Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Sudesh Singh

Department of Mechanical Engineering,


Indian Institute of Technology (BHU),
Varanasi, Uttar Pradesh 221005, India
e-mail: sudesh.sengar@gmail.com

Xinchun Chen Tribological Performance of Steel


Department of Mechanical Engineering,
State Key Laboratory of Tribology,
Tsinghua University,
With Multi-Layer Graphene
Beijing 100084, China
e-mail: chenxc1213@mail.tsinghua.edu.cn Grown by Low-Pressure Chemical
Chenhui Zhang
Department of Mechanical Engineering,
Vapor Deposition
State Key Laboratory of Tribology,
Tsinghua University, To explore the potential of directly grown multi-layer graphene as an agent in reducing fric-
Beijing 100084, China tion and wear of steel on steel tribo-pair, multi-layer graphene films were synthesized on
e-mail: chzhang@tsinghua.edu.cn GCr15 steel in a low-pressure chemical vapor deposition (LPCVD) setup using a
gaseous mixture of acetylene and hydrogen onto a bearing steel substrate. An interlayer
of electroplated nickel was deposited on steel to assist and accelerate the graphene deposi-
Rakesh Kumar Gautam tion. The tribological performance was evaluated using a ball-on-disc tribometer with an
Department of Mechanical Engineering,
average Hertzian pressure of 0.2, 0.28, 0.34, and 0.42 GPa over a stroke length of 5 mm
Indian Institute of Technology (BHU),
against GCr15 steel ball and compared with bare steel and nickel-plated steel. The
Varanasi, Uttar Pradesh 221005, India
results indicate that the friction coefficient is dependent on the applied load and decrease
e-mail: rkg.mec@itbhu.ac.in
with increasing load, and the minimum friction coefficient of ∼0.13 was obtained for an
applied normal load of 1 N; however, the coating failed after 250 cycles. The decrease in
Rajnesh Tyagi1 friction coefficient has been attributed to the homogenization of the deposited multi-layer
Department of Mechanical Engineering,
graphene along the sliding direction and transfer of graphene to counter-face ball
Indian Institute of Technology (BHU),
leading to inhibition of metal-metal contact. The investigation suggests that this kind of
Varanasi, Uttar Pradesh 221005, India
coating has the potential of improving the tribological performance of metal-metal tribo-
e-mail: rtyagi.mec@itbhu.ac.in
pairs. [DOI: 10.1115/1.4047458]
Jianbin Luo Keywords: coatings, friction, films, solid lubricant, wear
Department of Mechanical Engineering,
State Key Laboratory of Tribology,
Tsinghua University,
Beijing 100084, China
e-mail: luojb@tsinghua.edu.cn

1 Introduction growth [11]. The former describes the physical deposition of


carbon after dehydrogenation to form graphene, whereas the later
Graphene is the hexagonal arrangement of sp2 carbon atoms
one involves the synthesis of graphene by the dissolution of
forming a honeycomb shape [1] and is increasingly being explored
carbon atoms in the metal at high temperature followed by segrega-
as a solid lubricant to improve the tribological performance of
tion during cooling. Nickel and copper have been reported to be the
mechanical components leading to energy saving and prolongment
popular transition metals for graphene growth [12]. Copper follows
of their life [2–5]. The chemical vapor deposition (CVD) growth of
the surface-growth mechanism and fits best for monolayer graphene
graphene on transition metals has been reported to be the most pre-
growth [13]. However, nickel supports the segregation-growth
ferred method to synthesize high-quality graphene over a large area
mechanism and suits well for few to multi-layer growth of graphene
[6,7]. Based on operating pressure, the different versions for CVD
[14,15]. Owing to higher carbon solubility and diffusivity, nickel, as
synthesis of graphene have been proposed, such as atmospheric-
a transition metal, allows a faster synthesis of graphene than copper
pressure chemical vapor deposition (APCVD), low-pressure chem-
[16]. Methane (CH4), ethylene (C2H4), and acetylene (C2H2) are the
ical vapor deposition (LPCVD), and ultrahigh vacuum chemical
most recognized gaseous carbon sources for graphene deposition.
vapor deposition (UHV-CVD) [8–10]. For a compromise between
Out of these, acetylene offers a low pyrolysis temperature, shortens
quality and cost, the LPCVD method has been reported as the
the reaction time, and promotes the graphene growth due to its
most promising method of graphene deposition. LPCVD graphene
higher surface adsorption energy [17].
coatings typically exhibit excellent uniformity, high purity, and
The coating of graphene on steel has shown promising results in
good coverage over the substrate. At the same time, low pressure
different applications because of its unique physical and mechanical
during CVD decreases the influence of residual gas (O, C, H) by
properties. Pu et al. [18] and Stoot et al. [19] reported that the
avoiding unwanted gas-phase reactions to contaminate the coatings.
graphene-covered stainless steel exhibits outstanding corrosion
Based on transition metal, the surface-growth and segregation-
resistance, and it can considerably increase the lifetime of future-
growth are two renowned CVD mechanisms to favor graphene
generation bipolar plates for fuel cells. Currently, the tribological
potential of graphene-based coatings has attracted much attention
1
of scientists and engineers working in the area of tribology.
Corresponding author.
Contributed by the Tribology Division of ASME for publication in the JOURNAL OF
Several studies have reported the improvement in friction and
TRIBOLOGY. Manuscript received January 29, 2020; final manuscript received May 14, wear behavior of CVD grown graphene in a variety of experimental
2020; published online June 19, 2020. Assoc. Editor: Min Zou. combinations. Different types of graphene coatings (such as

Journal of Tribology Copyright © 2020 by ASME DECEMBER 2020, Vol. 142 / 122101-1
monolayer, few-layer, multi-layer, and reduced graphene oxide 2 Materials and Methods
coatings) have been reported to facilitate friction and wear reduction
2.1 Synthesis and Characterization of Graphene Films.
[20]. Berman et al. [4] concluded that the SPG-coatings over steel
The circular disc-shaped (ɸ 25 mm and thickness 7 mm) GCr15
reduced the friction and wear significantly by forming a conformal
steel samples (hardness ≥ 60 HRC) were prepared by grinding
protecting layer on sliding interfaces. Restuccia and Righi [21]
with different grades of emery papers (600#, 1200#, 2400#, and
revealed that graphene can efficiently lubricate the steel-on-steel
4000#, respectively) followed by cloth polishing with diamond
microscale contacts in dry as well as in the humid atmosphere
paste to obtain a smooth and bright surface (surface roughness Rq
and attributed the improvement in tribological properties to passiv-
≤ 20 nm). The specimens were cleaned in an ultrasonic bath with
ating effect and intense tribo-chemical action of graphene with
acetone and ethanol for about 15 min each to remove the surface
native irons. Kim et al. [22] deposited the graphene oxide
impurities, dust, oil, and organic contaminants. A 3D profilometer
nano-sheet (GONS) coatings on silicon substrates via electrody-
(Zygo NexView, AMETEK Inc., CT) was used to confirm the
namic spraying process and reported a reduction in friction coeffi-
surface finish. Prior to deposition, a 25 µm thick nickel layer was
cient to a value of ∼0.1 attributing it to transfer of GONS
electroplated over the steel surface. The surface roughness (Rq) of
coatings on to counter surface. Romani et al. [23] reported direct
the disc surface was ∼35 nm after nickel electroplating. Table 1
growth of graphene on API X80 steel using ethanol as a precursor
lists the results of surface composition analyzed by energy-
at different partial pressures and recorded about three times lower
dispersive spectroscopy (EDS) for steel samples before and after
friction coefficient for steel partially covered with graphene than
nickel electroplating. The graphene films were synthesized by a
bare steel. Hong et al. [24] deposited graphene on silicon wafer
thermal CVD method with acetylene as a hydrocarbon source and
by electrophoretic deposition under different combinations of
hydrogen gas as a reduction medium. Figure 1(a) presents a sche-
applied voltage and deposition time and reported over 80% reduc-
matic diagram of CVD setup consisting of a gas supply unit, a
tion in friction coefficient. The graphene coatings with more com-
split tube furnace, and a vacuum system. The heating–soaking–
pacted microstructure were reported to be more effective and
cooling cycle followed to synthesize graphene is illustrated in
durable as a solid lubricant.
Fig. 1(b). After placing the nickel-coated steel sample at the
According to previous research findings, the use of graphene
center of the split tube furnace, the quartz tube was pumped
as a solid lubricant is limited, partially by the continuous growth
down to obtain a vacuum of order 1.2 × 10−2 Torr (base pressure,
of graphene films over a large area, but more likely by weak adhe-
Pbase) using a rotary vacuum pump. The temperature was raised
sion of transferred graphene films to an arbitrary substrate. The
to 850 °C from room temperature over a time span of 30 min.
present study is in continuation of our earlier published work [25]
The steel sample was annealed for the next 40 min in a hydrogen
in which the multi-layer graphene was grown on nickel catalyzed
(20 sccm) rich environment. Hydrogen performs the cleaning of
steel substrate, and the optimum conditions of growth were deter-
nickel surface and maintains sites for dehydrogenation of hydrocar-
mined. In the same work, the friction and wear behavior of graphene
bon, yielding diffusion, and carbon segregation [26]. The hydrocar-
coating under unidirectional (rotary) sliding also reported by con-
bon source (acetylene) was flown into the quartz tube with a flow
ducting tests at a fixed load of 0.5 N and constant sliding speed
rate of 6 sccm, and the total pressure (Pbase + PH2 + PC2H2) of the
of 0.07 m/s. The current investigation is a detailed study on the fric-
tube was around 3.20 × 10−1 Torr for a period of 10 min to allow
tion and wear characteristics of multi-layer graphene, deposited on
the desired reaction to take place. Acetylene gas got decomposed
nickel-plated steel by chemical vapor deposition using the opti-
into carbon and hydrogen atoms at a higher temperature, and
mized growth conditions from our previous work [25], under recip-
carbon atoms were absorbed by the sample surface during the reac-
rocating sliding motion and is aimed at exploring the tribological
tion stage. The flow of acetylene was terminated after completing
performance of this coating by carrying out friction tests at different
the reaction stage, while the flow of hydrogen gas was continued
normal loads of 0.1, 0.3, 0.5, and 1 N and a frequency of 1 Hz.
till the cooling of the furnace to room temperature. The synthesis
of graphene occurred due to the segregation of carbon atoms over
the surface during cooling.
The synthesized graphene samples were characterized by optical
Table 1 EDS analysis of surface composition before and after microscopy, scanning electron microscopy (SEM), Raman spectro-
nickel electroplating scopy, and X-ray photoelectron spectroscopy (XPS). The morphol-
ogy and structural properties of as-deposited graphene were studied
Chemical composition (wt%) through optical microscopy (VHX-600K, Keyence, Japan) and field
EDS area C O Fe Cr Ni Mn Si emission gun scanning electron microscopy (FEG-SEM). Raman
spectroscopy was performed using a high-resolution Raman spec-
Steel 1.20 1.63 93.19 1.81 0.86 0.85 0.46 trometer (LabRAM HR Evolution, Horiba Jobin Yvon, Japan)
Nickel-plated steel 0.74 0.77 0.71 0.36 96.77 0.41 0.24 with a laser excitation wavelength of 532 nm, laser power of
13 mW, scan range of 800–3000 cm−1, and scanning time of

Fig. 1 (a) Schematic diagram of the thermal CVD setup used for graphene synthesis and (b) summary of the
CVD-graphene synthesis process

122101-2 / Vol. 142, DECEMBER 2020 Transactions of the ASME


0.1 s per spot. The sample was focused using a 50× objective lens, characterization of synthesized graphene. Raman spectrum pre-
and the size of the focal spot of the laser was 1.25 µm in diameter. sented in Fig. 3 clearly indicates the presence of three characteristics
The XPS analyses were performed using Quantera SXM peaks identified as D, G, and 2D peaks centered around
(ULVAC-PHI, Kanagawa, Japan) with a monochromatic Al Kα ∼1350 cm−1, ∼1580 cm−1, and ∼2700 cm−1, respectively [28,29].
X-rays source of energy 1486.6 eV. The applied power and beam The presence of a single and sharp 2D peak in the Raman spectrum
diameter of X-rays were 23.56 W and 100.0 µm, respectively. shown in Fig. 3 clearly indicates that the deposited film is of gra-
The angle of incidence was set at 45 deg for X-rays. To examine phene rather than graphite, which has two components in 2D
the elemental distribution versus depth, etching was performed by peak, as reported earlier [28]. As evidenced in the figure, the G
a sputtering argon gun. Peak fitting allowing to decompose the band is higher than 2D band, and I2D/IG approaches a value of
XPS spectra in different components assigned to different surface ∼0.64, which suggests that the steel surface was covered by multi-
species was performed using the CasaXPS processing software layer graphene. There is one more weak peak at ∼2460 cm−1 (G*).
with a Shirley background. A high-resolution transmission electron The value of the ratio of ID/IG is 0.07, which indicates the growth of
microscope (HRTEM, Tecnai G2 20 TWIN, FEI, CT) was utilized the high quality of multi-layer graphene due to very low-intensity
to measure the thickness and to reveal the structure of as-deposited D-peak. The surface roughness of multi-layer graphene-coated
graphene. steel as measured by a noncontact 3D optical profilometer is
found to be ∼56 nm (Rq).
The transmission electron microscopy was also used to character-
2.2 Tribological Behavior of Graphene Films. The tribolog- ize the synthesized multi-layer graphene, and TEM images are pre-
ical behavior of synthesized graphene films was investigated by sented in Fig. 4. FIB in situ lift-out technique was used to prepare
using a ball-on-disc type of tribometer (UMT-5, Bruker, Karlsruhe, the cross-sectional samples for TEM. Metallic layers of Cr and Pt
Germany) against the GCr15 steel ball of 6 mm diameter. The disc were deposited on the surface to prevent the damage of multi-layer
was set to reciprocate against the stationary counterpart (ball) with graphene before the FIB process. The cross-sectional TEM of the
a frequency of 1 Hz (sliding speed of 0.01 m/s) over a stroke length synthesized multi-layer graphene shown in Fig. 4(a) indicates that
of 5 mm. The applied normal loads of 0.1, 0.3, 0.5, and 1 N were the thickness of multi-layer graphene is about 21 ± 2 nm. Another
used for friction tests, corresponding to the average Hertzian high-resolution TEM image in Fig. 4(b) is presented to identify
contact pressures of 0.2, 0.28, 0.34, and 0.42 GPa, respectively. the layered structure of graphene films.
To ensure the repeatability of friction and wear results, the sliding Elemental composition and binding energy of multi-layer gra-
tests were repeated at least three times for every condition. The mor- phene were determined by X-ray photoelectron spectroscopy.
phologies of the worn tracks on the disc and wear scar on the coun- Figure 5(a) shows a broad survey XPS spectrum of as-deposited
terpart (ball) were systemically analyzed by Raman spectroscopy, multi-layer graphene with an intense peak at 284.6 eV correspond-
optical microscopy, scanning electron microscopy (SEM) equipped ing to C1s (∼92 at%). Two other low-intensity peaks were also
with an energy dispersive spectroscopy (EDS), and transmission observed at 532 eV and 854 eV representing O1s (∼7.4 at%) and
electron microscopy. Ni2p (∼0.6 at%), respectively. The O1s peak refers to molecular
oxygen (O2) that originated from the air. Further, a de-convolution
of the C1s spectrum of multi-layer graphene into a main peak at
3 Results and Discussion 284.6 eV and three other peaks centered at 285.9, 287.7, and
289 eV, respectively, was performed and presented in Fig. 5(b).
Figure 2 depicts the micrograph of the multi-layer graphene-
Peak fitting was done using Voigt approximation (Gaussian–
coated steel surface, as examined under SEM. Graphene could be
Lorentzian). The peak at 284.6 eV has been assigned to C=C and
observed to be present in the form of patches on nickel-plated
stands for sp2-hybridized graphite-like carbon atoms. The peak at
steel. The growth of graphene on steel without a specific catalyst
285.9 eV represents the presence of the sp3-hybridized amorphous
is quite challenging [5], and hence, we opted electroplating of a
carbon atoms. The remaining two peaks at 287.7 eV and 289 eV
nickel layer over the surface, which accelerates the synthesis
may be attributed to C=O (ketone, aldehyde) and O–C=O (ester,
process. Graphene deposition follows the mechanism of precipita-
acid, carboxylic) groups, respectively [30,31]. The binding energy
tion and growth. According to this, the carbon from acetylene
of each peak and the quantitative evaluation of the fitted curve
gets dissolved into nickel to form a Ni–C solid solution at the
are presented in Table 2. Overall, the results confirm that the multi-
higher temperature during the reaction stage. At the time of
layer graphene films were synthesized successfully using low-
cooling, the solubility of carbon in nickel decreases, and the
carbon atoms precipitate out from the Ni–C solid solution during
cooling and coalesce to form graphene film [27].
Since Raman spectroscopy has been proven as one of the
most widely used techniques for nondestructive characterization
of carbon-based material, we selected it for the initial

Fig. 3 Raman spectrum of synthesized graphene at 850 °C with


a gas mixture of 6 sccm C2H2 and 20 sccm H2 for 10 min reaction
Fig. 2 SEM micrograph of multi-layer graphene-coated steel time

Journal of Tribology DECEMBER 2020, Vol. 142 / 122101-3


Fig. 4 (a, b) Cross-sectional TEM images of synthesized multi-layer graphene films under low
and high magnifications for revealing thickness and layered structure, respectively

Fig. 5 XPS of as-deposited multi-layer graphene: (a) a broad XPS spectrum and (b) C1s XPS
spectrum showing different peak fittings

pressure CVD with few-oxygen containing functional groups on the was abated to some extent by a nickel-plating over the steel
surface of alloy steel. surface, as obvious from the figure. The friction coefficient was
The other aspect of this work is to study the tribological perfor- also reduced slightly and oscillated at around 0.59. For the multi-
mance of multi-layer graphene-coated bearing steel under recipro- layer graphene-coated steel, the friction coefficient was reduced sig-
cating sliding motion. For this, a set of experiments with different nificantly and followed a very smooth curve all over the friction test.
sliding pairs, namely, steel versus steel (as a reference pair), steel The friction of multi-layer graphene-coated steel is about five times
versus nickel-coated steel and steel versus multi-layer graphene- lower than that measured for bare steel, which may be attributed to
coated steel, were performed in the air environment under a the lubricious effect of graphene [21].
normal load of 0.5 N. The friction coefficient was measured contin-
uously during the sliding test for 600 cycles. The results for the
coefficient of friction are plotted against wear cycles in Fig. 6.
One can observe that the coefficient of friction for bare steel was ini-
tially low and then increased with wear cycles after the removal of
the oxide layer presented on the bare steel surface. Severe fluc-
tuations recorded throughout the friction test may be attributed to
the accumulation of wear debris in between the contact surfaces,
which might have initiated three-body abrasion. The fluctuation

Table 2 Binding energy and quantitative analysis of peak


fittings presented in Fig. 5(b)

Bonding Position (eV) Area % Area

C=C 284.6 8788.23 73.10


C–C 285.9 2486.40 20.68
C=O 287.7 505.89 4.21
O–C=O 289 241.20 2.01 Fig. 6 Evolution of friction coefficient as a function of sliding
cycles for different friction pairs

122101-4 / Vol. 142, DECEMBER 2020 Transactions of the ASME


(a) (b)

Fig. 7 (a) Variation of friction coefficient of graphene-coated steel against steel as a function of
sliding cycles for different loads and (b) average coefficient of friction as a function of applied
normal load for bare steel, nickel-plated steel, and multi-layer graphene-coated steel

To examine the durability of multi-layer graphene coatings The average values of friction coefficient throughout the friction test
under different loads, the tribological behavior of multi-layer for bare steel, nickel-plated steel, and multi-layer graphene-coated
graphene-coated steel was also investigated for different normal steel was also plotted as a function of normal load, as presented
loads of 0.1, 0.3, 0.5, and 1 N. Figure 7 presents the variation in Fig. 7(b). The average values of the friction coefficient for
of friction coefficient recorded throughout the friction tests for graphene-coated steel before and after coating failure are also
varying loads. The friction coefficient is found to decrease with presented in Fig. 7(b).
increasing load. Under a normal load of 0.1 N, the average value The worn surfaces of disc, as well as counterpart ball, were
of the friction coefficient was recorded as ∼0.19. As the load analyzed using optical microscopy, SEM, and EDS to explore the
increased to 0.3 N and 0.5 N, the friction coefficient reduced to mechanisms of wear for different tribo-pairs. Figures 8(a)–8(c)
∼0.17 and ∼0.15, respectively. The minimum average friction coef- show the worn surfaces of steel, nickel-plated steel, and multi-layer
ficient of 0.13 was recorded for 1 N normal load, but the graphene graphene-coated steel after sliding under a load of 0.5 N, whereas
film could only sustain this value up to 250 cycles, and the friction Fig. 8(d) presents the surface of multi-layer graphene-coated steel
coefficient was found to shoot up sharply to a value of 0.9 beyond worn under a load of 1 N. The width of wear track for multi-layer
that as indicated in Fig. 7(a) before attaining a value of ∼0.6 for graphene-coated steel is significantly less than other. The multi-
the remaining test. The sudden increase in friction coefficient layer graphene coating over steel provides the protection to the
after 250 cycles strongly suggested the damage of graphene coatings. underlying substrate by avoiding the direct metal-metal contact

Fig. 8 Optical micrographs of wear track on the surface of (a) bare steel, (b) nickel-plated steel,
(c, d) multi-layer graphene-coated steel after the sliding under the normal loads of (a–c) 0.5 N and
(d ) 1 N

Journal of Tribology DECEMBER 2020, Vol. 142 / 122101-5


and utilizes the property of low interlayer shearing strength of gra- energy-dispersive spectrums of the whole micrographs. The figure
phene in affecting a reduction in friction. It can be noted that when also contains the elemental analysis of the entire surface as an
the applied normal load is 1 N, the graphene coating was damaged inset in the EDS spectrum. Figure 9(a) shows fine scratches along
significantly, and the material beneath is exposed to the counterpart, the sliding direction, which appear to be covered by a layer of
which causes a sudden increase in friction. oxide probably of iron, as evidenced by the EDS spectrum along
The SEM micrographs of worn-out as-received steel, nickel- with some wear debris particles. It also presents some places from
plated steel, and multi-layer graphene-coated steel specimens at a where the layer might have been detached, leading to the creation
load of 0.5 N are presented in Fig. 9 along with the corresponding of voids and micro-cracks. Figure 9(b) clearly suggests a substantial

Fig. 9 SEM micrographs and EDS spectrums with elemental analysis of the worn wear track surface of (a) bare steel, (b) nickel-
plated steel, and (c) multi-layer graphene-coated steel

122101-6 / Vol. 142, DECEMBER 2020 Transactions of the ASME


Fig. 10 SEM micrographs of the worn counterparts (balls) surface slid against (a) steel,
(b) nickel-plated steel, (c, d) multi-layer graphene-coated steel after the friction tests under the
normal loads of (a–c) 0.5 N and (d) 1 N

difference in the morphology of nickel-plated steel with more metal and inhibits the direct metal–metal contact leading to a reduc-
obvious adhesion on the worn surface. The surface is covered by tion in friction as shown in Fig. 7. This increase in carbon may have
a relatively thick and well compacted glazed layer of nickel come from the remaining fragments of multi-layer graphene after
oxide, as indicated by the EDS spectrum. The glazed layer has the friction tests, which was further supported by Raman spectro-
been shown to be effective in reducing both friction and wear scopy and transmission electron microscopy.
[32]. However, this layer appears to have been delaminated at The SEM images of the counterparts (balls) after the friction test
few places leading to exposure of the underlying substrate pointing at a load of 0.5 N are presented in Figs. 10(a)–10(c), whereas
toward adhesion and delamination wear. For multi-layer graphene- Fig. 10(d) shows the worn surface of the ball after testing at 1 N.
coated steel, the graphene, as a solid lubricant, might have It could be observed that wear scar is reduced significantly for
reduced the plastic deformation as well as adhesion wear, and the
abrasive wear appears to have dominated as evidenced from rela-
tively deep tracks in Fig. 9(c). The elemental analysis shown in
Fig. 9(c) suggests a significant increase in carbon content for multi-
layer graphene-coated steel, which appears in the form of a compact
layer over the surface, which gives protection to the underlying

Table 3 EDS analysis of the different areas of the ball, as


mentioned in Fig. 10, after the friction tests of 600 cycles

Chemical composition (wt%)

EDS area C O Fe Cr Ni Mn Si

Ball (blank) 0.90 1.18 93.73 1.99 0.87 0.86 0.47


Point 1 1.80 2.52 92.62 1.71 0.38 0.65 0.32
Point 2 1.29 4.75 83.36 1.55 8.55 0.29 0.21
Point 3 2.80 1.43 90.44 2.10 1.59 1.21 0.43
Point 4 17.71 1.18 77.53 1.66 0.87 0.75 0.30
Point 5 1.38 3.63 86.13 1.93 4.93 1.51 0.49 Fig. 11 Raman spectra of the wear track and wear scar after the
Point 6 5.07 3.68 84.66 2.12 3.14 0.91 0.42 friction test under 0.5 N load for multi-layer graphene-coated
steel

Journal of Tribology DECEMBER 2020, Vol. 142 / 122101-7


(a) (b)

Fig. 12 (a) Average wear scar diameter and (b) wear volume of the counterpart steel balls as a
function of applied normal load after the friction tests

multi-layer graphene-coated steel for friction test at a normal load of wear scar as well as for wear track, which indicates the increase
0.5 N in comparison with both bare and nickel-coated pair with no in defects of synthesized multi-layer graphene films. This increase
obvious loss of ball material. However, when the load reaches to a in defects points toward the damaging and disordering of gra-
value of 1 N, an increase in loss of ball material was observed due to phene during sliding [33]. In brief, the Raman spectrum confirms
metal–metal contact brought about by the failure of the coating after the presence of the graphene layer with a slight increase in
250 cycles. Table 3 gives the chemical composition of different defects.
areas of the wear scar as obtained through EDS. The area named The wear scar diameter for different combinations of friction tests
ball represents the chemical composition of the original ball. An was measured and presented in Fig. 12(a), and based on these diam-
increase in carbon percentage for the steel ball sliding against multi- eters, the wear volume of steel ball was calculated and is shown in
layer graphene-coated steel with minimum oxidation could be Fig. 12(b) as a function of the normal load. The volumetric loss of
observed, indicating the transfer of carbon (originated from multi- ball material, i.e., wear volume after the friction tests were calcu-
layer graphene) to the ball surface during sliding. lated based on the measured wear scar diameter using the following
The surface protection of the disc by graphene layers and the equation:
transfer of the layer on the ball during sliding is confirmed by   2 
πh 3D
Raman spectroscopy (Fig. 11) of the wear track as well as of V= + h2
wear scar. It can be observed that there is a sharp D-peak for 6 4

Fig. 13 Cross-sectional TEM images of wear scar (a, b) on the steel ball and wear track (c, d) on
the disc after the wear test under the normal load of 0.5 N

122101-8 / Vol. 142, DECEMBER 2020 Transactions of the ASME


where “D” is effective wear scar diameter and h is the height of Data Availability Statement
material removed from the ball. If R is the original ball radius
The datasets generated and supporting the findings of this article
used for the friction test, then h can be calculated as follows:
are obtainable from the corresponding author upon reasonable
 request. The authors attest that all data for this study are included
D2 in the paper. Data provided by a third party listed in Acknowledg-
h = R − R2 − ments. No data, models, or code were generated or used for this
4
paper.
Figure 12(b) displays the wear volume of steel ball against steel,
nickel-plated steel, and multi-layer graphene-coated steel, and it is
evident from Fig. 12(b) that wear of steel ball sliding against steel References
and nickel-plated steel becomes more and more severe with an [1] Geim, A. K., 2009, “Graphene: Status and Prospects,” Science, 324(5934),
increase in applied load. On the other hand, the wear of steel ball pp. 1530–1534.
sliding against multi-layer graphene-coated steel decreases compar- [2] Bhowmick, S., Banerji, A., and Alpas, A. T., 2015, “Role of Humidity in
Reducing Sliding Friction of Multilayered Graphene,” Carbon, 87, pp. 374–384.
atively. The increase in normal load from 0.1 N to 0.5 N led to a [3] Xu, S., Liu, Y., Gao, M., Kang, K. H., Kim, C. L., and Kim, D. E., 2018,
decrease in wear scar diameter and, subsequently, the wear “Selective Release of Less Defective Graphene During Sliding of an
volume for multi-layer graphene-coated steel. This reduction in Incompletely Reduced Graphene Oxide Coating on Steel,” Carbon, 134,
wear volume may be attributed to the homogenization and smooth- pp. 411–422.
[4] Berman, D., Erdemir, A., and Sumant, A. V., 2013, “Few Layer Graphene to
ening of graphene coatings during the sliding with the increase in Reduce Wear and Friction on Sliding Steel Surfaces,” Carbon, 54, pp. 454–459.
load [3]. However, for a load of 1 N, the wear of the steel ball [5] Ye, X., Lin, Z., Zhang, H., Zhu, H., Liu, Z., and Zhong, M., 2015, “Protecting
increased due to failure of multi-layer graphene coating after Carbon Steel From Corrosion by Laser In Situ Grown Graphene Films,”
∼250 cycles, and for the remaining cycles, the severity of wear Carbon, 94, pp. 326–334.
[6] Li, X., Cai, W., An, J., Kim, S., Nah, J., Yang, D., Piner, R., Velamakanni, A.,
increased due to a direct metal to metal contact. Undoubtedly, it Jung, I., Tutuc, E., Banerjee, S. K., Colombo, L., and Ruoff, R. S., 2009,
can be clearly observed that the wear for multi-layer graphene- “Large-Area Synthesis of High-Quality and Uniform Graphene Films on
coated steel up to the load of 0.5 N, is much lower than that of Copper Foils,” Science, 324(5932), pp. 1312–1314.
bare steel and nickel-plated steel, alike to the trends observed for [7] Huang, L., Chang, Q. H., Guo, G. L., Liu, Y., Xie, Y. Q., Wang, T., Ling, B., and
Yang, H. F., 2012, “Synthesis of High-Quality Graphene Films on Nickel Foils by
the coefficient of friction. Rapid Thermal Chemical Vapor Deposition,” Carbon, 50(2), pp. 551–556.
Further, TEM was used to characterize the wear track and wear [8] Vlassiouk, I., Fulvio, P., Meyer, H., Lavrik, N., Dai, S., Datskos, P., and Smirnov,
scar. Cross-sectional TEM samples were prepared using a similar S., 2013, “Large Scale Atmospheric Pressure Chemical Vapor Deposition of
procedure, as mentioned before for TEM of as-deposited multi- Graphene,” Carbon, 54, pp. 58–67.
[9] Li, X., Magnuson, C. W., Venugopal, A., Tromp, R. M., Hannon, J. B., Vogel,
layer graphene. TEM images of wear scar shown in Figs. 13(a) E. M., Colombo, L., and Ruoff, R. S., 2011, “Large-area Graphene Single
and 13(b) reveal the presence of a transferred tribo-layer, which Crystals Grown by Low-Pressure Chemical Vapor Deposition of Methane on
consists of fragments of graphene and amorphous carbon. The Copper,” J. Am. Chem. Soc., 133(9), pp. 2816–2819.
presence of a tribo-layer confirmed that the surface was well pro- [10] Mueller, N. S., Morfa, A. J., Abou-Ras, D., Oddone, V., Ciuk, T., and Giersig, M.,
2014, “Growing Graphene on Polycrystalline Copper Foils by Ultra-High
tected. The analysis of wear track reveals that the thickness of Vacuum Chemical Vapor Deposition,” Carbon, 78, pp. 347–355.
multi-layer graphene was reduced to 10 ± 2 nm and consists [11] Batzill, M., 2012, “The Surface Science of Graphene: Metal Interfaces, CVD
mainly of the fragments of graphene, as evident from Figs. Synthesis, Nanoribbons, Chemical Modifications, and Defects,” Surf. Sci. Rep.,
13(c) and 13(d). 67(3–4), pp. 83–115.
[12] Wintterlin, J., and Bocquet, M.-L., 2009, “Graphene on Metal Surfaces,” Surf.
Sci., 603(10–12), pp. 1841–1852.
[13] Li, X., Cai, W., Colombo, L., and Ruoff, R. S., 2009, “Evolution of Graphene
Growth on Ni and Cu by Carbon Isotope Labeling,” Nano Lett., 9(12),
4 Conclusions pp. 4268–4272.
[14] Yu, Q., Lian, J., Siriponglert, S., Li, H., Chen, Y. P., and Pei, S. S., 2008,
The present study was conducted for the evaluation of the tri- “Graphene Segregated on Ni Surfaces and Transferred to Insulators,” Appl.
bological potential of directly grown multi-layer graphene coating Phys. Lett., 93(11), p. 113103.
on GCr15 steel in rubbing against GCr15 steel ball under recip- [15] Reina, A., Jia, X., Ho, J., Nezich, D., Son, H., Bulovic, V., Dresselhaus, M. S.,
and Kong, J., 2009, “Large Area, Few-Layer Graphene Films on Arbitrary
rocating motion under different loads. The multi-layer graphene Substrates by Chemical Vapor Deposition,” Nano Lett., 9(1), pp. 30–35.
was successfully grown via the CVD method on GCr15 steel [16] Baraton, L., He, Z. B., Lee, C. S., Cojocaru, C. S., Châtelet, M., Maurice, J. L.,
by incorporating an interlayer of electroplated nickel. Multi-layer Lee, Y. H., and Pribat, D., 2011, “On the Mechanisms of Precipitation of
graphene films on steel enhance the tribological performance Graphene on Nickel Thin Films,” EPL, 96(4), p. 46003.
[17] Yang, M., Sasaki, S., Suzuki, K., and Miura, H., 2016, “Control of the Nucleation
significantly, and a transfer layer consisting of fragments of gra- and Quality of Graphene Grown by Low-Pressure Chemical Vapor Deposition
phene sheets and amorphous carbon was observed over the With Acetylene,” Appl. Surf. Sci., 366, pp. 219–226.
counter surface (ball). The results indicate that the friction coeffi- [18] Pu, N.-W., Shi, G.-N., Liu, Y.-M., Sun, X., Chang, J.-K., Sun, C.-L., Ger, M.-D.,
cient is dependent on the applied load and decreases with increas- Chen, C.-Y., Wang, P.-C., Peng, Y.-Y., Wu, C.-H., and Lawes, S., 2015,
“Graphene Grown on Stainless Steel as a High-Performance and Ecofriendly
ing load. The minimum friction coefficient of ∼0.13 was recorded Anti-Corrosion Coating for Polymer Electrolyte Membrane Fuel Cell Bipolar
for an applied normal load of 1 N; however, the coating failed Plates,” J. Power Sources, 282, pp. 248–256.
after 250 cycles. The decrease in friction coefficient has been [19] Stoot, A. C., Camilli, L., Spiegelhauer, S.-A., Yu, F., and Bøggild, P., 2015,
attributed to the homogenization of the deposited multi-layer gra- “Multilayer Graphene for Long-Term Corrosion Protection of Stainless Steel
Bipolar Plates for Polymer Electrolyte Membrane Fuel Cell,” J. Power Sources,
phene along the sliding direction and transfer of graphene films to 293, pp. 846–851.
counter-face ball leading to inhibition of metal–metal contact. [20] Zhai, W., Srikanth, N., Kong, L. B., and Zhou, K., 2017, “Carbon Nanomaterials
The results indicate that this kind of coating has excellent poten- in Tribology,” Carbon, 119, pp. 150–171.
tial for improving the tribological performance of metal–metal [21] Restuccia, P., and Righi, M. C., 2016, “Tribochemistry of Graphene on Iron and
Its Possible Role in Lubrication of Steel,” Carbon, 106, pp. 118–124.
tribo-pairs. However, more studies need to be conducted to eval- [22] Kim, H. J., Penkov, O. V., and Kim, D. E., 2015, “Tribological Properties of
uate the tribological behavior of these coatings under different Graphene Oxide Nanosheet Coating Fabricated by Using Electrodynamic
speeds and environments. Spraying Process,” Tribol. Lett., 57(3), p. 27.
[23] Romani, E. C., Larrude, D. G., Nachez, L., Vilani, C., de Campos, J. B., Peripolli,
S. B., and Freire, F. L., 2017, “Graphene Grown by Chemical Vapour Deposition
on Steel Substrates: Friction Behavior,” Tribol. Lett., 65(3), p. 96.
[24] Hong, H., Chen, S., Chen, X., Zhang, Z., and Shen, B., 2018, “Study on the
Acknowledgment Friction Reducing Effect of Graphene Coating Prepared by Electrophoretic
Deposition,” Proc. CIRP, 71, pp. 335–340.
The authors are thankful to the National Natural Science Founda- [25] Singh, S., Chen, X., Zhang, C., Gautam, R. K., Tyagi, R., and Luo, J., 2020,
tion of China (Grant No. 51975314) for financial support. “Nickel-catalyzed Direct Growth of Graphene on Bearing Steel (GCr15) by

Journal of Tribology DECEMBER 2020, Vol. 142 / 122101-9


Thermal Chemical Vapor Deposition and Its Tribological Behavior,” Appl. Surf. [30] Wang, X., You, H., Liu, F., Li, M., Wan, L., Li, S., Li, Q., Xu, Y., Tian, R., Yu,
Sci., 502, p. 144135. Z., Xiang, D., and Cheng, J., 2009, “Large-Scale Synthesis of Few-Layered
[26] Chen, C. S., and Hsieh, C. K., 2015, “Effects of Acetylene Flow Rate and Graphene Using CVD,” Chem. Vap. Deposition, 15(1–3), pp. 53–56.
Processing Temperature on Graphene Films Grown by Thermal Chemical [31] Ago, H., Kugler, T., Cacialli, F., Salaneck, W. R., Shaffer, M. S., Windle,
Vapor Deposition,” Thin Solid Films, 584, pp. 265–269. A. H., and Friend, R. H., 1999, “Work Functions and Surface Functional
[27] Lavin-Lopez, M. P., Valverde, J. L., Ruiz-Enrique, M. I., Sanchez-Silva, L., and Groups of Multiwall Carbon Nanotubes,” J. Phys. Chem. B, 103(38),
Romero, A., 2015, “Thickness Control of Graphene Deposited Over pp. 8116–8121.
Polycrystalline Nickel,” New J. Chem., 39(6), pp. 4414–4423. [32] Torgerson, T. B., Harris, M. D., Alidokht, S. A., Scharf, T. W., Aouadi, S. M.,
[28] Ferrari, A. C., Meyer, J. C., Scardaci, V., Casiraghi, C., Lazzeri, M., Mauri, F., Chromik, R. R., Zabinski, J. S., and Voevodin, A. A., 2018, “Room and
Piscanec, S., Jiang, D., Novoselov, K. S., Roth, S., and Geim, A. K., 2006, “Raman Elevated Temperature Sliding Wear Behavior of Cold Sprayed Ni-WC
Spectrum of Graphene and Graphene Layers,” Phys. Rev. Lett., 97(18), p. 187401. Composite Coatings,” Surf. Coat. Technol., 350, pp. 136–145.
[29] Nanda, S. S., Kim, M. J., Yeom, K. S., An, S. S. A., Ju, H., and Yi, D. K., 2016, [33] Zheng, D., Cai, Z. B., Shen, M. X., Li, Z. Y., and Zhu, M. H., 2016, “Investigation
“Raman Spectrum of Graphene With Its Versatile Future Perspectives,” TrAC, of the Tribology Behaviour of the Graphene Nanosheets as Oil Additives on
Trends Anal. Chem., 80, pp. 125–131. Textured Alloy Cast Iron Surface,” Appl. Surf. Sci., 387, pp. 66–75.

122101-10 / Vol. 142, DECEMBER 2020 Transactions of the ASME


Applied Surface Science 502 (2020) 144135

Contents lists available at ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

Full Length Article

Nickel-catalyzed direct growth of graphene on bearing steel (GCr15) by T


thermal chemical vapor deposition and its tribological behavior
⁎ ⁎
Sudesh Singha,b, Xinchun Chenb, Chenhui Zhangb, , Rakesh Kumar Gautama, Rajnesh Tyagia, ,
Jianbin Luob
a
Department of Mechanical Engineering, Indian Institute of Technology (BHU), Varanasi 221005, India
b
State Key Laboratory of Tribology, Tsinghua University, Beijing 100084, China

A R T I C LE I N FO A B S T R A C T

Keywords: The present work investigates the friction and wear behavior of graphene synthesized on bearing steel. To avoid
Graphene the deterioration in the quality of graphene during transfer and to avail the advantage of better adhesion be-
Tribology tween graphene and substrate, the graphene films were grown directly by thermal chemical vapor deposition on
Coating GCr15 bearing steel catalyzed with a layer of electroplated nickel. A gaseous mixture of H2 and acetylene (C2H2)
CVD
was used to synthesize the graphene films. The effect of growth temperature, reaction time and flow rate on the
formation of graphene films was investigated by Raman spectroscopy and scanning electron microscopy. The
results indicated that few to multilayer graphene can be grown across the surface of Ni/steel. It was concluded
that the graphene quality could be enhanced by optimizing the growth temperature, acetylene flow rate and
time. Further, comparative study of tribological behavior of steel, Ni/steel, and G/Ni/steel was performed
against steel ball for shorter as well as longer duration and it was concluded that graphene films over nickel
catalyzed steel can enhance the tribological performance effectively. Graphene films can reduce the friction
coefficient to 0.15 with 2–3 order reduction in wear.

1. Introduction method for accomplishing large area with high-quality graphene


growth on different substrates utilizing metal catalysts. Metal catalysts
The polymorphs of carbon in inorganic form, like graphite, dia- are required to interact with the hydrocarbons from the carbon source
mond, fullerene, CNTs, and graphene are gaining popularity in various for the synthesis of graphene films through CVD. The quality of gra-
fields such as biomedical, energy harvesting and storage industry, phene is dependent on the interaction of transition metal with a carbon
electronic, composite and coatings due to their superb characteristics precursor, which further decides the growth mechanism. Copper and
[1–4]. During the previous decade, these carbon-based nanomaterials nickel are most frequently used as metal catalysts for CVD synthesis of
have drawn much attention from researchers to be used as lubricant graphene, although other metal catalysts such as Ru, Rh, Re, Pd, Ir, Co,
additives, as well as a solid lubricant. Graphene, a 2D crystalline na- and Pt also have been used in several investigations [14]. Nickel offers
nosheet of carbon atoms packed into a honeycomb lattice material [5], important advantages over copper as a catalyst, such as stronger cata-
has emerged out as a promising candidate for better tribological per- lytic action for typical hydrocarbon precursors and subsequently, the
formance and hence, extending the active life of mechanical elements. graphene can be grown at comparatively low temperature [15].
However, moving forward towards to actualize the use of graphene in Since the carbon has a higher solubility in nickel, the synthesized
tribology applications relies heavily on the existence of a growth pro- graphene is mostly few to multi-layer. Graphene growth mechanism on
cess that can produce graphene in large scale with high quality. nickel catalyst consists of the decomposition of the carbon-based pre-
Graphene can be synthesized through a number of ways, such as cursor at high temperature, followed by carbon diffusion into the cat-
micromechanical cleavage [6], exfoliation of graphite intercalation alyst and precipitation over catalyst while cooling [16]. A lot of studies
compounds [7], unzipping carbon nanotubes [8], thermal decomposi- reported the influence of key factors such as growth temperature,
tion of bulk silicon carbide [9] and chemical vapor deposition [10–13]. growth time and hydrocarbon concentration (flow rate) in a CVD
Among all the processes that have been studied so far for graphene growth of graphene [17–22]. Chen et al. [23] proposed the low-tem-
synthesis, chemical vapor deposition (CVD) is the most promising perature growth of graphene films over pure nickel foil using C2H2 as a


Corresponding authors.
E-mail address: chzhang@tsinghua.edu.cn (C. Zhang).

https://doi.org/10.1016/j.apsusc.2019.144135
Received 19 March 2019; Received in revised form 5 July 2019; Accepted 20 September 2019
Available online 18 October 2019
0169-4332/ © 2019 Elsevier B.V. All rights reserved.
S. Singh, et al. Applied Surface Science 502 (2020) 144135

carbon source and further, regulated the C2H2 flow rate to improve the maintained. The samples were gradually heated from room temperature
quality of graphene. Yang et al. [24] synthesized graphene at different to various growth temperatures (from 650 °C to 950 °C, in the interval
flow rates of acetylene and studied its effect on domain size, defect of 100 °C) in 30 min and then annealed for 40 min under 20 sccm H2 gas
density, layer number and sheet resistance of graphene. Chae et al. [11] flow at the growth temperature. Annealing in hydrogen environment
reported the growth of highly crystalline few-layer graphene on a poly- helps in cleaning the native oxide over nickel layer and thus, promotes
Ni substrate with an optimized condition of temperature, growth time, the C2H2 dehydrogenation over nickel surface. The precursor i.e.
and gas mixing ratio (acetylene to hydrogen). Some wrinkles were acetylene (C2H2) was allowed to flow for 20 min at 6 sccm along with
observed in the graphene layers due to a difference in thermal expan- H2 (20 sccm). At higher temperature, the decomposed carbon atoms
sion of Ni and graphene layers. Gullapali, John and Pu et al. have re- from hydrocarbons were absorbed by the nickel. After completing the
ported the direct growth of graphene films on stainless steel, however, reaction stage, the acetylene gas was terminated and the samples were
these authors did not discuss the effect of various growth parameters allowed to cool down naturally to room temperature in the hydrogen-
and tribological performance of as-grown graphene films [25–27]. rich environment. During the cooling stage, the formation of graphene
Graphene, as a solid lubricant reduces the friction forces between occurs as a result of carbon precipitation over the surface. Further,
the contact surfaces on nano-, micro- and macro-scale while protecting some experiments were also accomplished with various acetylene flow
the coated surface. The solid lubrication effect of graphene is due to its rates (6, 8, and, 10 sccm) to grow graphene at 850 °C for 20 min and
lamellar structure with low interlayer shear strength. Berman et al. [28] keeping the other conditions the same. Graphene growth was also in-
studied wear performance of monolayer CVD graphene (transferred on vestigated for different growth time (10 and 20 min) at 850 °C under an
steel) in dry hydrogen environments and found that hydrogen passi- acetylene flow rate of 6 sccm while keeping the other conditions same.
vation of graphene helps to avoid disintegration of graphene and hence, Table 1 presents a brief introduction of the different conditions of
extending the wear life. Lee et al. [29] compared the nanoscale fric- temperature, flow rate, and reaction time variation for graphene
tional properties of atomically thin sheets of graphene to bulk graphite growth.
and reported that four layers of graphene sheets show almost similar
friction behavior as that of bulk. Kim et al. [30] reported that as-de-
posited multilayer graphene films over nickel can achieve the friction 2.2. Characterization of graphene films
coefficient as low as 0.03 at the microscale. This improved behavior
was explained by the tight bonding of graphene to the nickel substrate Graphene films on the surface were confirmed by a high-resolution
and low shear strength between graphene layers. Shi et al. [31] studied Raman spectroscopy (LabRAM HR Evolution, Horiba Jobin Yvon,
the tribological properties of few-layer graphene on textured M2 steel France), using a laser excitation wavelength of 532 nm, laser power of
surfaces with different groove area ratios and explained the enhance- 13 mW, the scan range of 800–3000 cm−1, the scanning time is 0.1 s
ment of wear resistance due to the entrapment of wear debris in grooves per spot. The size of the focal spot of the laser was 1.25 µm in diameter.
and the formation of a protective layer of carbon over the contact The morphology and structural properties of as-deposited graphene
surfaces. were studied by means of optical microscopy (VHX-6000, Keyence,
In the present study, we report the direct synthesis of graphene films Japan).
to avail the benefit of better adhesion between the graphene and sub-
strate and its tribological response. Graphene was synthesized by the 2.3. Tribological behavior of synthesized graphene films
thermal CVD method on GCr15 bearing steel electroplated with nickel
as a transition metal catalyst. The main benefit of direct growth is the The friction and wear behaviors of graphene films were explored by
avoidance of the possibility of degrading the quality of graphene during using a ball-on-disc type of a tribometer (UMT-5, Bruker, USA) against
the transfer process. Acetylene was used as carbon source to avail the GCr15 steel ball of 6 mm diameter. The disc was rotated against the
advantage of its low pyrolysis temperature. The growth temperature, stationary counterpart (ball) with an average sliding speed of 0.07 m/s
the flow rate of acetylene, and the reaction time were tuned. The effects under the applied normal load of 0.5 N (average Hertz contact pressure
of these parameters were also observed on the quality of graphene. The of 0.34 GPa). To confirm the repeatability of the obtained results, each
grown graphene was characterized by optical microscopy and Raman condition was repeated three times. After the friction experiments, the
spectroscopy. Friction and wear behavior has been investigated to ex- morphologies of the wear track on the disc and wear scar on the
amine the effectiveness of directly synthesized graphene improving the counterpart (ball) were systemically inspected by Raman spectroscopy,
tribological performance of tribo-pair. optical microscopy and scanning electron microscopy (SEM) equipped
with an energy dispersive spectroscopy (EDS). The wear volume of the
2. Materials and methods ball having a wear scar of effective diameter ‘D’ after tests were de-
termined as follows:
2.1. Synthesis of graphene films
πh 3D 2
V = ⎛ ⎞⎛ ⎜ + h2⎞ ⎟

For experiments, GCr15 alloy steel (hardness about 62 HRC) discs of ⎝ 6 ⎠⎝ 4 ⎠


24 mm diameter were used to synthesize graphene films. After me-
chanical grounding, the steel discs were polished and surface roughness where h, is the height of material removed from the ball and can be
of Rq ≤ 20 nm was measured by a non-contact 3D optical surface calculated as follows:
profiler (Zygo NexView, AMETEK Inc., USA). The discs were immersed
in acetone and isopropanol, respectively, and sonicated to get a con- D2
h=R− R2 −
tamination free surface. A nickel layer (25 µm thick) was electroplated 4
on the surface of steel disc to facilitate the graphene deposition. The
graphene films were prepared by a thermal CVD method with acetylene where R, is the original ball radius used for the friction test.
as a carbon source and hydrogen gas as a reduction medium. A sche- Further, the wear rate was calculated using the following equation:
matic diagram of CVD set up and a summarized temperature-time cycle V
of a typical CVD growth of graphene are shown in Fig. 1. After posi- wear rate =
d ·L
tioning the discs in a ceramic boat, it was pushed inside the quartz to
place in the center of the furnace. Then, a vacuum of the order of where V, L, and d are the total wear volume, normal load, and the total
1.2 × 10−2 Torr was accomplished by using a vacuum pump and sliding distance, respectively.

2
S. Singh, et al. Applied Surface Science 502 (2020) 144135

Fig. 1. (a) Schematic representation of the thermal CVD set-up used for graphene synthesis, (b) Various stages of the CVD-graphene synthesis process.

Table 1
Different conditions of graphene growth over nickel catalyzed steel.
Effects Temperature (°C) Reduction stage Reaction stage Cooling stage

Temperature variation 650 20 sccm (H2) for 40 min 6 sccm (C2H2) for 20 min H2(20 sccm)
750
850
950
Acetylene flow rate variation 850 6 sccm (C2H2) for 20 min
8 sccm (C2H2) for 20 min
10 sccm (C2H2) for 20 min
Reaction time variation 850 6 sccm (C2H2) for 10 min
6 sccm (C2H2) for 20 min

3. Results and discussion and independent of G peak [32,33]. There occur two more weak peaks
at ~2460 cm−1 (G*) and ~2944 cm−1 (D + D′) attributed to the inter-
3.1. Effect of temperature on growth of CVD graphene valley double resonant Raman process and the combination of D and G
peaks respectively [26]. The ratio of ID/IG is used to evaluate the quality
Since, the growth temperature is the key parameter for the growth of graphene. Lower the ID/IG ratio, the higher the quality of graphene
of high-quality graphene, the effect of growth temperature on the due to a decrease in D-peak intensity. The I2D/IG is directly relevant to
synthesis of graphene was investigated. Here, the growth temperature the number of graphene layers (the value of I2D drop with an increase in
was varied between 650 °C and 950 °C, while the reaction time was the number of the layers).
fixed to 20 min with an acetylene flow rate of 6 sccm followed by air Although the steel specimen with pre-plated nickel exhibited a
cooling. Graphene can be grown on nickel at a lower temperature than Raman signal indicative of graphene at every temperature, it can be
that required for copper. The CVD growth of graphene on Ni/steel can observed (Fig. 3a) that a strong D peak was observed for a growth
be described in terms of the diffusion of carbon atoms into the elec- temperature of 650 °C, which indicates the presence of a significant
troplated Ni layer at high temperature, followed by segregation on the number of defects in graphene. The poor graphitization at low tem-
Ni surface during cooling. According to phase diagrams, the solubility perature is the reason behind the strong D peak. A decrease in D peak
of carbon in nickel increases at higher temperature. A solid solution of intensity was recorded with an increase in temperature and D peak is
nickel and carbon was formed beyond 500 °C due to absorption of almost absent above a growth temperature of 850 °C, which confirms
carbon atoms into the nickel layer. the growth of high-quality graphene. Fig. 3b presents the ID/IG and I2D/
Fig. 2 presents the optical microscopic images of synthesized gra- IG ratios obtained from Fig. 3a. The ID/IG ratios for the graphene grown
phene at various temperatures. It is not possible to grow graphene with at 650 °C, 750 °C, 850 °C, 950 °C are 1.56, 1.12, 0.04 and 0.06 re-
a precise and uniform layering throughout the surface using nickel as a spectively. The I2D/IG ratios for the graphene grown at 650 °C, 750 °C,
transition metal. Here, we select the mostly occurred region in each 850 °C, 950 °C are 0.36, 0.42, 0.58 and 0.44 respectively. This ratio
surface as a representative micro-images of graphene grown at different increases from 650 °C to 850 °C and further increasing the temperature
growth temperatures. From the figure, it can be observed that the to 950 °C cause a decrease in the ratio. As I2D/IG value ~1 is the re-
graphene synthesized in the form of patches with varying thickness. It is presentative of bilayer graphene, the I2D/IG ratios present that the steel
well known that the carbon nano-materials are best characterized by surface was covered by few-layer graphene. Since the lowest ID/IG and
Raman spectroscopy. Fig. 3a presents the Raman spectra of graphene highest I2D/IG was observed for a growth temperature of 850 °C, we
grown on surfaces at different growth temperatures. It can be observed selected this temperature as optimum temperature to study further the
in the Raman spectrum that there exists mainly a D peak, a G peak, and, effect of acetylene flow rate and reaction time. Fig. 4a and b present the
a 2D peak. The D peak represents zone-boundary phonons, which can surface roughness after nickel electroplating and after CVD growth of
be identified at ~1350 cm−1 in defected graphite. The defect can be in graphene at the growth temperature of 850 °C, respectively. It is ob-
the form of points defects, dislocation like defects, bending and over- vious that roughness increases after the graphene deposition and ap-
lapping of layers, substitutional impurities, and carbon adatoms, etc. proaches to a value of ~56 µm (Rq) as shown in Fig. 4b.
The G peak is representative of doubly degenerate phonon made of sp2
carbon bonds and occurs at ~1580 cm−1. The 2D peak, centered at
~2700 cm−1, is caused by the second order of zone-boundary phonon

3
S. Singh, et al. Applied Surface Science 502 (2020) 144135

Fig. 2. Optical images of CVD-grown graphene for various growth temperatures: (a) 650 °C, (b) 750 °C, (c) 850 °C, and (d) 950 °C under an acetylene flow rate of 6
sccm for 20 min.

3.2. Effect of acetylene flow rate on the growth of CVD graphene precipitation. The decrease in 2D peak may be due to the availability of
more and more carbon atoms at the increased flow rate to be dissolved
In order to understand the influence of flow rate on the growth of into Ni/steel, further their precipitation over the surface during the
graphene on Ni/steel surface, the flow rate of acetylene was varied cooling and subsequently increasing the number of graphene layers.
during the reaction stage to grow the graphene keeping the H2 flow rate I2D/IG ratios were reported as 0.58, 0.37 and 0.28 for corresponding
constant. The growth temperature was fixed to 850 °C for each condi- acetylene flow rate of 6, 8 and 10 sccm, while ID/IG remains under 0.05
tion of flow rate variation. Fig. 5 shows the Raman spectra and peak for each and every flow rate. Thus, low defect graphene can be grown
intensity ratios for different acetylene flow rates. With increase in with reduced acetylene flow rate [19]. Since the flow rate was not
acetylene flow rate, the peak position of 2D peak shifts from 2709 cm−1 varied by a large amount, there was only a slight increase in defect was
(6 sccm) to 2714 cm−1 (8 sccm), 2719 cm−1 (10 sccm). Increasing the recorded.
acetylene flow rate also causes the suppression of 2D and G peak in-
tensities and strengthens the D peak. This is mainly due to the presence
of sp3 carbon atoms simultaneously with sp2 carbon atoms after

Fig. 3. Raman spectra (a) and the ratios of ID/IG and I2D/IG (b) of graphene grown on Ni/steel at various growth temperatures under a gas mixture of 6 sccm C2H2 and
20 sccm H2 for 20 min.

4
S. Singh, et al. Applied Surface Science 502 (2020) 144135

Fig. 4. Surface roughness measured by WLI of (a) Ni/steel and (b) G/Ni/steel-850 °C.

3.3. Effect of reaction time on the growth of CVD graphene Overall, the result indicates that the quality of grown graphene can
be controlled by regulating the reaction parameters and the optimized
Reaction time has also a major influence on the growth of high- condition for graphene over nickel was obtained at a growth tem-
quality graphene. We investigated the effect of reaction time (10 and perature of 850 °C for 10 min reaction time under an acetylene flow rate
20 min), while the growth temperature and acetylene flow rate were of 6 sccm. Further, the tribological behavior of the optimized grown
fixed at 850 °C and 6 sccm respectively. An optical image of synthesized graphene films was investigated for shorter as well as for longer dura-
graphene for 10 min and 20 min is shown in Fig. 6a and Fig. 2c. Fig. 6b tion.
presents the Raman spectra of graphene on Ni/steel for various reaction
time. It can be seen that strong G and 2D peak signals appear for all 3.4. Friction and wear behavior of grown graphene during initial sliding
growth times. The shorter growth time 10 min is preferred since too
long a growth time may also expose the crystals to etching [11,21]. The friction and wear behavior of graphene coated bearing steel
With decreasing growth time, the peak position of 2D peak shifts from were investigated for an initial test of 800 cycles. The friction tests were
2709 cm−1 to 2701 cm−1, while the intensity of 2D peak increases. carried out in an air environment under a load of 0.5 N (average Hertz
Based on measured data, I2D/IG ratio was calculated and the results contact pressure of 0.34 GPa) for the contact between the rotating steel
indicated that the I2D/IG ratio increased slightly with the decrease in disc and stationary steel ball. As shown in Fig. 7, the coefficient of
reaction time from 20 min to 10 min. This concludes the reduction in friction of steel ball against bare steel is unstable and highly fluctuating.
the number of layers of graphene as the reaction time reduces. The These fluctuations in friction behavior are attributed to the accumula-
comparatively thinner graphene films can be synthesized at lower de- tion of wear debris as a third body between the friction surfaces. For the
position time. These results are identical to earlier investigations for the initial sliding of 800 cycles, the friction coefficient was highest for steel
growth of graphene on Ni substrates by CVD [34]. With the increase in against steel and was approximate 0.89. Electroplating of nickel over
reaction time, more numbers of carbon atoms may available to be steel disc can reduce friction coefficient to 0.66. A significant reduction
dissolved into Ni/steel. These dissolved carbon atoms precipitated in friction coefficient can be observed for graphene grown on steel disc
while cooling and subsequently increasing the number of graphene (G/Ni/steel), which can be explained by the lubrication effect of gra-
layers. An increase in the intensity of D peak was also observed for short phene. An average coefficient of friction of 0.15 can be achieved
reaction time. throughout the friction test of 800 cycles.

Fig. 5. Raman spectra (a) and the ratios of ID/IG and I2D/IG (b) of graphene grown on Ni/steel for various acetylene flow rates at 850 °C growth temperature and
20 min reaction time.

5
S. Singh, et al. Applied Surface Science 502 (2020) 144135

Fig. 6. (a) Optical image of CVD-grown graphene for 10 min reaction time at 850 °C under an acetylene flow rate of 6 sccm and (b) Raman spectra of graphene grown
on Ni/steel at 850 °C under a gas mixture 6 sccm C2H2 and 20 sccm of H2 for 10 and 20 min growth time, respectively.

Fig. 8 shows the optical micrographs of friction pairs after the test to counter-body occurred during the sliding process. Some rugged areas
put across another important aspect, i.e., the effect of graphene on wear can be clearly observed in the wear track of steel. For Ni/steel (Fig. 9b),
performance of steel. For each friction pair, it can be observed that the the worn surface is relatively less worn out without significant de-
wear scar diameter is larger than respective wear track because the ball formation and containing comparatively smaller adhesion of wear
is subjected to higher Hertzian pressure with regular contact with debris and reaction products (oxides). For G/Ni/steel (Fig. 9c), almost
counter surfaces, while the wear track comes in contact periodically. negligible adhesion of wear debris was observed. The wear track was
The observation of wear surfaces revealed that bare steel and Ni/steel relatively smooth and mainly abrasion wear was observed with the
suffered severe damage. Wear scar diameter and wear track length for evidence of grooves along the sliding direction. The results of EDS
steel against steel was observed 545.03 µm and 428.69 µm, respec- analysis performed to get the chemical composition inside the wear
tively. Graphene coating on steel is able to reduce the wear significantly tracks are shown in Table 2. It can be seen that steel and Ni/steel
to 125.88 µm and 89.21 µm respectively for wear scar and wear track. surfaces were strongly oxidized, while G/Ni/steel exhibited the
Low interlayer shear strength and the strong adhesion between the minimum oxidation. A comparison of EDS result of point 1 and 2 of ball
graphene and surface contribute to friction reduction as well as wear indicates that a transfer layer of carbon was formed on the counter ball
reduction significantly. The wear volume and wear rate of steel ball surface after the friction tests.
against steel, Ni/steel, and G/Ni/steel are displayed in Fig. 7b. It can be Fig. 10 presents the typical Raman spectra of worn out surfaces after
seen clearly that the steel ball against steel has undergone the max- the friction tests of 800 cycles. Raman spectra from different locations
imum wear. By all of the tested pairs, the steel ball against G/Ni/steel on wear track are presented in Fig. 10b with respective locations
shows exceptional anti-wear performance. marked on Fig. 10a. Points 1 and 10 are outside the wear tracks, while
Fig. 9 shows SEM images of wear tracks on the disc and wear scar on other points are inside the wear track. The clear presence of D, G, and
the ball. Severe damage of the steel surfaces was observed, while G/Ni/ 2D peak in the Raman spectrum indicate that the graphene still existed
steel suffered the minimum wear. In the SEM micrographs of bare steel over the steel surface even after sliding of 800 cycles. Although an in-
Fig. 9a, the wear includes mainly adhesive wear, abrasive wear, crease in D peak intensity was also observed for each and every point.
plowing and wear induced oxidation. Adhesion wear involves the de- This concludes an increase in the defect in the graphene during the
tachment of material from the ball as well as disc and transfer to the sliding process. Further, the Raman spectrum was also measured for

Fig. 7. The coefficients of friction as a function of sliding cycles (a) and wear volumes and wear rates (b) of steel ball against bare steel, Ni/steel, G/Ni/steel for 800
cycles.

6
S. Singh, et al. Applied Surface Science 502 (2020) 144135

Fig. 8. Optical micrographs of wear morphologies of ball (a, b, c) and disc (d, e, f) counter-pairs for steel against steel, steel against Ni/steel, steel against G/Ni/steel,
respectively after the friction tests for 800 cycles.

wear scar over the ball (Fig. 10d), which confirms the transfer of gra- protective layer could prevent direct contact at the interface of tribo-
phene layers over the ball surface. pairs. As a result, the coefficient of friction as low as 0.15 can be con-
Based on the analysis of worn surfaces after the friction tests, a wear tinued throughout the friction test with minimum wear, while com-
mechanism for explaining the solid lubrication effect can be proposed. paring to other friction pairs.
Low shear stress between the graphene layers is mainly the reason for
such low friction. On the other hand, a partial transfer of graphene
layers occurs to the counter surface (ball) during the sliding process. 3.5. Friction and wear behavior of grown graphene for long duration sliding
Thus, a protective layer was formed over the counter surface. This
To investigate the sturdiness over time of graphene film, the

Fig. 9. SEM images of wear track of bare steel (a), Ni/steel (b), G/Ni/steel (c), and steel ball wear scar against G/Ni/steel (d) after the friction test of 800 cycles.

7
S. Singh, et al. Applied Surface Science 502 (2020) 144135

Table 2 different combinations of tribo-pairs. It is clear from the figure that the
EDS analysis of wear surface after the friction test of 800 cycles. high friction coefficient causes substantial wear of rubbing surfaces for
EDS area Chemical composition (wt.%) steel against steel. A wear scar diameter of 850 µm and a wear track
width of 690 µm were recorded for steel against steel, while these va-
C O Fe Cr Ni Mn Si lues reduced to 167 µm and 149 µm, respectively for G/Ni/steel. The
wear volume and wear rate of the ball against different counter surfaces
Steel 1.89 11.39 83.54 1.82 0.47 0.71 0.18
Ni/steel 3.07 6.79 3.83 0.37 85.53 0.33 0.08 are presented in Fig. 11b. It can be seen clearly that the minimum wear
G/Ni/steel 7.81 2.72 0.72 0.50 87.46 0.44 0.35 volume and wear rate occur for steel against G/Ni/steel while the
Ball point 1 2.01 1.81 92.48 2.06 0.70 0.58 0.30 maximum for steel against steel.
Ball point 2 0.90 1.18 93.73 1.99 0.87 0.86 0.47 Fig. 13 presents the typical Raman spectra of worn out surfaces of
the ball and graphene-coated steel after the friction tests of 5600 cycles.
The presence of D, G, and 2D peak in the Raman spectra indicate that
tribotests were conducted for a longer duration at 0.5 N normal load in
the graphene still exists over the steel surface and ball even after sliding
the ambient environment for 5600 cycles. Fig. 11a presents the friction
of 5600 cycles. Although an intense D peak occurs for both ball and
behavior of steel against steel, steel against Ni/steel and, steel against
plate, which indicates a significant increase of defects in graphene.
G/Ni/steel tribo-pairs. For steel against steel, the friction coefficient
Table 3 presents the ball wear volume and wear rate for short (800
remains unsteady and fluctuating with an average value of 0.8. It can be
cycles) and long (5600 cycles) duration friction tests for a normal load
observed from Fig. 11a that for the initial 1000 cycles, the friction
of 0.5 N. The tabulated values show that electroplating of nickel over
coefficient of steel against Ni/steel is lower than steel against steel, but
steel reduces the ball wear by 4.4 times for short duration (800 cycles)
after that, it increases slightly when compared to the average friction
friction tests, while for long duration (5600 cycles) friction tests, this
coefficient of steel against steel. A significant reduction in friction
reduction is only 1.5 times. The CVD growth of graphene over nickel
coefficient can be observed for steel against G/Ni/steel and an average
electroplated steel can reduce the wear significantly by 2 to 3 orders
value of 0.18 was recorded for friction coefficient throughout the fric-
lower than bare steel for short and long duration as well. The reduction
tion test, which is slightly more than that of short duration sliding. This
in wear volume justifies the solid lubrication effect of graphene in an air
increase is mainly due to the loss of the beneficiary effect of graphene
environment.
with time.
Fig. 12 presents the rubbing surfaces after the friction tests for

Fig. 10. Optical micrographs and the corresponding Raman analysis of wear track (a, b) and wear scar (c, d) for 800 cycles.

8
S. Singh, et al. Applied Surface Science 502 (2020) 144135

Fig. 11. (a) Coefficients of friction as a function of sliding cycles and (b) wear volumes and wear rates of steel ball against steel, Ni/steel, G/Ni/steel for 5600 cycles.

4. Conclusions

Graphene films were synthesized on bearing steel directly by


thermal CVD assisted with an electroplated nickel layer. The study re-
ported the effects of growth temperature, acetylene flow rate and re-
action time on graphene growth over nickel catalyzed steel substrate.
Electroplating was chosen as a cheap and quick alternative to coat a
nickel layer over steel. The quality of graphene films was effectively
improved with the increase in temperature. Growth temperature,
acetylene flow rate and reaction time were optimized to control the
quality of graphene. Further, the tribological response of graphene
coated steel against steel was studied. Graphene films on steel serve as a
protective layer, which can reduce friction as well as wear significantly
compared to high and fluctuating friction of bare steel. For the initial
800 cycles, the average friction coefficient can be lowered to a value of
0.15 under the protection from graphene films. Low shear strength of
the interlayer sliding is the reason for such a low friction. Directly
grown graphene layers stick firmly on the surface and can protect the Fig. 13. Raman spectra of the wear track (plate) and wear scar (ball) for 5600
steel for more than 5600 cycles with an average friction coefficient of cycles.
0.18. Moreover, the wear volume for short and long duration tests are
reduced reasonably, i.e., 2–3 orders less in magnitude than that of bare

Fig. 12. Optical micrographs of wear morphologies of ball (a, b, c) and disc (d, e, f) counter-pairs for steel against steel, steel against Ni/steel, steel against G/Ni/
steel, respectively, after the friction test for 5600 cycles.

9
S. Singh, et al. Applied Surface Science 502 (2020) 144135

Table 3
Ball wear volume calculations for 800 and 5600 cycles friction test at 0.5 N load.
Experiment conditions Wear volume (V) (800 cycles) (mm3) Wear rate (V/(d∙L)) Wear volume Wear rate (V/(d∙L))
(800 cycles) (mm3/N∙m) (V) (5600 cycles) (mm3) (5600 cycles) (mm3/N∙m)

Steel Vs. steel 1.448 × 10−3 8.23 × 10−5 8.591 × 10−3 4.65 × 10−5
Steel Vs. Ni/steel 3.29 × 10−4 1.87 × 10−5 5.404 × 10−3 2.92 × 10−5
Steel Vs. G/Ni/steel 4.11 × 10−6 2.34 × 10−7 1.28 × 10−5 6.94 × 10−8

steel. quality graphene at low temperatures, ACS Nano 6 (2012) 9996–10003, https://
Overall, our work presents direct growth of graphene films over doi.org/10.1021/nn303674g.
[16] X. Li, W. Cai, L. Colombo, R.S. Ruoff, Evolution of graphene growth on Ni and Cu by
steel to avoid the problem regarding the graphene damage during carbon isotope labeling, Nano Lett. 9 (2009) 4268–4272, https://doi.org/10.1021/
transfer and demonstrates the role of graphene films as effective solid nl902515k.
lubricants for a considerably longer duration. [17] K. Al-Shurman, H. Naseem, CVD graphene growth mechanism on nickel thin films,
Excerpt from Proc. 2014 COMSOL Conf. Bost. 446 (2014) 7131. doi: 10.1063/1.
2836265.
Acknowledgment [18] N. Ramli, N.A. Nayan, H.W. Lee, S.S. Embong, Analysis of the effect of growth
parameters on graphene synthesized by chemical vapor deposition, J.
Nanoelectron. Optoelectron. 10 (2015) 50–55, https://doi.org/10.1166/jno.2015.
This work is partially supported by the National Natural Science
1689.
Foundation of China (Grant No. 51605247). We acknowledge Technos [19] M. Regmi, M.F. Chisholm, G. Eres, The effect of growth parameters on the intrinsic
Instruments for providing the CVD system for the graphene growth. properties of large-area single layer graphene grown by chemical vapor deposition
on Cu, Carbon N. Y. 50 (2012) 134–141, https://doi.org/10.1016/j.carbon.2011.
07.063.
References [20] Y. Yao, Z. Li, Z. Lin, K.S. Moon, J. Agar, C. Wong, Controlled growth of multilayer,
few-layer, and single-layer graphene on metal substrates, J. Phys. Chem. C. 115
[1] S. Bahrami, A. Solouk, H. Mirzadeh, A.M. Seifalian, Electroconductive poly- (2011) 5232–5238, https://doi.org/10.1021/jp109002p.
urethane/graphene nanocomposite for biomedical applications, Compos. Part B [21] R. Papon, C. Pierlot, S. Sharma, S.M. Shinde, G. Kalita, M. Tanemura, Optimization
Eng. 168 (2019) 421–431, https://doi.org/10.1016/j.compositesb.2019.03.044. of CVD parameters for graphene synthesis through design of experiments, Phys.
[2] X. Ma, Y. Liu, H. Liu, L. Zhang, B. Xu, F. Xiao, Fabrication of novel slurry containing Status Solidi Basic Res. 254 (2017), https://doi.org/10.1002/pssb.201600629.
graphene oxide-modified microencapsulated phase change material for direct ab- [22] D. Su, M. Ren, X. Li, W. Huang, Synthesis of graphene by chemical vapor deposition:
sorption solar collector, Sol. Energy Mater. Sol. Cells 188 (2018) 73–80, https://doi. effect of growth conditions, J. Nanosci. Nanotechnol. 13 (2013) 6471–6484,
org/10.1016/j.solmat.2018.08.021. https://doi.org/10.1166/jnn.2013.7920.
[3] A. Sarkar, A.K. Chakraborty, S. Bera, NiS/rGO nanohybrid: an excellent counter [23] C.S. Chen, C.K. Hsieh, Effects of acetylene flow rate and processing temperature on
electrode for dye sensitized solar cell, Sol. Energy Mater. Sol. Cells 182 (2018) graphene films grown by thermal chemical vapor deposition, Thin Solid Films. 584
314–320, https://doi.org/10.1016/j.solmat.2018.03.026. (2015) 265–269, https://doi.org/10.1016/j.tsf.2014.12.012.
[4] S. Yu, L. Zhao, R. Liu, C. Zhang, H. Zheng, Y. Sun, L. Li, Performance enhancement [24] M. Yang, S. Sasaki, K. Suzuki, H. Miura, Control of the nucleation and quality of
of Cu-based AZO multilayer thin films via graphene fence engineering for organic graphene grown by low-pressure chemical vapor deposition with acetylene, Appl.
solar cells, Sol. Energy Mater. Sol. Cells 183 (2018) 66–72, https://doi.org/10. Surf. Sci. 366 (2016) 219–226, https://doi.org/10.1016/j.apsusc.2016.01.089.
1016/j.solmat.2018.04.008. [25] H. Gullapalli, A.L. Mohana Reddy, S. Kilpatrick, M. Dubey, P.M. Ajayan, Graphene
[5] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, M.I. Katsnelson, I.V. Grigorieva, growth via carburization of stainless steel and application in energy storage, Small 7
S.V. Dubonos, Two-dimensional gas of massless Dirac fermions in graphene 438 (2011) 1697–1700, https://doi.org/10.1002/smll.201100111.
(2005) 197–200. doi: 10.1038/nature04233. [26] R. John, A. Ashokreddy, C. Vijayan, T. Pradeep, Single-and few-layer graphene
[6] K.S. Novoselov, Electric field effect in atomically thin carbon films, Science (80-.) growth on stainless steel substrates by direct thermal chemical vapor deposition,
306 (2004) 666–669, https://doi.org/10.1126/science:1102896. Nanotechnology 22 (2011), https://doi.org/10.1088/0957-4484/22/16/165701.
[7] C. Vallés, C. Drummond, H. Saadaoui, C.A. Furtado, M. He, O. Roubeau, L. Ortolani, [27] N.W. Pu, G.N. Shi, Y.M. Liu, X. Sun, J.K. Chang, C.L. Sun, M. Der Ger, C.Y. Chen,
M. Monthioux, A. Pé nicaud, Solutions of negatively charged graphene sheets and P.C. Wang, Y.Y. Peng, C.H. Wu, S. Lawes, Graphene grown on stainless steel as a
ribbons 09 (2018) 17. doi: 10.1021/ja808001a. high-performance and ecofriendly anti-corrosion coating for polymer electrolyte
[8] L. Jiao, L. Zhang, X. Wang, G. Diankov, H. Dai, Narrow graphene nanoribbons from membrane fuel cell bipolar plates, J. Power Sources. 282 (2015) 248–256, https://
carbon nanotubes, Nature 458 (2009) 877–880, https://doi.org/10.1038/ doi.org/10.1016/j.jpowsour.2015.02.055.
nature07919. [28] D. Berman, S.A. Deshmukh, S.K.R.S. Sankaranarayanan, A. Erdemir, A.V. Sumant,
[9] H. Hibino, H. Kageshima, M. Nagase, Epitaxial few-layer graphene: towards single Extraordinary macroscale wear resistance of one atom thick graphene, Layer.
crystal growth, J. Phys. D. Appl. Phys. 43 (2010), https://doi.org/10.1088/0022- (2014) 6640–6646, https://doi.org/10.1002/adfm.201401755.
3727/43/37/374005. [29] C. Lee, Q. Li, W. Kalb, X.Z. Liu, H. Berger, R.W. Carpick, J. Hone, Frictional char-
[10] X. Wang, H. You, F. Liu, M. Li, L. Wan, S. Li, Q. Li, Y. Xu, R. Tian, Z. Yu, D. Xiang, acteristics of atomically thin sheets, Science (80-.) 328 (2010) 76–80, https://doi.
J. Cheng, Large-scale synthesis of few-layered graphene using CVD, Chem. Vap. org/10.1126/science.1184167.
Depos. 15 (2009) 53–56, https://doi.org/10.1002/cvde.200806737. [30] K.S. Kim, H.J. Lee, C. Lee, S.K. Lee, H. Jang, J.H. Ahn, J.H. Kim, H.J. Lee, Chemical
[11] S.J. Chae, F. Güneş, K.K. Kim, E.S. Kim, G.H. Han, S.M. Kim, H. Shin, S.M. Yoon, vapor deposition-grown graphene: the thinnest solid lubricant, ACS Nano 5 (2011)
J.Y. Choi, M.H. Park, C.W. Yang, D. Pribat, Y.H. Lee, Synthesis of large-area gra- 5107–5114, https://doi.org/10.1021/nn2011865.
phene layers on poly-nickel substrate by chemical vapor deposition: wrinkle for- [31] Z. Shi, P. Shum, A. Wasy, Z. Zhou, L.K.Y. Li, Tribological performance of few layer
mation, Adv. Mater. 21 (2009) 2328–2333, https://doi.org/10.1002/adma. graphene on textured M2 steel surfaces, Surf. Coatings Technol. 296 (2016)
200803016. 164–170, https://doi.org/10.1016/j.surfcoat.2016.04.031.
[12] Y. Zhang, L. Gomez, F.N. Ishikawa, A. Madaria, K. Ryu, C. Wang, A. Badmaev, [32] A.C. Ferrari, J.C. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri, F. Mauri, S. Piscanec,
C. Zhou, Comparison of graphene growth on single-crystalline and polycrystalline D. Jiang, K.S. Novoselov, S. Roth, A.K. Geim, Raman spectrum of graphene and
Ni by chemical vapor deposition, J. Phys. Chem. Lett. 1 (2010) 3101–3107, https:// graphene layers, Phys. Rev. Lett. 97 (2006) 1–4, https://doi.org/10.1103/
doi.org/10.1021/jz1011466. PhysRevLett. 97.187401.
[13] M. Losurdo, M.M. Giangregorio, P. Capezzuto, G. Bruno, Graphene CVD growth on [33] K.J. Peng, C.L. Wu, Y.H. Lin, Y.J. Liu, D.P. Tsai, Y.H. Pai, G.R. Lin, Hydrogen-free
copper and nickel: Role of hydrogen in kinetics and structure, Phys. Chem. Chem. PECVD growth of few-layer graphene on an ultra-thin nickel film at the threshold
Phys. 13 (2011) 20836–20843, https://doi.org/10.1039/c1cp22347j. dissolution temperature, J. Mater. Chem. C. 1 (2013) 3862–3870, https://doi.org/
[14] C.M. Seah, S.P. Chai, A.R. Mohamed, Mechanisms of graphene growth by chemical 10.1039/c3tc30332b.
vapour deposition on transition metals, Carbon N. Y. 70 (2014) 1–21, https://doi. [34] Y.-B. Guo, S.-W. Zhang, The tribological properties of multi-layered graphene as
org/10.1016/j.carbon.2013.12.073. additives of PAO2 oil in steel-steel contacts, Lubricants 4 (2016) 30, https://doi.
[15] R.S. Weatherup, B. Dlubak, S. Hofmann, Kinetic control of catalytic CVD for high- org/10.3390/lubricants4030030.

10
Vacuum 166 (2019) 307–315

Contents lists available at ScienceDirect

Vacuum
journal homepage: www.elsevier.com/locate/vacuum

Investigation on the lubrication potential of graphene oxide aqueous T


dispersion for self-mated stainless steel tribo-pair
Sudesh Singha,b, Xinchun Chenb, Chenhui Zhangb,∗, Rajnesh Tyagia,∗∗, Jianbin Luob
a
Department of Mechanical Engineering, Indian Institute of Technology (BHU), Varanasi, 221005, India
b
State Key Laboratory of Tribology, Tsinghua University, Beijing, 100084, China

A R T I C LE I N FO A B S T R A C T

Keywords: Due to high concerns for environmental protection and resource conservation, water-based lubrication with
Graphene some additives is gaining the attention of researchers. Hence, the present study has been conducted to explore
Tribology the tribological potential of lubricant containing graphene oxide (GO) as an additive in water. The friction and
Water lubrication wear performance was evaluated for self-mated SS 304 under reciprocating sliding using a ball-on-disc con-
Friction
figuration under high contact stress (∼1.42 GPa or higher). The results indicate that 0.1 wt% is the optimum
Wear
content of graphene oxide under the conditions used in the study and has the capability to reduce friction
coefficient from 0.56 (pure water) to 0.12 with an attendant reduction in wear volume by 68%. The enhance-
ment in tribological performance has been attributed to the formation of a low shear strength tribo-layer of
graphene on the surface of both the ball and the disc. The effect of normal load and sliding speed for optimized
concentration was also investigated. The average coefficient of friction was observed to increase with load
whereas, it was observed to decrease with increase in speed till 0.05 m/s beyond which it increased.

1. Introduction wear for self-mated 440C steel lubricated with solution-processed gra-
phene both in the air and dry nitrogen. Shen et al. [14] explored the
Water-based lubricants are gaining attention of researchers due to tribological behavior of in-situ exfoliated graphene aqueous dispersion
their various advantages such as low cost, non-flammable and en- and reported a reduction in friction coefficient and wear scar diameter
vironmental friendly nature. Nowadays, it is being used as a lubricant by 81.3% and 61.8%, respectively. Zhang et al. [15] reported that the
to replace the oil lubricants in numerous manufacturing, automotive, dispersion of graphene modified with oleic acid in PAO9 improved the
ship-building, food, and biomedical industries. But water lubrication tribological performance through the formation of a graphene protec-
has a drawback of limited load carrying capacity due to its low visc- tive layer on the surface of the specimen and the counterface as well.
osity. So, additives are required to enhance the lubricating capability of Yang et al. [16] dissolved the liquid-like graphene into deionized water
water [1,2]. Carbon-based nanomaterials as additives offer exceptional and observed a reduction in both the friction and the wear due to the
properties such as high chemical stability, high thermal conductivity, formation of tribo-film by electrostatic adsorption and rearrangement
low surface energy, high load-bearing capacity, weak inter-molecular, of graphene molecules. Guo et al. [17] investigated the tribological
and strong intra-molecular bonding, etc. Few studies have reported that performance of multilayer graphene sheets as an additive in PAO2 and
the addition of carbonaceous nanomaterials such as nano-diamonds reported a significant improvement. Graphene oxide has a number of
[3], fullerenes [4], graphite nano-sheets [5], CNTs [6], carbon nano- carbon-oxygen functional groups. These functional groups help in its
horns [7] and graphene [8], etc., as an additive in lubricants results in easy dispersion in water without any hydrophilic treatment, which is
improved tribological behavior of the base lubricant. normally required in case of graphene. Kinoshita et al. [18] suggested
Graphene, a two dimensional flaked carbon nanomaterial with sp2- that graphene oxide-water dispersion can achieve a very low friction
hybridization, offers better tribological properties both in atomic scale coefficient and high anti-wear performance after 60,000 cycles due to
and macro-scale [9–12]. Graphene has the capability to reduce friction the adsorption of graphene oxide sheets on the surfaces of WC ball as
and wear when used either as a solid lubricant or lubricant additive. well as steel plate to form a protective coating. Hae-Jin Kim and Dae-
Berman et al. [11,13] reported a significant reduction in friction and Eun Kim [2] observed a stable friction coefficient of 0.1 and a three


Corresponding author.
∗∗
Corresponding author.
E-mail addresses: chzhang@tsinghua.edu.cn (C. Zhang), rtyagi.mec@itbhu.ac.in (R. Tyagi).

https://doi.org/10.1016/j.vacuum.2019.05.015
Received 29 March 2019; Received in revised form 11 May 2019; Accepted 11 May 2019
Available online 14 May 2019
0042-207X/ © 2019 Published by Elsevier Ltd.
S. Singh, et al. Vacuum 166 (2019) 307–315

times lower wear rate in water lubrication when compared to oil lu- The morphologies of the worn surfaces of plate and ball were ex-
brication for rGO coated 440C stainless steel ball slid against 440C amined under optical microscopy (VHX-6000, Keyence, Japan), Raman
stainless steel. spectroscopy (LabRAM HR Evolution, Horiba Jobin Yvon, France) and
Till date, to the best of our knowledge, when compared to nano- high-resolution scanning electron microscopy (HR-SEM) (Nova Nano
scale and micro-scale, the macro-scale tribological properties of gra- SEM 450, FEI, USA) equipped with energy dispersive X-ray spectro-
phene oxide water dispersion under reciprocating motion have not been scopy (EDS). The wear volume of the ball having a wear scar of effective
much explored for considerably higher contact stresses. Hence, the diameter ‘D’ after tests was determined as mentioned in ASTM standard
present work is aimed at exploring the potential of graphene oxide G99-17 [19] using the following equation:
nano-sheets as an additive in water based lubrication. To fulfill the 2
πh 3D
objective, the dispersions of graphene oxide (GO) nano-sheets having V = ⎛ ⎞⎛ ⎜ + h2⎞ ⎟

different concentrations in pure water were prepared and the optimum ⎝ 6 ⎠⎝ 4 ⎠


concentration of graphene oxide was determined. Further, to explore Where h is the height of material removed from the ball and can be
the tribological behavior for the optimized concentration of graphene calculated as follows:
oxide in water, the effect of load and sliding speed was studied in detail
by carrying out the wear tests using a ball-on-flat configuration under D2
h = R− R2 −
reciprocating conditions. 4

Where R is the original ball radius used for the friction test.
2. Experimental procedure

Commercially available graphene oxide nano-sheets of thickness 3. Results and discussion


0.8–1.2 nm and diameter 0.5–5 μm were used as a graphene source
procured from XFNANO Inc., China. Graphene oxide was initially 3.1. Characterizations of graphene oxide
characterized by Raman spectrometer (LabRAM HR Evolution, Horiba
Jobin Yvon, France). The dispersions were prepared by sonication of It is well known that the nano-crystalline materials are best char-
graphene oxide in pure water with different weight concentrations for acterized by Raman spectroscopy. Fig. 2a presents the Raman spectrum
about 2 h. Fig. 1a presents the digital images of graphene oxide-water of graphene oxide powder, which shows the presence of a D peak at
dispersion after sonication. To characterize the thickness and average 1349 cm−1, a G peak at 1589 cm−1 and a weak intensity 2D peak at
diameter of graphene oxide in dispersion, an AFM (Dimension Icon, 2715 cm−1. Another weak intensity peak D + D’ could also be observed
Bruker, USA) and a transmission electron microscope (TEM, JEOL- at 2932 cm−1.
2010F, Japan) were utilized. Rheological behavior of dispersions was Atomic force microscope and transmission electron microscopy
measured at 25 °C using a rheometer (MCR 302, Anton Paar, Austria) (TEM) were used to characterize the dimensions of graphene oxide
having cone and plate geometry with a 50 mm diameter plate and 1° nano-sheets dispersed in water. Fig. 2b shows a 2D AFM image and
cone angle over a range of shear rates from 1 to 1000 s−1. corresponding height profile (at a particular section lined in the 2D
Tribological tests were performed at different concentrations (0.01, image) of some dispersed nano-sheets. One could observe that the
0.05, 0.1, and 0.5 wt%) of GO in water, different normal loads (5, 10, thickness of sheets is around 1.2 nm. However, a higher thickness of
15, and 20 N) and different speeds (0.005, 0.01, 0.05, and 0.1 m/s) as about 2.4 nm, observed at the center in AFM image may be due to the
mentioned in Table 1 using a multi-functional tribometer (Rtec in- overlapping of two layers which is clearly visible in the image. The
struments, USA) at room temperature with a ball-on-plate contact rheological properties play a significant role in lubrication. Therefore,
geometry under reciprocating motion. The stroke length was kept as the rheological behavior of the dispersions, having different con-
5 mm and the test duration was 3600 cycles. The tests were also run centrations of graphene oxide was determined by using a rheometer
with pure water as a lubricant for the purpose of comparison. SUS 304 with cone-plate geometry. Fig. 2c illustrates the variation of dynamic
stainless steel ball of 4 mm diameter and the counterface plate also viscosity with shear rate and one may infer that the viscosity of water
made of SUS 304 was selected as tribo-pair and plates were polished to increases with increasing amount of addition of graphene oxide. Fig. 2d
obtain an average roughness (Rq) of 10 nm. Both the plate and the ball presents a TEM image of graphene oxide sheets confirming that the size
were sonicated in acetone and isopropanol to remove the contaminants is lower than 5 μm. Dark areas indicate the thick stacking of several
from their surface. Fig. 1b presents the schematic diagram of ball-on- graphene oxide nano-sheets whereas the transparent area indicates
plate contact assembly used for the experiments. A dispersion volume of much thinner films of few-layer graphene oxide.
100 μl was applied on the highly polished surface before starting the
test, and no additional dispersion was supplied during the friction test. 3.2. Effect of concentration on the friction coefficient
Each experimental condition was tested thrice in order to minimize the
experimental error, and the average value is reported. Friction and wear tests under reciprocating motion were conducted

Fig. 1. (a) Digital images of graphene oxide-water dispersion after 2 h sonication with different graphene oxide concentrations and (b) Schematic of a ball-on-plate
contact assembly for friction and wear tests.

308
S. Singh, et al. Vacuum 166 (2019) 307–315

Table 1
Experimental conditions.
Effects Load (N) Max. Contact pressure (GPa) Lubricants Sliding speed (m/s) Number of cycles

Concentration variation 5 1.42 water 0.01 3600


0.01 wt% GO in water
0.05 wt% GO in water
0.1 wt% GO in water
0.5 wt% GO in water
Load variation 5 1.42 0.1 wt% GO in water 0.01 3600
10 1.79
15 2.05
20 2.26
Speed variation 5 1.42 0.1 wt% GO in water 0.005 3600
0.01
0.05
0.1

to investigate the effect of a change in concentration of graphene oxide was able to significantly reduce the coefficient of friction (COF) of the
in pure water on the macro-scale tribological performance of self-mated system as evident from Fig. 3a. This reduction in COF may be attributed
stainless steel tribo-pair under a load of 5 N and sliding speed of to the low interlayer shear stress between graphene oxide nano-sheets.
0.01 m/s in ambient condition. The pure water as a lubricant was The variation of average steady-state COF of the system with varying
chosen as a reference for comparing the lubrication capabilities of concentration of graphene oxide in water is presented in Fig. 3b. It is
different concentrations of graphene oxide in water. Fig. 3a shows the found that average coefficient of friction decreases with an increasing
variation of friction coefficient with time for different concentrations of amount of graphene oxide from 0.01 to 0.1 wt% and beyond that in-
graphene oxide in water and for pure water under a load of 5 N. Every creases slightly till 0.5 wt% indicating that 0.1 wt% is the optimum
experimental condition was tested at least 3 times to ensure the re- content of graphene oxide under the conditions used in the present
peatability of friction tests. It is clear from Fig. 3 that the pure water has investigation. Further, one could see that the optimal concentration of
large fluctuation and a high friction coefficient throughout the friction graphene oxide (i.e., 0.1 wt%) exhibited a comparatively consistent,
test. This is mainly because of low viscosity and low pressure-viscosity and low friction coefficient throughout the test. The lower and higher
coefficient of water. The friction coefficient of the system using pure concentration of graphene oxide have shown relatively higher friction
water was approximately 0.56. However, one could observe that even a coefficients in comparison to 0.1 wt% as evident from Fig. 3b, which
small concentration (0.01 wt%) of graphene oxide nano-sheets in water might have been caused by the deterioration in lubricating performance

Fig. 2. (a) Raman spectrum of graphene oxide powder, (b) 2D AFM image and height profile across the line, (c) Variation of dynamic viscosity as a function of shear
rate for different concentrations of graphene oxide in water, (d) TEM image.

309
S. Singh, et al. Vacuum 166 (2019) 307–315

Fig. 3. (a)Time-resolved variation of friction coefficient for different concentrations of graphene oxide in water and for pure water under the load of 5 N and (b) The
average friction coefficients as a function of graphene oxide concentration in water.

either due to inadequate or excess amounts of graphene oxide in water and wear track width is about 34% and 36%, respectively, when
depending on the amount of addition of graphene oxide. When com- compared with pure water. During the sliding process, the graphene
pared with pure water lubrication, the lubrication with 0.1 wt% gra- oxide nano-sheets get adsorbed on the friction surfaces and form a
phene oxide in water reduced the friction coefficient of the system by protective tribo-layer. This tribo-layer hinders metal-metal contact and
78% (i.e., from 0.56 to 0.12). provides easy shearing interface reducing thereby the friction coeffi-
cient.
Fig. 5 gives the wear volume of steel ball after the friction tests for
3.3. Worn surface analysis different concentrations of graphene oxide in water. One may observe
that wear volume gets significantly reduced by the addition of just
The present section discusses the effect of graphene oxide con- 0.01 wt% graphene oxide in water as evident from Fig. 5. Among the
centration on the wear of tribo-pair i.e. ball and plate. Wear scar dia- lubricants containing graphene oxide, the worn volume decreases as the
meter and wear track width are considered the key parameters for amount of graphene oxide increases from 0.01 to 0.1 wt%, showing a
evaluating the anti-wear performance of the lubricants. Fig. 4 shows the minimum at 0.1 wt% and increases thereafter till 0.5 wt%, the largest
optical micrographs of wear scar on the ball and corresponding wear amount of additive used in present work. The wear volume is found to
track on plates, respectively, after the friction test under a load of 5 N decrease by 68% for 0.1 wt % graphene oxide concentration, in com-
and sliding speed 0.01 m/s for 3600 cycles. Under the lubrication parison to that of pure water. The lubricant containing 0.1 wt% GO in
condition of pure water, the severe wear of steel ball, as well as plate, water has shown the lowest wear volume and the friction coefficient
occurred. Figs.4 (a and b). Show the wear scar and wear tracks on the under the conditions used in the present study as seen from Figs. 3 and
counter-face plate with pure water as a lubricant. It can be seen that 5. Hence, it may be concluded that 0.1 wt% GO addition is the optimum
wear scar diameter and wear track width for pure water lubrication are amount. Therefore, further testing at different loads and speeds was
about 266.8 μm and 281.7 μm, respectively. Further, one could observe conducted for this lubricant.
a significant reduction in both the wear scar diameter on the ball and It can be revealed from the optical images (Fig. 4) that there oc-
the wear track width on the corresponding plate with the addition of curred a chemical reaction between the steel and carbon (graphene) for
0.01 wt% GO as evident from a comparison of Figs. 4(a and c and b & d) each concentration of graphene oxide in water. The severity of the
which reflects the effect of GO nano-sheets addition in improving the chemical reaction is observed to increase with increasing concentration
performance of the lubricant. A comparison of wear scars shown in of graphene oxide. For 0.5 wt%, there might be a strong chemical re-
Figs. 4 (c, e, g and i) indicates that wear scar diameter reduces with an action accelerated by a large number of carbon atoms between the
increasing amount of addition of GO nano-sheets till 0.1 wt% and in- carbon (graphene) and steel. The degradation in anti-wear performance
creases thereafter. A similar trend is observed for wear track widths for 0.5 wt% graphene oxide in water may be ascribed to the agglom-
shown in Figs.4 (d, f, h, and j). The addition of 0.1 wt% of GO in water eration of excessive graphene nano-sheets resulting in the formation of
has shown the smallest diameter and wear track width in the present a lump and making it difficult for graphene to enter between the
study as evident from Fig. 4 and the reduction in wear scar diameter

Fig. 4. Optical micrographs of wear scars and wear tracks on the ball and plates, respectively, after tests (a & b) pure water, (c & d) 0.01 GO-water, (e & f) 0.05 GO-
water (g & h) 0.1 GO-water and (i & j) 0.5 GO-water.

310
S. Singh, et al. Vacuum 166 (2019) 307–315

surfaces with slightly increased defects concentration.


The formation of tribo-layer in the contact area was also confirmed
by examining the worn surface under TEM. The cross-sectional samples
for TEM were prepared by FIB in-situ lift-out technique from the worn
friction surfaces of steel surfaces. Additional metallic layers of Cr and Pt
were sequentially deposited on the surface of both the ball and the plate
before FIB process to prevent damage, spurious sputtering of the top
portion of the specimen and to delineate the location of the area of
interest as well. Fig. 8 (b and c) show the cross-sectional TEM image of
tribo-film with a thickness of 25 ± 5 nm on the steel plate. Figs. 8 (d, e,
and f) present the cross-sectional TEM images of tribo-film on the steel
ball with a film thickness of 50 ± 10 nm (approximately). The contact
area contains the graphene stacked layers, fragments of graphene oxide
sheets and a matrix of amorphous carbon as evident from TEM images
shown in Fig. 8. The d-spacing for graphene layers is found to be about
0.42 nm, which is larger than that of pristine graphite which has d
(002) value of 0.33 nm. This increase in the interlayer spacing may
probably be due to the introduction of oxygen functional groups.

3.4. Effect of load on tribo-performance


Fig. 5. Wear volume of stainless steel balls as a function of the concentration of
graphene oxide in water. In the study presented above, the friction tests were performed at a
constant load of 5 N with varying concentration of graphene oxide in
friction pairs. Moreover, excessive concentration has also been reported pure water to determine the optimum concentration of graphene oxide.
to reduce the dispersion behavior of graphene as an additive [17]. It has Based on the above investigation, a concentration of 0.1 wt% graphene
also been indicated that the agglomerations due to a higher con- oxide in water has been found to be optimum as it showed the least
centration of graphene/graphene-based nano-particle can also cause friction and wear under the tested condition. Hence, further tests were
the abrasive wear (even though these particles are soft) of the counter conducted at loads different loads of 10, 15 and 20 N to explore the
surface, leading to an increase in wear of surfaces [14,21]. friction and wear behavior of tribo-pair for the optimized concentration
Fig. 6 shows the typical HRSEM micrographs of worn tracks of plate (i.e., 0.1 wt%) at a speed of 0.01 m/s for the duration of 3600 s corre-
lubricated with pure water and with graphene oxide (0.1 wt% and sponding to a distance of 36 m. The variation of coefficient of friction
0.5 wt%) in water after the friction tests and corresponding energy with time is illustrated in Fig. 9a. Almost the same frictional behavior
dispersive spectroscopy (EDS) of the whole micrograph. It can be no- could be observed for normal loads of 5, 10, and 15 N during the initial
ticed that surface under pure water lubrication (Fig. 6a) has consider- 700 s. After that, the friction coefficient for 10 and 15 N was found to
able high wear (mainly abrasive) and higher surface roughness when increase gradually while the friction coefficient for 5 N load was com-
compared to graphene oxide in water lubrication (Figs. 6 b & c). Some paratively low and steady. For a normal load of 20 N, the friction
darker regions can be seen clearly in HRSEM image (Fig. 6b and c), coefficient was higher in the beginning and this behavior continued to
which are due to the adsorption of graphene oxide nano-sheets on the the completion of the test with a gradual increase in friction coefficient
flat plate surface. The presence of an intense carbon peak in EDS cor- with time. However, the amplitude of fluctuations is also observed to
responding to 0.1 and 0.5 wt% graphene oxide containing lubricant increase with increasing load as evident from Fig. 9a.
(Fig. 6 b and c) reflects the increased amount of carbon, which basically For a better understanding of the load-friction coefficient relation,
came from the lubricant through the adsorption of graphene oxide the variation of the average friction coefficient with the normal load is
nano-sheets on the surface during the sliding process. These adsorbed presented in Fig. 9b. The average friction coefficient is observed to
nano-sheets formed a protective layer on the surface, which further increase with increasing load. From the above results, it is quite clear
smoothened the surface and resulted in improved tribological perfor- that graphene oxide in water has a stable and low friction coefficient
mance by inhibiting the metal-metal contact. The elemental of the due to the presence of a tribo-layer between sliding pair which protect
presented HRSEM images is also shown in Fig. 6, which clearly shows the underlying materials from being penetrated by the asperities of due
the increased carbon content with an increase in the concentration of to lower loads. Since the protection provided by the tribo-layer is
graphene oxide. limited by its load carrying capability, the damage to tribo-layers occurs
Further, Raman spectroscopy analysis was performed to confirm the with an increasing load where asperities are able to penetrate through
formation of graphene tribo-layer on both the contact surfaces and the tribo-layer giving rise to metal-contact and resulting thereby, in
compared with the Raman spectrum of pristine graphene oxide used in increased friction and wear.
the present study. Fig. 7 presents the Raman spectrum of graphene Figs. 10a and 10b shows the optical micrographs of wear scar on the
oxide on the worn surface of the ball as well as on plate lubricated with ball and the corresponding wear track width on the plate after the
0.1 wt% graphene oxide in water and of pristine graphene oxide. The friction test under a load of 20 N and sliding speed 0.01 m/s for 3600
pattern of peaks for pristine graphene oxide, the graphene oxide ad- cycles. It can be observed that wear scar diameter and wear track width
sorbed on the wear track and wear scar is found to be similar with only for 20 N load are about 313 μm and 340 μm, respectively, which are
change in ID/IG ratio which is 0.93 for pristine graphene oxide and 1.11 larger than those observed for 5 N load (Figs. 4g and 4h). Fig. 10c
and 1.13, respectively, for graphene oxide on ball and plate worn sur- presents the relationship between the wear volume and normal load
faces. An increase in D peak intensity in Raman spectrum of the worn and it is found that wear volume increases linearly with the load. As
surface of the ball, as well as plate, could also be seen which points discussed above, with the increase in normal load, the loss of protective
toward an increase in the defect and disordering of graphene oxide tribo-layer also increases. So, there occurs a reduction in anti-wear
during the sliding process due to the presence of high contact pressure behavior of tribo-layer, which causes an increase in wear volume with
on graphene oxide nano-sheets [20]. Thus, Raman Spectrum of ball and the load. As shown in Fig. 10c, the wear volume of the ball under 5 N
plate confirms the presence of the graphene layer on the rubbing load is about 83.4% less compared to the that under 20 N. So, it is clear
that lubrication with graphene oxide in water is more effective and

311
S. Singh, et al. Vacuum 166 (2019) 307–315

Fig. 6. HRSEM micrographs and energy dispersive spectroscopy (EDS) of wear track lubricated with (a) pure water, (b) 0.1 wt% of graphene oxide in water, and (c)
0.5 wt% of graphene oxide in water.

beneficial under relatively lower loads.

3.5. Effect of speed on tribo-performance

In order to compare the variation in the friction and wear behavior


with sliding speed, additional friction tests were conducted with 0.1%
graphene oxide-water dispersion at a normal load of 5 N, a duration of
3600s corresponding to a sliding distance of 36 m and at different
sliding speeds of 0.005, 0.01, 0.05, and 0.1 m/s, corresponding to 0.5,
1, 5, and 10 Hz reciprocating frequency. Fig. 11a shows the variation of
friction coefficient with sliding speed at the maximum Hertzian contact
pressure of 1.42 GPa. A decrease in the value of friction coefficient
could be seen with an increase in sliding speed from 0.005 to 0.05 m/s
and the minimum friction coefficient occurred at a speed of 0.05 m/s.
Fig. 11b presents the variation of average friction coefficient against
sliding speeds. It is evident that the average friction coefficient de-
creases with increase in speed up to 0.05 m/s and after that, it increases
for 0.1 m/s. The tendency of entrainment of graphene oxide between
Fig. 7. Raman spectra of the pristine graphene oxide and worn surface of ball the contact pairs increases with increasing sliding speed due to better
and plate lubricated with 0.1 wt% graphene oxide in water. circulation of graphene oxide sheets in water between the rubbing
surfaces. It may also due to reduced time for asperity contact and so, the
less deformation of asperities occurs. Hence, with an increase in the

312
S. Singh, et al. Vacuum 166 (2019) 307–315

Fig. 8. (a) Schematic of the contact area of friction pair, (b, c) cross-sectional TEM image of the tribo-film on the plate, (d, e, f) cross-sectional TEM image of the
tribo-film on the ball.

speed, the real contact area decrease which causes a decline in friction surfaces which increases the chance of forming the protective layer on
coefficient and wear as well. However, for the highest sliding speed of contact pairs as elaborated earlier. However, an increase in wear vo-
0.1 m/s used in the present study, it was observed that lubricant moved lume beyond 0.05 m/s may be due to the unavailability of the lubricant
out of the contact surface during the experiments because of high speed between the contacting surfaces as discussed above.
resulting in discontinuous lubrication due to inadequate supply of lu-
bricant between the rubbing surfaces. This might have given rise to 3.6. Anti-friction and anti-wear mechanism
metal-metal contact and resulted in an increase in both the friction
coefficient and wear at the highest speed. Based on the above results, the anti-wear and anti-friction me-
Figs. 12a and 12b presents the optical micrographs showing wear chanism of graphene oxide as lubricant additives can be proposed.
scar diameter on ball and corresponding wear track width on plates Under pure water lubrication, the ball slides over the plate causing
after the testing under a load of 5 N and sliding speed of 0.05 m/s for abrasive wear mainly. However, the addition of graphene oxide in
3600 cycles. It can be observed that wear scar diameter and wear track water brings about a sudden decrease in both friction and wear due to
width for 0.05 m/s load are about ∼168 μm and ∼173 μm, respec- the adsorption of GO on contacting surface and subsequent formation of
tively, which are slightly smaller than those found for 0.01 m/s a protective tribo-film as evident from Fig. 8. This film plays a crucial
(Figs. 4g and 4h). Fig. 12c presents the relationship between the wear role in reducing both the friction and the wear by protecting the un-
volume and sliding speed and it is observed that the wear volume de- derlying metals from direct metal-metal contact and providing an easy
creases with increasing sliding velocity from 0.005 m/s to 0.05 m/s and to shear interface between the contacting pair. The observed friction
increases thereafter till 0.1 m/s. The minimum wear volume loss occurs and wear behavior may be explained on the basis of the amount of
for a sliding speed of 0.05 m/s. The decrease in wear volume with in- graphene oxide nano-sheets in the lubricant. At relatively low con-
creasing sliding velocity till 0.05 m/s may be explained on the basis of centrations (< 0.1 wt%) adequate nano-sheets are not available to form
better circulation of graphene oxide nano-sheets between the rubbing a continuous tribo-film whereas larger concentration (> 0.1 wt%) leads

Fig. 9. (a)Time-resolved variation of friction coefficient for different normal load lubricated with 0.1 wt% graphene oxide in water and (b) The average friction
coefficients as a function of the normal load.

313
S. Singh, et al. Vacuum 166 (2019) 307–315

Fig. 10. (a) Optical micrographs of wear scars on the ball, (b) corresponding wear tracks on the plate for 20 N load and (c) wear volume of steel balls as a function of
normal load for 0.1 wt% graphene oxide in water.

to a strong chemical interaction between graphene and steel, and thus surfaces and subsequent formation of a protective tribo-film which
increasing the friction and wear. Another reason for increased friction prevented the surfaces to come in direct contact resulting in im-
and wear for 0.1 wt% GO containing water may be the aggregation and proved tribological performance.
clustering of nano-sheets that take place at a concentration higher than 2) The results indicate that 0.1 wt% concentration of GO in water is the
0.1 wt%, which hinders the entrapment of graphene nano-sheets be- optimum, as it resulted in the least friction coefficient of 0.12 as
tween the contacting surfaces as indicated by earlier studies [8,21]. against 0.56 for pure water under the conditions of load and speed
Both the conditions lead to an increase in friction as indicated by Fig. 4. used in the present study. Also, it showed the least wear scar dia-
An increase in friction with the increasing load may also be explained meter as well as wear track width compared to those observed for
by the increasing penetration of asperities through the tribo-layer at pure water lubrication.
relatively higher loads due to increased stress which results in the loss 3) Addition of graphene oxide in water resulted in improved tribolo-
of effectiveness of this tribo-layer in reducing the friction. For sliding gical performance in terms of both the reduced friction coefficient
speed, up to 0.05 m/s, the tribological properties improved due to en- and wear volume at relatively lower loads compared to that at
trainment of more sheets between the rubbing surfaces with increased higher loads due to the loss of effectiveness of tribo-layer in pre-
velocity. However, for the speed beyond 0.05 m/s, lubricant squeezed venting asperity contacts under high contact pressures.
out of the contact surface due to high speed resulting in discontinuous 4) For 0.1 wt% graphene oxide in water, the sliding speed was opti-
lubrication and inadequate supply of lubricant between the rubbing mized to 0.05 m/s which had the least friction coefficient and wear
surfaces. Since, the lubrication can only be provided by the entrapped volume due to better circulation of dispersed nano-sheets in water
graphene nano-sheets between the surfaces, the discontinuous supply of and thus enhancing the formation of tribo-layer.
lubricant might have led to an increase in friction and wear at the
highest speed of 0.1 m/s used in the present investigation. The results demonstrated that graphene oxide as an additive in
water holds promise in improving the tribological performance of water
4. Conclusions as an environmental friendly lubricant and the potential needs to be
tapped for a cleaner and greener environment.
A series of experiments were conducted to investigate the friction
and wear performance of SS 304 steel under the water lubrication
containing well-dispersed graphene oxide nano-sheets as an additive Acknowledgment
and the following conclusions were drawn:
The authors are thankful to the National Natural Science
1) A significant reduction in COF was observed even for the addition of Foundation of China (Grant No. 51605247) for financial support.
a small concentration of graphene oxide in water. This has been
attributed to the adsorption of graphene nano-sheets on the sliding

Fig. 11. (a) Friction coefficient as a function of cycles for different sliding speeds for 0.1 wt% graphene oxide in water, (b) The variation of average friction
coefficient against sliding speed.

314
S. Singh, et al. Vacuum 166 (2019) 307–315

Fig. 12. (a) Optical micrographs of wear scar, (b) corresponding wear tracks for sliding speed of 0.05 m/s, and (c) wear volume of steel balls as a function of sliding
speed for 0.1 wt% graphene oxide in water.

References 10.1016/j.carbon.2012.11.061.
[12] D. Berman, A. Erdemir, A.V. Zinovev, A.V. Sumant, Nanoscale friction properties of
graphene and graphene oxide, Diam. Relat. Mater. 54 (2015) 91–96, https://doi.
[1] A. Tomala, A. Karpinska, W.S.M. Werner, A. Olver, H. Störi, Tribological properties org/10.1016/j.diamond.2014.10.012.
of additives for water-based lubricants, Wear 269 (2010) 804–810, https://doi.org/ [13] D. Berman, A. Erdemir, A.V. Sumant, Reduced wear and friction enabled by gra-
10.1016/j.wear.2010.08.008. phene layers on sliding steel surfaces in dry nitrogen, Carbon N Y 59 (2013)
[2] H.-J. Kim, D.-E. Kim, Water lubrication of stainless steel using reduced graphene 167–175, https://doi.org/10.1016/j.carbon.2013.03.006.
oxide coating, Sci. Rep. 5 (2015) 17034, https://doi.org/10.1038/srep17034. [14] S. Liang, Z. Shen, M. Yi, L. Liu, X. Zhang, S. Ma, In-situ exfoliated graphene for high-
[3] X. Tao, Z. Jiazheng, X. Kang, The ball-bearing effect of diamond nanoparticles as an performance water-based lubricants, Carbon N Y 96 (2016) 1181–1190, https://
oil additive, J. Phys. D Appl. Phys. 29 (1996) 2932–2937, https://doi.org/10.1088/ doi.org/10.1016/j.carbon.2015.10.077.
0022-3727/29/11/029. [15] W. Zhang, M. Zhou, H. Zhu, Y. Tian, K. Wang, J. Wei, et al., Tribological properties
[4] J. Lee, S. Cho, Y. Hwang, C. Lee, S.H. Kim, Enhancement of lubrication properties of of oleic acid-modified graphene as lubricant oil additives, J. Phys. D Appl. Phys. 44
nano-oil by controlling the amount of fullerene nanoparticle additives, Tribol. Lett. (2011), https://doi.org/10.1088/0022-3727/44/20/205303.
28 (2007) 203–208, https://doi.org/10.1007/s11249-007-9265-2. [16] J. Yang, Y. Xia, H. Song, B. Chen, Z. Zhang, Synthesis of the liquid-like graphene
[5] C.G. Lee, Y.J. Hwang, Y.M. Choi, J.K. Lee, C. Choi, J.M. Oh, A study on the tribo- with excellent tribological properties, Tribol. Int. 105 (2017) 118–124, https://doi.
logical characteristics of graphite nano lubricants, Int. J. Precis. Eng. Manuf. 10 org/10.1016/j.triboint.2016.09.040.
(2009) 85–90, https://doi.org/10.1007/s12541-009-0013-4. [17] Y.-B. Guo, S.-W. Zhang, The tribological properties of multi-layered graphene as
[6] C. Espejo, F.J. Carrion, D. Martinez, M.D. Bermudez, Multi-walled carbon nanotube- additives of PAO2 oil in steel–steel contacts, Lubricants 4 (2016) 30, https://doi.
imidazolium tosylate ionic liquid lubricant, Tribol. Lett. 50 (2013) 127–136, org/10.3390/lubricants4030030.
https://doi.org/10.1007/s11249-012-0082-x. [18] H. Kinoshita, Y. Nishina, A.A. Alias, M. Fujii, Tribological properties of monolayer
[7] V. Zin, F. Agresti, S. Barison, L. Colla, M. Fabrizio, Influence of Cu, TiO2 nano- graphene oxide sheets as water-based lubricant additives, Carbon N Y 66 (2014)
particles and carbon nano-horns on tribological properties of engine oil, J. Nanosci. 720–723, https://doi.org/10.1016/j.carbon.2013.08.045.
Nanotechnol. 15 (2015) 3590–3598, https://doi.org/10.1166/jnn.2015.9839. [19] ASTM G99-17 Standard Test Method for Wear Testing with a Pin-On-Disk
[8] S.S.N. Azman, N.W.M. Zulkifli, H. Masjuki, M. Gulzar, R. Zahid, Study of tribolo- Apparatus, ASTM International, West Conshohocken, PA, 2017https://doi.org/10.
gical properties of lubricating oil blend added with graphene nanoplatelets, J. 1520/G0099-17.
Mater. Res. 31 (2016) 1932–1938, https://doi.org/10.1557/jmr.2016.24. [20] D. Zheng, Z bing Cai, M xue Shen, Z yang Li, M hao Zhu, Investigation of the tri-
[9] P.F. Li, H. Zhou, X.H. Cheng, Nano/micro tribological behaviors of a self-assembled bology behaviour of the graphene nanosheets as oil additives on textured alloy cast
graphene oxide nanolayer on Ti/titanium alloy substrates, Appl. Surf. Sci. 285 iron surface, Appl. Surf. Sci. 387 (2016) 66–75, https://doi.org/10.1016/j.apsusc.
(2013) 937–944, https://doi.org/10.1016/j.apsusc.2013.09.019. 2016.06.080.
[10] G. Paolicelli, M. Tripathi, V. Corradini, A. Candini, S. Valeri, Nanoscale frictional [21] O. Elomaa, V.K. Singh, A. Iyer, T.J. Hakala, J. Koskinen, Graphene oxide in water
behavior of graphene on SiO2and Ni(111) substrates, Nanotechnology 26 (2015), lubrication on diamond-like carbon vs. stainless steel high-load contacts, Diam.
https://doi.org/10.1088/0957-4484/26/5/055703. Relat. Mater. 52 (2015) 43–48, https://doi.org/10.1016/j.diamond.2014.12.003.
[11] D. Berman, A. Erdemir, A.V. Sumant, Few layer graphene to reduce wear and
friction on sliding steel surfaces, Carbon N Y 54 (2013) 454–459, https://doi.org/

315

You might also like