trial that sustained a

thicker tribofilm for longer as shown in Figure 2, exhibits a lower friction for the duration of the

test and a larger depression in friction after the removal of free particles. This could allude to the

experimental variability of the tribofilm generation and wear, or a more robust tribofilm being

of
generated.

ro
-p
re
lP

0.16
na

0.12
Friction Coefficient

ur
Jo

0.08

0.04

0
0 60 120 180 240
Time (min)

Trial 1 Trial 2 Trial 3 Removal


Figure 10. Friction coefficient over time for tribofilm growth and wear.

22
4) DISCUSSION

The preceding results illustrate the wear processes of a sintered tribological film. As free

nanoparticles were removed at the midpoint of the tribological testing, tribofilm growth was

of
successfully negated, and the tribofilm wore at a maximum rate of roughly 3.5 nm/min on the

ro
ball surface until a thinner, enduring film remained at a thickness of 30-60 nm shown in Figure
-p
2. The slight increase in thickness of the enduring tribofilm was noted, which, while
re
inconclusive, could be attributed to tribofilm transfer from the disk to the ball through an
lP

adhesive process, or worn tribofilm returning to the contact and being sintered into the tribofilm.
na

This resilient tribofilm was corroborated posttest using a mechanical stylus and SEM

micrographs. SEM of the worn tribofilm in Figures 6 and 7 yielded what appeared to be a
ur

plateaued, sintered film with uniform microporosity. The presence of porosity was not alarming,
Jo

as they commonly occur in sintered materials when cross-sectioned [49]. The appearance of the

tribofilms being plateaued in microscopy was corroborated with the tribofilm roughness data of

Figure 5, as a reduction by 25% was quantified when undergoing the wear process. The reduced,

plateaued state of the film was strong enough to mitigate further wear and maintain tribological

equilibrium under the boundary lubrication conditions.

In Figure 2 during the wear portion of the test, trials 1 and 2 experienced aggressive wear

immediately from 120 to 170 min. Trial 3 experienced mild wear during this duration, which

alludes to the film being resilient to wear until the film was compromised. At 170 min, wear

23
transitioned to a more aggressive rate similar to the other two trials. For all three trials, the wear

period of aggressive tribofilm wear was followed by a slightly recovering tribofilm thickness.

This phenomena could be attributed to low-wear friction theory stating that the strength of the

surface layers or films should increase in material strength with depth [50]. Thus, as the tribofilm

wore, the material strength improved, which permitted the transition to a low or minimal wear

state.

of
The documented porosity in unworn and worn tribofilms is believed to be a product of adhesive

ro
material transfer between the two surfaces, causing porosity as a function of the material transfer

-p
size as opposed to the nanoparticle size. As both surfaces are AISI 52100 steel with sintered

tribofilms, they are expected to have similar material properties in accordance with multi-layered
re
theory [51]. Thus, tribofilm material is likely transferred between surfaces as well as re-entrained
lP

into the fluid as three body wear particles. Prior investigation found the severity of adhesive
na

material transfer and the friction force were a product of applied load, which would support this
ur

hypothesis [2]. This process could further elucidate the particle clusters observed in the used

nanofluid sedimentation TEM analysis of a prior investigation [1].


Jo

Under a rolling contact, there is a central zone of stick with slip in the direction of rotation at the

trailing edge. Outside the central zone, there is stick with slip on the trailing edge in the opposite

direction of rotation [52]. However, the boundary between stick and slip will differ for this

tribological test due to the 50 % SRR, as the sliding component will contribute to an interfacial

shear term. This can explain the abrasive and adhesive wear mechanisms.

Posttest analysis of the worn tribofilm found severe pitting and crack formation in the center

region of the tribocontact as shown in Figure 6B. The centered portion of the tribocontact is

where the applied pressure is highest and interfacial shear is relatively low, which would

24
promote an adhesive wear process as opposed to abrasion. Low cycle fatigue pitting and crack

formation would also be more likely to occur in this region as the pressure further exceeds the

elastic limits of the materials [52]. As the SRR was held constant for all testing at 50 %, the

rolling element of the contact could influence the location of the stick-slip boundary, shear

distribution, and wear mechanisms. The dominant wear mechanism may tend towards an

abrasive wear mode approaching pure sliding and adhesive, pitting and crack formation wear

modes approaching pure rolling.

of
ro
Advantageously, the porosity throughout the sintered film could be beneficial in preventing

-p
scuffing under the severity of the boundary lubrication conditions through oil retention in the

contact preventing squeeze out. Oil impregnated parts have been used for nearly 100 years
re
including porous sintered bronze commonly referred to as oilite. The worn film did not reveal
lP

any underlying machining marks of the substrate material, thus, the tribofilm provided full
na

protection of the substrate steel. The significance of the tribofilm was emphasized with the
ur

testing of neat PAO4 in identical test conditions without a tribofilm from a previous study, where

scuffing failure would occur depicted as sharp rises in friction coefficient occurred within 10 min
Jo

for all three trials [1]. This sharp rise in friction was accompanied by severe adhesive wear and

plastic deformation, which are also indicators of scuffing.

While the film thickness was reduced by roughly 94 nm on the ball and only 45 nm on the disk

per Figure 4, this may be misleading because the path lengths are significantly different.

Multiplying by the path length and tribofilm width yields average tribofilm volumetric loss of

0.0015 mm3 and 0.0017 mm3 for the ball and disk, respectively. Thus, the disk experienced

slightly more volumetric wear than the ball even though the ball had a larger tribofilm thickness

reduction. Hardness as a function of sintered tribofilm depth was characterized by Khare et al.,

25
where thinner films correlated with higher hardness converging to the hardness of the substrate

material [3]. In context to the present study, the slightly thinner tribofilm on the ball could be

related to higher hardness and higher wear resistance. Comparatively for the disk, less

revolutions was countered by the softer AISI 52100 disk and thicker tribofilm to result in more

volumetric wear. Considering Archard’s wear equation for an adhesive contact, the relative

sliding distance is equally experienced by both surfaces even though the ball and disk speeds

differ. The applied load is constant for this investigation and can also be ignored. The reduced

of
form of Archard’s wear equation would directly correlate the volumetric loss to a

ro
nondimensional wear coefficient and inversely to the material hardness. The nondimensional
-p
wear coefficient would be simultaneously experienced by both surfaces, but would be expected
re
to change based on tribofilm thickness. Hardness of the layered mediums (i.e. softer tribofilm on
lP

harder substrate) are a function of tribofilm thickness as well, making it difficult to deconvolute.
na

However, in a general sense, the 720-780 HV substrate hardness disk with a thicker tribofilm

would be expected to have more volumetric wear than the 800-920 HV substrate hardness ball
ur

with a thinner tribofilm, which is in good agreement with the data.


Jo

The proposed tribofilm failure mechanism was depicted in Figure 9. The microcrack was

characterized as a product of multiple micropores in the tribofilm being joined together along the

sliding direction through adhesive wear. The microcrack appeared to widen with evidence of

material necking as well as similar geometry features and respective voids appearing opposite

each other along the crack. Cracks of this nature were documented to be centered in the tribofilm

thickness where the Hertzian pressure was greatest, best displayed in Figure 6B. While the

tribofilm endured the 2 hour wear process, the test could be extended to quantify the endurance

26
life of the thinner, wear-resistant film. However, the observable pore coalescence indicates the

film has a finite lifetime, as the cracks would propagate over time until failure.

SEM efforts also found the presence of a micron-scale scratch on the surface of a ball that

preceded the tribological test. Further examination of the boundary between the micron wide

scratch and tribofilm revealed that the scratch extended from one side of the tribofilm to the

other, but was covered by the tribofilm in the contact. In the transition areas from no tribofilm to

of
thick tribofilm, evidence of tribofilm on the leading edge of the scratch suggested tribofilm

ro
attempting to flow across the chasm. The tribofilm growth was not sufficient to mend this area,

-p
which could potentially result in local delamination of the tribofilm and re-entrainment into the

fluid.
re
lP

The friction coefficient during tribofilm growth of Figure 10 showed repeatable results for all

trials and with prior investigations [1-2]. A highly transient period was observed for the first 16
na

min, which was followed by a downward sloped trend to a friction coefficient of roughly 0.105.
ur

Subsequent wear testing for 120 min in neat PAO4 observed a depression in friction coefficient.
Jo

Trial 3, which maintained a film greater than 120 nm until 170 min in Figure 2, harmoniously

had the lowest friction throughout the full test duration during the wear process. Based on the

friction coefficient and tribofilm thickness over time data, trial 3 was believed to have created a

tribofilm with less imperfections, taking longer duration to transition from mild to severe wear.

As tribosintering was improved with increasing temperature and load [2-3], the test variability

could be potentially reduced by increasing test temperature or load to assist in particle-to-particle

diffusion mechanisms creating more consistent films.

Fujita and Spikes found that primary and secondary ZDDP tribofilms were resilient to wear

when undergoing boundary lubrication conditions at a lambda ratio of 0.5 in the neat base oil

27
[30]. While the sintered tribofilm of the present study experienced aggressive wear, the more

severe lambda ratio of 0.265 could be a contributing factor. Regardless, the existence of an

enduring sintered tribofilm is promising for future study.

The interface between the tribofilm and the AISI 52100 steel is not well understood. Khare et al.

theorized the steel surface would undergo a wear process to remove an oxidation layer and any

adsorbed contaminants because they noted a significant duration before tribofilm formation

of
could occur [3]. Considering tribosintering is a mechanical process, it is likely that there would

ro
be a limited transition thickness between the steel surface and tribofilm as there might be some

-p
diffusion between the two materials. An extensive posttest characterization effort is proposed as

future work to understand this interface further.


re
lP

5) CONCLUSIONS

A wear study was conducted on sintered tribofilms in boundary lubrication. The following
na

conclusions were made:


ur

1. Approximately 120 nm tribologically sintered films were generated on both AISI 52100
Jo

ball and disk test specimens.

2. The sintered tribofilm was smoothened by 25% through an adhesive wear process

combined with minor abrasive wear.

3. The film experienced a maximum wear process of roughly 3.5 nm/min on the ball

reduced the film thickness from 120 to approximately 40 nm. The plateau at 40 nm was

in agreement with a higher composite hardness based on layered media theory. The

higher material hardness would correlate to increased wear resistance in accordance with

Archard’s wear equation. This resulted in the transition to a low or minimal wear state.

28
4. The microscale porosity was fairly homogenous, but evidence of pores being connected

along the sliding direction creating microcracks concentrated towards the center of the

tribofilm width was found. If prolonged, the propagation of microcracks could result in

failure of the film.

5. Observation of a micron-wide scratch preceding the tribological study on either side of

the worn tribofilm revealed a thin tribofilm which attempted to traverse the gap in the

transition areas. The scratch was mended by tribofilm in the area of Hertzian pressure.

of
6. Friction coefficient and optical interference images over time elucidated that the tribofilm

ro
reached a new equilibrium state after the major wear process. When the same test
-p
conditions were repeated without the presence of an existing tribofilm in neat PAO4,
re
scuffing occurred within the first 10 min of test.
lP

6) ACKNOWLEDGEMENTS
na

N/A
ur

7) REFERENCES
Jo

(1) Thrush, S.J., Comfort, A.S., Dusenbury, J.S., Xiong, Y., Qu, H., Han, X., Schall, J. D.,
Barber, G.C. and Wang, X. (2020), “Stability, Thermal Conductivity, Viscosity, and Tribological
Characterization of Zirconia Nanofluids as a Function of Nanoparticle Concentration,” Tribology
Transactions, 63:1, pp. 68-76
(2) Thrush, S.J., Comfort, A.S., Dusenbury, J.S., Xiong, Y., Qu, H., Han, X., Schall, J. D.,
Barber, G.C. and Wang, X. (2021), “Study of Pressure Dependence on Sinterable Zircona
Nanoparticle Tribofilm Growth,” Tribology International, 154, 106683.
(3) Khare, H. S., Lahouij, I., Jackson, A., Feng, G., Chen, Z., Cooper, G. D. and Carpick, R. W.
(2018), “Nanoscale Generation of Robust Solid Films from Liquid-Dispersed Nanoparticles via
in Situ Atomic Force Microscopy: Growth Kinetics and Nanomechanical Properties,” ACS
Applied Materials & Interfaces 10 (46), pp. 40335-40347.
(4) Ganapathy Pandian, S. (2014) "Tribological Characteristics of Yttria Stabilized Zirconia
Nanolubricants," SAE Technical Paper 2014-01-2790, doi:10.4271/2014-01-2790.

29
(5) Gullac, B. and Akalin, O. (2010), “Frictional Characteristics of IF-WS2 Nanoparticles in
Simulated Engine Conditions,” Tribology Transactions, 53:6, pp. 939-947
(6) Kato, H. and Komai, K. (2006), “Tribofilm formation and mild wear by tribo-sintering of
nanometer-sized oxide particles on rubbing steel surfaces,” Wear, 262, pp. 36-41
(7) Lee, K., Hwang, Y., Cheong, S., Choi, Y., Kwon L., Lee, J. and Kim, S.H. (2009),
“Understanding the role of nanoparticles in nano-oil lubrication,” Tribology Letters, 35, 2, pp.
127–131.
(8) Tang, G., Sun, J., Chen, W., Tang, H., Wang, Y. and Li, C. (2013) “Surfactant-assisted
hydrothermal synthesis and tribological properties of flower-like MoS2 nanostructures,” Micro &
Nano Letters, 8:3, pp. 164-168.
(9) Celik, O.N., Ay, N. and Goncu, Y. (2013), “Effect of Nano Hexagonal Boron Nitride

of
Lubricant Additives on the Friction and Wear Properties of AISI 4140 Steel,” Particulate Science
and Technology, 31, pp. 501-506.

ro
(10) Sgroi, M., Gili, F., Mangherini, D., Lahouij, I., Dassenoy, F., Garcia, I., Odriozola, I. and
-p
Kraft, G. (2015), “Friction Reduction Benefits in Valve-Train System Using IF-MoS2 Added
Engine Oil,” Tribology Transactions, 58, pp. 207-214.
re
(11) Kaplan-Ashiri, I. and Tenne, R. (2016), “On the Mechanical Properties of WS2 and MoS2
Nanotubes and Fullerene-Like Nanoparticles: In Situ Electron Microscopy Measurements,” The
lP

Minerals, Metals & Materials Society, 68:1, pp. 151-167.


(12) Dou, X. (2018), “Self-Dispersed Crumpled Graphene Nanoparticles for Lubrication,” Ph.D.
na

Dissertation, Northwestern University, Evanston, Illinois.


(13) Jiao, D., Zheng, S., Wang, Y., Guan, R. and Cao, B. (2011), “The tribology properties of
ur

alumina/silica composite nanoparticles as lubricant additives,” Applied Surface Science, 257, pp.
5720-5725.
Jo

(14) Kheireddin, B.A., Lu, W., Chen, I. and Akbulut, M. (2013), “Inorganic nanoparticle-based
ionic liquid lubricants,” Wear, 303, pp. 185-190.
(15) Gara, L. and Zou, Q. (2013), “Friction and Wear Characteristics of Oil-Based ZnO
Nanofluids,” Tribology Transactions, 56, pp. 236-244.
(16) Zhang, Y., Xu, Y., Yang, Y., Zhang, S., Zhang, P. and Zhang, Z. (2015), “Synthesis and
tribological properties of oil-soluble copper nanoparticles as environmentally friendly lubricating
oil additives,” Industrial Lubrication and Tribology, 67:3, pp. 227-232.
(17) Ran, X., Yu, X. and Zou, Q. (2017), “Effect of Particle Concentration on Tribological
Properties of ZnO Nanofluids,” Tribology Transactions, 60:1, pp. 154-158.
(18) Borda, F., Oliveira, S., Lazaro, L. and Leiroz, A. (2018), “Experimental investigation of the
tribological behavior of lubricants with additive containing copper nanoparticles,” Tribology
International, 117, pp. 52-58.

30
(19) Harta, I., Owens, K., Santiago, S. D., Schall, J. D., Thrush, S., Barber, G. and Zou, Q.
(2013), “Tribological Performance of ZnO-Oil Nanofluids at Elevated Temperatures,” SAE
International Journal of Fuels and Lubricants”, 6, 1, pp. 126-131
(20) Han, X., Barber, G.C., Zhang, Z., Thrush, S., Schall, J.D. and Li, Z. (2019), “Tribological
performance of oil-based ZnO and diamond nanofluids,” Lubrication Science, 31:3, pp. 73-84
(21) Thrush, S. (2012), “Investigation of Dispersion, Stability, and Tribological Performance of
Oil-Based Aluminum Oxide Nanofluids,” Master’s Thesis, Oakland University, Department of
Mechanical Engineering, Michigan, USA.
(22) Pena-Paras, L., Taha-Tijerina, J., Garcia, A., Maldonado, D., Gonzalez, J.A., Molina, D.,
Palacios, E. and Cantu, P. (2014), “Antiwear and Extreme Pressure Properties of Nanofluids for
Industrial Applications,” Tribology Transactions, 57:6, pp. 1072-1076.

of
(23) Thottackkad, M.V., Rajendrakumar, P.K. and Prabhakaran, N.K. (2014), “Tribological

ro
analysis of surfactant modified nanolubricants containing CeO2 nanoparticles,” Tribology:
Materials, Surfaces & Interfaces, 8:3, pp. 125-130.
-p
(24) Alves, S.M., Mello, V.S., Faria, E.A. and Camargo, A.P.P. (2016), “Nanolubricants
developed from tiny CuO nanoparticles,” Tribology International, 100, pp. 263-271.
re
(25) Olofsson, J., Bexell, U. and Jacobson, S. (2012), “Tribofilm formation of lightly loaded self
mated alumina contacts,” Wear, 289, pp. 39-45.
lP

(26) Liu, S.Y., Wang, Y., Zhou, C. and Pan, Z.Y. (2015), “Mechanical properties and
tribological behavior of alumina/zirconia composites modified with SiC and plasma treatment,”
na

Wear, 332-333, pp. 885-890.


(27) Pena-Paras, L., Taha-Tijerina, J., Garcia, A., Maldonado-Cortes, D., Michalczewski, R. and
ur

Lapray, C. (2015), “Effect of CuO and Al2O3 nanoparticle additives on the tribological behavior
of fully formulated oils,” Wear, 332-333, pp. 1256-1261.
Jo

(28) Battez, A.H., Gonzalez, R., Viesca, J.L., Fernandez, J.E., Fernandez, J.M.D., Machado, A.,
Chou, R. and Riba, J. (2008), “CuO, ZrO2 and ZnO nanoparticles as antiwear additive in oil
lubricants,” Wear, 265, pp. 422-428.
(29) Spikes H. (2004), “The history and mechanisms of ZDDP,” Tribology Letters, 17:3.
(30) Fujita, H., Glovnea, R.P. and Spikes, H. (2005), “Study of Zinc Dialkyldithiophosphate
Antiwear Film Formation and Removal Processes, Part I: Experimental,” Tribology
Transactions, 48, pp. 558-566.
(31) Fujita, H. and Spikes, H. (2005), “Study of Zinc Dialkyldithiophosphate Antiwear Film
Formation and Removal Processes, Part II: Kinetic Model,” Tribology Transactions, 48, pp. 567-
575.
(32) Zhang, J. and Spikes, H. (2016), “On the Mechanism of ZDDP Antiwear Film Formation,”
Tribology Letters, 63:24.
(33) Topolovec-Miklozic, K., Lockwood, F. and Spikes, H. (2008), “Behavior of boundary
lubricating additives on DLC coatings,” Wear, 265, pp. 1893-1901.

31
(34) Vengudusamy, B., Green, J., Lamb, G. and Spikes, H. (2011), “Tribological properties of
tribofilms formed from ZDDP in DLC/DLC and DLC/steel contacts,” Tribology International,
44, pp. 165-174.
(35) German, R.M. (2014), “Sintering: From Empirical Observations to Scientific Principles,”
Elsevier Science.
(36) Kendall, K. (2012), “Problems of particle aggregation in ceramics,” Journal of the European
Ceramic Society, 32, pp. 2589-2596.
(37) Ghyngazov, S.A. and Frangulyan, T.S. (2017), “Impact of pressure in static and dynamic
pressing of zirconia ultradisperse powders on compact density and compaction efficiency during
sintering,” Ceramics International, 43, pp. 16555-16559.
(38) Balakrishnan, A., Pizette, P., Martin, C.L., Joshi, S.V. and Saha B.P., (2010), “Effect of

of
particle size in aggregated and agglomerated ceramic powders,” Acta Materialia, 58, pp. 802-
812.

ro
(39) Sadeghi, B., Shamanian, M., Ashrafizadeh, F., Cavaliere, P., Sanayei, M. and Szpunar, J.A.
-p
(2018), “Microstructural behaviour of spark plasma sintered composites containing bimodal
micro- and nano-sized Al2O3 particles,” Powder Metallurgy, 61:1, pp. 50-63.
re
(40) Miyake, K., Hirata, Y., Shimonosono, T. and Sameshima, S. (2018), “The Effect of Particle
Shape on Sintering Behavior and Compressive Strength of Porous Alumina,” Materials, 11,
lP

1137.
(41) Kocjan, A., Logar, M. and Shen, Z. (2017), “The agglomeration, coalescence and sliding of
na

nanoparticles, leading to the rapid sintering of zirconia nanoceramics,” Scientific Reports,


7:2541.
ur

(42) Castro, R.H.R. and Benthem, K. (2013), “Sintering: Mechanisms of Convention


Nanodensification and Field Assisted Processes,” Engineering Materials, vol. 35.
Jo

(43) He, R., Niu, F. and Chang, Q. (2018), “Tribological properties of PI-SiC nanocomposite
prepared by hot dynamic compaction,” Journal of Thermoplastic Composite Materials, 31:8, pp.
1066-1077.
(44) Rahmani, K., Majzoobi, G.H. and Atrian, A. (2018), “A novel approach for dynamic
compaction of Mg–SiC nanocomposite powder using a modified Split Hopkinson Pressure Bar,”
Powder Metallurgy, 61:2, pp. 164-177.
(45) Alba-Baena N.G. (2007), “Mechanical behavior of shock-wave consolidated nano and
micron-sized Al/SIC and Al/Al2O3 two-phase systems characterized by light and electron
metallography,” Ph.D. Dissertation, University of Texas at El Paso.
(46) Wu, Y., Bandyopadhyay, A. and Bose, S. (2004), “Processing of alumina and zirconia nano-
powders and compacts,” Materials Science and Engineering A, 380, pp. 349-355.
(47) Stump, B.C., Zhou, Y., Viola, M.B., Xu, H., Parten, R.J. and Qu, J. (2018), “A rolling-
sliding bench test for investigating rear axle lubrication,” Tribology International, 121, pp. 450-
459.

32
(48) Elinski, M.B., LaMascus, P., Zheng, L., Jackson, A., Wiacek, R.J. and Carpick, R.W.
(2020), “Cooperativity Between Zirconium Dioxide Nanoparticles and Extreme Pressure
Additives in Forming Protective Tribofilms: Toward Enabling Low Viscosity Lubricants,”
Tribology Letters, 68:107.
(49) Yu, C., Cao, P. and Jones, M.I. (2017), “Titanium Powder Sintering in a Graphite Furnace
and Mechanical Properties of Sintered Parts,” Metals, 7, 67.
(50) Kragelski, I.V. (1965), “Friction and Wear,” Butter Worth, London.
(51) Bhattacharya, A.K. and Nix, W.D. (1988), “Analysis of elastic and plastic deformation
associated with indentation testing of thin films on substrates,” International Journal of Solids
and Structures, 24(12), pp.1287-1298.

of
(52) Halme, J. and Andersson, P. (2010), “Rolling contact fatigue and wear fundamentals for
roller bearing diagnostics - state of the art,” Proceedings of the Institution of Mechanical

ro
Engineers, Part J: Journal of Engineering Tribology, 224(4), pp. 377–393.

-p
re
lP
na
ur
Jo

33
• A wear study was conducted on sintered tribofilms in boundary lubrication.

• A thinner enduring film remained after a wear process was observed.

• Microcracks concentrated towards the center of the worn tribofilm were found, which

could be a possible failure mechanism.

• Without the presence of a tribofilm in neat PAO4, scuffing occurred within the first 10

min of test.

of
ro
-p
re
lP
na
ur
Jo

You might also like