Você está na página 1de 11

Carbon 123 (2017) 7e17

Contents lists available at ScienceDirect

Carbon
journal homepage: www.elsevier.com/locate/carbon

Novel tertiary dry solid lubricant on steel surfaces reduces significant


friction and wear under high load conditions
Abdullah A. Alazemi a, 1, Arthur D. Dysart b, 1, Steven J. Shaffer c, Vilas G. Pol b, *,
Lars-Erik Stacke d, Farshid Sadeghi a, **
a
School of Mechanical Engineering, Purdue University, West Lafayette, Indiana 47907, United States
b
Davidson School of Chemical Engineering, Purdue University, West Lafayette, Indiana 47907, United States
c
Bruker Nano Surfaces Division, 61 Daggett Dr, San Jose, CA 95134-2109, United States
d €teborg SE-415 50, Sweden
SKF Digital Business Technology, Go

a r t i c l e i n f o a b s t r a c t

Article history: A novel graphene-zinc oxide composite film is created and studied as a solid-state lubricant for friction
Received 19 April 2017 and wear reduction under extreme load conditions. The liquid-free composite is made from a slurry of
Received in revised form graphene, zinc oxide, and polyvinylidene difluoride spin-coated onto a stainless steel substrate.
9 June 2017
Enhanced tribological performance was measured under ambient conditions using a ball-on-disk trib-
Accepted 10 July 2017
ometer with contact pressures up to 1.02 GPa and sliding distances up to 450 m. The graphene-rich
Available online 11 July 2017
lubricant demonstrates substantial friction and wear reduction (ca. 90%) compared to unlubricated
sliding. The composite film is able to maintain its lubricating effects under extreme operating conditions
including 15 N normal load and 450 m sliding distance. Following tribological testing, optical and
spectroscopic analysis of the formed wear scars reveal a persistent protective film on the ball and disk
surfaces. The excellent tribological performance of this graphene-rich composite is attributed to the
adhesion effect from zinc oxide: zinc adheres graphene to the contact interface, maintaining improved
tribological performance under high contact pressure. The durability and resilience of this adhesive
coating suggest exceptional potential as a dry lubricant for high load-bearing applications.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction and fire. In this context, dry solid-state lubricants are a viable
alternative to their liquid counterparts in extreme operating
Mechanical systems with greater energy efficiency and lower environments.
environment impact require enhanced performance at moving Graphene, due to its distinct material properties, is a promising
interfaces. The fundamental causes of mechanical failure are fric- candidate for solid-state lubrication [9,10]. Graphene is a unique
tion and wear: reducing these energy losses improves performance carbon allotrope, consisting of a flat lattice of aromatic carbon rings
and lifetime of many mechanical systems. Conventional methods to only one atom thick. This two-dimensional arrangement enables
improve energy efficiency utilize liquid lubricants, including superior thermal conductivity [11], extreme mechanical strength
organic oils [1], to reduce friction between contacting surfaces in [12], and ultralow friction [13e15]. In particular, the low coefficient
relative motion. Furthermore, the addition of nanoparticles to of friction for graphene materials has been demonstrated at the
lubricating oils has been shown to further enhance tribological nanoscale by atomic force microscopy [16e18]. These uncommon
performance [2e7]. However, despite their convenience and utility, properties, combined with continuous production processing,
liquid lubricants cannot be used in situations of high temperature distinguish graphene from other nanomaterials for friction and
or low pressure (viz., vacuum) [8], due to the risk of volatilization wear reduction [19]. However, existing tribological studies of gra-
phene at the microscale [15,20] and macroscale [21,22] show that
friction and wear reductions only occur under low contact pres-
* Corresponding author.
sures: friction and wear rapidly increase under high contact pres-
** Corresponding author. sure (i.e., >0.5 GPa [9,22]). The failure of bare graphene is attributed
E-mail addresses: vpol@purdue.edu (V.G. Pol), sadeghi@purdue.edu (F. Sadeghi). to poor adhesion with substrate surface, enabling graphene ejection
1
These authors contributed equally.

http://dx.doi.org/10.1016/j.carbon.2017.07.030
0008-6223/© 2017 Elsevier Ltd. All rights reserved.
8 A.A. Alazemi et al. / Carbon 123 (2017) 7e17

under excessive pressure. Therefore, durable adhesion between the Preparation of composite coating: The solid-phase composite
lubricant film and contact surfaces is critical for application of gra- lubricant slurry was prepared by ultrasonic homogenization. A
phene as a solid lubricant [23]. viscous mixture of 85.5 %-wt. graphene (United Nanotech Innovations
To overcome poor adhesion between graphene and contact PVT Ltd.), 9.5 %-wt. zinc oxide, and 5 %-wt. polyvinylidene difluoride
surfaces, zinc oxide can be added to the dry lubricant. Zinc oxide (PVDF, Sigma-Aldrich) was prepared with solvent N-Methyl-2-pyr-
(ZnO) is a common additive for paints and coatings to improve both rolidone (NMP, Sigma-Aldrich). Homogenization was performed in a
metallic adhesion and corrosion resistance [24,25]. Wu et al. [26] sealed borosilicate scintillation vial (Thermo-Fisher Scientific Co.)
demonstrated that the ZnO film improves the interface contact dispersed using an ultrasonic bath (RPI Corp).
and adhesion of graphene sheets with the surface of silver Lamination of the ultrasonically-mixed dispersion onto the
electrode-coated glass substrate. Furthermore, zinc compounds are contact surface was performed using a spin-coating technique. The
widely used in oil lubricants as anti-wear additives [27e29]. Zinc homogenized mixture was transferred to the center of the stainless
dialkyldithiophosphate (ZDDP) in particular is a common additive steel disk substrate (Bruker Scientific Co.). Immediately, the loaded
for automotive oil lubricants to prevent wear of machine compo- disk was accelerated to a constant rotational speed 1000 rev min 1
nents [28,30]. Under an applied load, ZDDP forms a protective tri- using a tribometer (UMT-3, Bruker Corp.). The disk was rotated at
bofilm that prevents direct contact between sliding surfaces this speed for ca. 2 min. After deceleration to rest, the disk surface
[31e33]. Ex situ X-ray characterization has identified zinc-oxygen was uniformly covered by a black thin film (ca. 10 mm thick). The
bonds as a primary component of formed tribofilms [30]. coated disk was then dried at temperature 80  C for at least 12 h to
In this work, a graphene-rich composite is studied as a solid remove the NMP solvent.
lubricant to minimize friction and wear losses under high contact To understand the tribological role of zinc oxide, two reference
pressures and long sliding distances. The proposed composite con- composite coatings were prepared: (I) graphene and PVDF, and (II)
sists of graphene, zinc oxide, and polyvinylidene difluoride binder. graphene only. The coatings were prepared using modified versions
Spin coating applies the graphene-based composite as a ca. 10 mm of the above procedure. To produce the graphene and PVDF coating,
thick film onto a stainless steel surface. Tribological performance was a viscous mixture of 95 %-wt. graphene and 5 %-wt. PVDF was
measured in the ball-on-disk configuration under ambient condi- prepared with solvent NMP (Sigma-Aldrich) using the ultrasonic
tions. It was found that the composite film significantly improves dispersion technique previously described. To produce the gra-
friction and wear reduction (ca. 90%) relative to unlubricated contact. phene coating, 100 %-wt. graphene was ultrasonically homogenized
Following tribological testing, Raman spectroscopic analysis of pro- with solvent NMP. Lamination of the reference composites onto the
duced wear scars reveals a persistent protective film on both contact disk substrate were performed using the spin-coating technique
surfaces. It is theorized that zinc oxide enables durable binding of previously described.
graphene to the contact surfaces, enabling friction and wear reduc-
tion under the unusual conditions of high contact pressure inside the
contact area. 2.2. Characterization techniques

X-Ray powder diffraction (XRD) was performed with an X-ray


2. Experimental diffractometer (Smartlab III, Rigaku Corp.) with a cross-beam optics
system. For powder materials, approximately 2 mg of graphene,
2.1. Material synthesis zinc oxide, or PVDF were packed into the cavity of borosilicate
sample holders (Rigaku) to packing depth ca. 2 mm. For the com-
Synthesis of zinc oxide: Zinc oxide powder was prepared by posite coating, the mixed composite dispersion was dried inside the
calcination of zinc acetate dihydrate (Sigma Aldrich Corp.). The zinc cavity at temperature 80  C in vacuo. Loaded sample holders were
precursor was loaded into a rectangular aluminum oxide crucible then mounted into the theta-theta goniometer (Rigaku). Mono-
(MTI Corp.) and placed within a horizontal quartz tube furnace chromatic Cu-Ka radiation was produced with a 9 kW rotating
(MTI) under continuous compressed air flow at a rate of ca. anode X-ray source, and collected with a sodium iodide scintillation
100 mL min 1. The furnace was heated at uniform temperature rate detector (Rigaku). Spectral patterns were produced in the 2q scat-
10  C min 1 to a dwell temperature of 500  C for 2 h. After cooling to tering angle range 2e150 at scanning rate 0.5 min 1. Reported
room temperature, the product was ground and homogenized using spectral patterns are smoothed for clarity of interpretation, but not
a mortar and pestle. The collected zinc oxide powder was utilized in reduced for background. For diffractograms of materials containing
the following procedures without further processing (Fig. 1). graphene, the intense (002) feature has been truncated to improve

Fig. 1. Synthesis of zinc oxide, graphene, and binder composite film. Zinc acetate, heated in a continuous air stream, is oxidized to zinc(II) oxide. This product is homogenized
with graphene and polyvinylidene fluoride binder using N-Methyl-2-pyrrolidone solvent. The resulting mixture is laminated and dried to produce the solid lubricant film. (A colour
version of this figure can be viewed online.)
A.A. Alazemi et al. / Carbon 123 (2017) 7e17 9

visibility of less intense features. sliding contact between the stationary ball and the rotating disk. The
Scanning Electron Microscopy (SEM) was performed using a dual- stationary specimen was a stainless steel ball with diameter of
beam scanning electron microscope (Quanta 3D FEG, FEI Co.). For 6.3 mm and surface roughness Ra of 60 nm, and the rotating spec-
powder materials, approximately 2 mg of graphene, zinc oxide, or imen was a stainless steel disk with diameter of 70 mm and surface
PVDF were adhered to an aluminum sample stage using double-sided roughness Ra of 20 nm. Applied normal load was varied from 5 to
carbon tape (3 M Corp). For the composite coating, the mixed 15 N (average Hertz contact pressure 0.71e1.02 GPa) and sliding
dispersion was dried on the stage at temperature 80  C in vacuo. distance was varied from 145 to 450 m. Tribology tests were repeated
Loaded sample stages were placed inside the microscope chamber at least three times with error of measured friction and wear below
and evacuated to high vacuum (i.e., <2.6 nbar). Micrographs were 5%. Prior to testing, all specimens were cleaned with anhydrous
recorded at various magnifications after thorough optimization of acetone (Sigma Aldrich) to remove surface contamination.
electron beam alignment, stigmation, focus, brightness, and contrast.
Energy dispersive X-ray spectroscopy (EDXS) was performed using
3. Results and discussion
an 80 mm2 area silicon drift detector (Oxford Instruments PLC) at
energy level 10 keV. Electron pixel maps were produced using the
3.1. Formation of graphene-based composite
AZTEC analysis software suite (Oxford Instruments).
Thermogravimetric Analysis (TGA) was performed using a
The composite film is composed of graphene, zinc oxide, and
simultaneous thermal analyzer (Q600, TA Instruments Inc).
polyvinylidene fluoride (PVDF). It is expected that the primary
Approximately 4 mg zinc acetate dehydrate were loaded into a
lubrication effects are due to graphene, while synergistic adhesion
cylindrical aluminum oxide crucible (TA Instruments). The weight
is due to zinc oxide and PVDF. Specifically, zinc oxide provides
of the crucible was tared prior to sample loading. The loaded cru-
intraphase adhesion (that is, between the composite coating and
cible was placed inside the horizontal furnace chamber under
the substrate surface) while PVDF provides interphase adhesion
continuous compressed air flow at rate 100 mL min 1. Sample mass
(within the composite coating). Characterization of this system has
was recorded during heating at uniform temperature rate 10  C
been performed to understand the chemical and material proper-
min 1 to temperature 1000  C. Reported differential thermograms
ties of the film relative to its constituents.
are smoothed for clarity of interpretation.
The synthesis of zinc oxide from zinc acetate dihydrate is a result
Raman Spectroscopy was performed using a Raman microscope
of thermal decomposition [36,37]. Thermogravimetric analysis in-
(DXR, Thermo-Fisher Scientific). The apparatus was calibrated us-
dicates the macroscale mechanism is comprised of two indepen-
ing a polystyrene calibration standard (Thermo-Fisher). For powder
dent steps (Fig. 2). The first mass loss of 15.6 %-wt. occurs at
materials, approximately 2 mg of graphene, zinc oxide, or PVDF
powders were evenly dispersed across a borosilicate microscope
slide (Fisher). For the composite coating, the mixed dispersion was
dried on the slide at temperature 80  C in vacuo. The loaded slide
was then placed inside the microscope chamber. Spectral patterns
were produced using an aperatured green laser with wavelength
532 nm, beam diameter 25 mm, and power 8 mW. A single spectral
pattern is the average of at least 3 exposures, with a collection time
of 20 s per exposure. Reported spectral patterns are smoothed and
background-reduced for clarity of interpretation. The areal D-G
intensity ratio RD/G, a relative measure of sp2 and sp3 hybridized
carbon, is calculated as the ratio of the area of the D mode spectral
peak AD to the area of the G mode spectral peak AG [34]. Raman
spectra were de-convoluted into constituent spectral peaks by
fitting each excitation feature to the pseudo-Voigt function.
Following tribology testing, ex situ Raman spectral maps were
collected for the ball and disk specimens after sliding contact under
applied normal load 10 N to a sliding distance 145 m. The wear scar
on the ball specimen was characterized both before and after
removing the visible surface film from the wear scar. The film was
removed by gently sweeping a dry fiber cloth across the specimen
surface. Each Raman spectral map was produced from at least 169
sampled points represented as a color-scaled cluster map. Each
Raman point spectrum was produced with aperatured beam
diameter 1.5 mm, and vertical and horizontal step size 2.5 mm. All
color maps represent the spectral intensity at excitation frequency
1580 cm 1 [35].

2.3. Tribological investigation

Tribological testing was performed at a steel-steel interface using


a universal mechanical tribometer (UMT-3, Bruker Corp.). An optical
surface profilometer was used to measure the arithmetic average
surface roughness Ra of the specimens and wear measurements of Fig. 2. Thermogravimetric analysis of zinc acetate. The synthesis of zinc oxide oc-
curs through a two-step mechanism. (a) Initial mass losses occur at temperature
the tested specimens. Tribological performance was measured at 117  C, corresponding to moisture evaporation, and 316  C, corresponding to oxidative
ambient conditions (i.e., 27  C and 1 atm) in the ball-on-disk decomposition. (b) The greatest mass loss of 87.2 %-wt. occurs due to oxidative
configuration: friction and wear were measured during pure decomposition of zinc acetate. (A colour version of this figure can be viewed online.)
10 A.A. Alazemi et al. / Carbon 123 (2017) 7e17

temperature 117  C due to vaporization of hygroscopic adsorbed process (Fig. 3b). Graphene exhibits intense Raman excitation at
moisture. The second mass loss of 71.6 %-wt. occurs at temperature frequencies 1332, 1577, and 2667 cm 1 corresponding to vibra-
316  C due to the thermal decomposition of the acetate anion. The tional modes D, G, and 2D, respectively [35]. The areal intensity
broad, overlapping features of the differential thermogram (Fig. 2a) ratio of the D and G peaks is calculated as 0.942 for graphene and
near 316  C suggests this decomposition includes several elemen- 1.073 for the composite film: high similarity in these values suggest
tary steps [37]. The overall mass loss of this reaction is 87.2 %-wt. the distribution of sp2 and sp3 hybridization of graphene is
producing a product yield of 12.8 %-wt. conserved in the composite film. It is important to note that the
The prepared composite coating is considered a physical spectral features of PVDF [42] and ZnO [43] are absent in the
mixture of its components. X-ray powder diffraction (XRD) dem- composite spectrum because these features produce negligible
onstrates that the crystallographic profile for the composite is a Raman intensity compared to those of graphene. Therefore, the
linear combination of the profiles for the components (Fig. 3a). The presence of the composite film can be related to intense excitation
synthesized zinc oxide exhibits sharp, spectral features of hexag- of the D or G vibrational modes (Fig. 3b).
onal zincite [38]. Graphene exhibits spectral features representing The composite film contains a uniform distribution of zinc oxide
the 002 and 100 facets of hexagonal carbon [38]. The additional and graphene (Fig. 4). Scanning electron micrographs show that the
features near scattering angles 22, 42 and 44 respectively repre- film is primarily comprised of broad micro-scale graphene parti-
sent the 009, 101, and 015 facets. These secondary features are cles. Element pixel maps from energy dispersive X-ray spectros-
characteristic of nitrate functionalization, a common consequence copy (EDS) confirm the high carbon content of the planar graphene
of graphene production through chemical exfoliation [39]. PVDF particles. Furthermore, smaller nano-scale particles are observed
exhibits diffuse spectral features, indicative of disordered materials, across the basal graphene surfaces: these high zinc content parti-
that represent a mixture of the a and b phase [40,41]. In compari- cles are zinc oxide (Fig. 4c).
son, the scattering spectrum of the composite is comprised of the From its characterization, the composite film can be interpreted
principle intensity features of its constituents. The broad, shallow, as a mono-disperse mixture of zinc oxide and PVDF among a
and irregular baseline that occurs between 12 and 42 is due to the random arrangement of graphene. Zinc oxide and PVDF are nano-
low intensity of PVDF relative to the more crystalline components. scale particles, while graphene is a micro-scale particle. Finally,
The intense features that occur near 23, 44, and 50 correspond to the applied preparation techniques do not chemically modify gra-
the high crystallinity of graphene. The remaining spectral features phene, PVDF, or zinc oxide: rather, each component retains its
correspond to the crystal facets of zinc oxide. The spectral features crystallographic order in the composite following preparation.
of the composite X-ray spectrum are a direct consequence of the
discrete nature of the composite film. 3.2. Influence of film composition on friction reduction
Similarly, Raman spectroscopy shows that the binding of carbon
within graphene remains unchanged from the film preparation At the steel-steel point interface, the tribological performance of

Fig. 3. X-ray powder diffraction (left) and Raman spectroscopy (right) of the composite coating. Preparation of the composite film does not chemically change the identity of
the film precursors. (a) The characteristic x-ray spectral features of the dried composite coating are identical to those of its constituent components. (b) However, the Raman
spectral features of the dried composite coating only match those of graphene, which demonstrates an overwhelming Raman excitation response. (A colour version of this figure can
be viewed online.)
A.A. Alazemi et al. / Carbon 123 (2017) 7e17 11

Fig. 4. Scanning electron microscopy micrographs and energy-dispersive x-ray spectroscopy of the composite coating. The composite coating shows a homogeneous dis-
tribution of zinc within a carbon-rich graphene phase. (a) SEM of the composite shows a homogeneous distribution of micro-scale sheet-like particles and nano-scale aggregates. (b)
High carbon content suggests the micro-scale particles are attributed to graphene. (c) High zinc content suggests the nano-scale particles are attributed to zinc oxide. (A colour
version of this figure can be viewed online.)

the graphene-rich lubricant is superior to the performance in composite, 0.14 for graphene and PVDF, and 0.10 for graphene only.
unlubricated contact. Under normal load of 10 N (Hertz contact The friction coefficient for the graphene/PVDF composite is greater
pressure 0.89 GPa) to sliding distance of 145 m, the graphene-based than that of graphene alone, which could be due to the additional
composite demonstrates ca. 90% friction reduction relative to friction from the rough polymer. In comparison, the friction coef-
unlubricated contact (Fig. 5). During unlubricated steel-steel ficient for the graphene/PVDF/ZnO composite is less than that of
sliding, the coefficient of friction erratically increases from ca. graphene alone, suggesting that zinc oxide simultaneously en-
0.20 to 0.90 after 30 m with unsteady behavior. This irregular hances the native friction reduction of graphene and mitigates
behavior is attributed to the generation of wear particles at the friction introduced by PVDF.
sliding surfaces. During graphene-lubricated sliding to 145 m, in In addition to reducing friction, the composite coating reduces
contrast, the coefficient of friction reaches a maximum of ca. 0.13 both wear scar formation and surface roughening. After testing
and steadily decreases to stability at 0.08 with stable and steady with normal load 10 N to sliding distance of 145 m, optical mi-
behavior during the entire tribo-test. crographs and 3D surface scans of the ball specimen illustrate that
The role of zinc oxide in the graphene-based composite was the wear scar decreases with the addition of components to the
investigated by testing alternative coating compositions: (I) gra- composite coating (Fig. 7). After unlubricated sliding, the formed
phene, zinc oxide, and PVDF binder; (II) graphene and PVDF binder; wear scar has diameter 656 mm and roughness Ra 540 nm. After
and (III) graphene only. The graphene only composition measures lubricated sliding with graphene only, the wear scar diameter and
the native friction reduction of graphene alone, while the gra- roughness are reduced to 221 mm and 87 nm, respectively. After
phene/PVDF composition measures friction reduction of graphene lubricated sliding with graphene and PVDF, the wear scar diameter
in the presence of PVDF. Under applied normal load of 10 N and and roughness are further reduced to 158 mm and 72 nm, respec-
for 145 m sliding distance, the graphene/PVDF/ZnO composite tively. In this case, enhanced wear reduction is attributed to the
demonstrates the greatest friction reduction (Fig. 6). The equilib- increased coating strength from PVDF. Finally, after lubricated
rium coefficients of friction were measured as 0.08 for the total

Fig. 6. Effect of film composition on friction reduction. While graphene shows


Fig. 5. Friction reduction from composite coating. Unlubricated sliding results in a exceptional lubrication properties alone, zinc oxide critically reduces the measured
large coefficient of friction that persists over a long sliding distance. In comparison, friction in the composite coating. Zinc oxide also mitigates the increased friction
lubricated sliding results in a drastically reduced coefficient of friction across the same resulting from the addition of PVDF to graphene. (A colour version of this figure can be
sliding distance. (A colour version of this figure can be viewed online.) viewed online.)
12 A.A. Alazemi et al. / Carbon 123 (2017) 7e17

Fig. 7. Optical micrographs and 3D surface reconstruction of composition-dependent wear scars. On the stationary specimen, dry lubricants improve the wear resistance, but
the presence of zinc oxide prevents formation of an appreciable scar. (a, e) Unlubricated sliding produces a deep, rough wear scar. (b, f) Lubricated sliding with graphene produces a
smaller, smoother wear scar. (c, g). Lubricated sliding with graphene and PVDF produces an even smaller, smoother wear scar. (d, h) The composite coating prevents formation of an
appreciable wear scar, generally preserving the specimen surface. (A colour version of this figure can be viewed online.)

sliding with graphene, zinc oxide, and PVDF, an appreciable wear


scar was not formed: the specimen surface demonstrates only a few
discernable scratches. In fact, a dark opaque film is observed in
place of a wear scar (Fig. 7d). It is expected that this residual coating
enhances friction and wear reduction by preventing direct contact
to the sliding disk surface.
The tertiary composition of the coating significantly improves
tribological performance. Compared to unlubricated steel-steel
sliding, lubricated sliding with graphene alone results in drastic
friction reduction to 10 N load and 145 m sliding distance tribo-test.
The addition of PVDF to the film also reduces wear and friction
compared to unlubricated sliding, however, friction reduction is
greater than that of graphene alone. The addition of zinc oxide to
the composite film reduces friction below that measured in either
lubricant compositions. These findings suggest the importance of
zinc oxide to the composition of this adhesive composite.

Fig. 8. Effect of applied load on friction reduction. The steady state coefficient of
3.3. Influence of normal load on tribological performance friction for the composite coating is constant under an applied load up to 15 N. (A
colour version of this figure can be viewed online.)
To quantify durability of the composite coating, the coefficient of
friction was measured under normal loads 5, 10, and 15 N, corre-
sponding to Hertz contact pressures of 0.71, 0.89, and 1.02 GPa, the coefficient of friction increases to a maximum of only 0.17, then
respectively (Fig. 8). After extended time in ball-on-disk sliding, the remains at 0.1 to long sliding distances 145 and 300 m. At sliding
coefficient of friction of the composite coating remains constant. In
all trials, the steady state coefficient of friction is initially 0.15, then
decreases to 0.09 at ca. 200 s. This value is maintained from ca. 200
s to the conclusion of the test. The consistent lubricating ability of
the composite coating is attributed to two factors: the native me-
chanical strength of graphene, with Young modulus of 1 TPa and
intrinsic stress strength of 130 GPa [12]; and the strong adhesive
effects of zinc oxide and PVDF. The high-bearing load of the com-
posite coating validates its potential application for high pressure
loads [22].

3.4. Influence of sliding distance on tribological performance

To quantify endurance of the composite coating under high load,


the coefficient of friction was measured to sliding distances of 145, Fig. 9. Effect of sliding distance on friction reduction. Under a 10 N load, the
300, and 450 m under normal load 10 N (Fig. 9). During unlu- composite coating maintains constant friction reduction up to a sliding distance of ca.
300 m. After this critical distance, friction increases while the coefficient of friction
bricated steel-steel sliding, the coefficient of friction rapidly in- becomes unstable. However, the friction of the composite coating is still well below
creases to 0.9 after ca. 25 m, then remains constant with unsteady that of the unlubricated interface. (A colour version of this figure can be viewed
behavior to a sliding distance of 145 m. During lubricated sliding, online.)
A.A. Alazemi et al. / Carbon 123 (2017) 7e17 13

Fig. 10. 3D surface reconstruction of sliding distance-dependent wear scars. On the stationary specimen, lubricated sliding enables wear protection up to ca. 300 m but
continues to prevent excessive roughness. (a) Unlubricated sliding to 145 m produces a deep, rough scar. (b) Lubricated sliding up to 145 m does not show formation of an
appreciable wear scar. (c). Lubricated sliding to 300 m produces a developing, smooth wear scar. (d) Lubricated sliding to 450 produces a shallow, smooth wear scar on the contact
surface. (A colour version of this figure can be viewed online.)

distance 450 m, the composite coating increases from 0.1 to 0.4 Raman spectroscopy, the remaining composite film is tracked by
with poor stability: this is attributed to generation of wear particles the intense vibration mode of graphene at frequency 1580 cm 1
at the sliding surfaces after natural degradation of the coating [22]. (Fig. 12). The digital photograph of the wear scar after tribology
However, even after a long sliding distance of 450 m under 10 N testing shows a dark opaque film covering the specimen surface.
load (Hertz contact pressure 0.89 GPa), the frictional losses using Correlation of the spectral color map to the photograph indicates
the composite coating is still 50% less than that of the dry unlu- the protective film contains high graphene content. In addition,
bricated configuration. While friction reduction is maintained to a scratches both within and surrounding the wear scar contain high
maximum service distance of at least 450 m, a critical distance of ca. graphene content. It is theorized that this persistent film prevents
300 m defines a significant departure in the lubrication ability at excessive wear on the ball surfaces during sliding, resulting in
the interface of the sliding surfaces. friction reduction. After removing the film, the digital photograph
In addition to reducing friction coefficient, the composite film of the wear scar does not show the dark opaque coating; however,
decreases both wear scar formation and surface roughening. After the spectral color map indicates that the edges of the wear scar and
testing with normal load 10 N, 3D surface scans of the ball specimen the surrounding scratches retain high carbon content. This signal is
indicate a change in lubrication effects after a critical distance of attributed to adhered graphene due to the strength of the signal
300 m (Fig. 10). After unlubricated sliding to 145 m, the formed relative the pristine steel surface. The persistence of graphene
wear scar has diameter 656 mm and roughness 540 nm. After around the wear scar is attributed to the zinc oxide additive: the
lubricated sliding to 145 m, the contact surface does not form an zinc-containing compound serves as a binding agent between
appreciable wear scar; this result is consistent to previous wear graphene and the stainless steel surface.
observations (Fig. 7h). After lubricated sliding to 300 m, the contact Similarly, the high graphene content of the wear track on the
surface begins to develop slight wear with diameter 135 and rotating disk specimen evidences the durability of the composite
roughness 50 nm. After lubricated sliding to 450 m, an appreciable film. The wear track on the disk was formed after application of
scar is observed with diameter 494 mm and roughness 56 nm. normal load 10 N to sliding distance 145 m. The digital photograph
Interestingly, the wear scar developed after lubricated sliding to of the wear track after tribology testing shows a rectangular track
450 m is 25% smaller but 90% smoother than that developed after populated by linear scratches (original machining marks) and black
unlubricated sliding. The perseverance of a smooth surface after particles (Fig. 13). The edges of the track are adorned by more small
300 me450 m further evidences a persistence of graphene around black particles; these materials are attributed to the decomposition
the wear scar even after sliding distance 300 m. of the composite film. The spectral color map indicates the
Sliding distance, to a critical value, has a significant influence on randomly-oriented scratches on the wear scar retain appreciable
friction reduction and wear scar formation on the stationary ball carbon content after tribology testing. Interestingly, the sliding
specimen. However, with graphene-based composite lubricant, track remains wholly undamaged: aside from surface scratches, no
sliding distance does not influence formation of the wear features significant indicators of wear are observed on the disk surface.
on the rotating disk specimen. After testing with normal load 10 N, Raman spectral mapping of the ball and disk specimens after
optical micrographs and 3D surface reconstructions of the rotating tribology testing evidence that the graphene-rich composite film is
disk specimen show no appreciable wear scar after 450 m in retained under extreme operating conditions. Furthermore, the
lubricated sliding (Fig. 11). After unlubricated sliding to 145 m persistence of the film after testing evidences the importance of a
sliding distance, a wear track with 0.4 mm in depth and about zinc-based binding agent between the graphitic carbon and the
650 mm in width was formed on the disk specimen. After lubricated contact surface.
sliding to 145 m sliding distance, the contact surface does not form This macroscale tribological study has demonstrated that low
an appreciable wear scar. This observation is consistent for lubri- coefficients of friction, under high applied loads, are viable with the
cated sliding to 300 m and 450 m. In all surface reconstructions for tertiary composite. To better assess its application, the tertiary
lubricated sliding, the intense contour features adjacent to either composite was compared to other graphene-containing solid lu-
side of the wear track are due to solid particles produced from the bricants reported in the literature according to measured coeffi-
graphene-based composite film. cient of friction and applied load (Fig. 14; a full compilation of the
reported literature can be found in the supporting information).
Since 2003, several studies have reported the use of solid lubricants
3.5. Post-diagnostic surface characterization containing graphene, graphene cognates (i.e., graphene oxide,
reduced graphene oxide), or other carbon allotropes. Across all
The high graphene content of the wear scar on the stationary studied contact forces, the friction reduction from the tertiary
ball specimen suggests the mechanical and chemical durability of composite ranks among the lowest observed in the literature to
the composite lubricant. The wear scar was formed after applica- date (Fig. 14). In fact, in the domain surrounding 101 N, the tertiary
tion of normal load 10 N to sliding distance 145 m. Identified using
14 A.A. Alazemi et al. / Carbon 123 (2017) 7e17

Fig. 11. Optical micrographs and 3D surface reconstructions of sliding distance-dependent wear tracks. On the rotating specimen, lubricated sliding enables wear protection up
to at least 450 m. (a) Unlubricated sliding up to 145 m produces a deep, rough scar. In contrast, lubricated sliding up to (b) 145 m, (c) 300 m, and (d) 450 m does not show formation
of an appreciable wear track. (A colour version of this figure can be viewed online.)

composite demonstrates the lowest coefficient of friction observed The tertiary composite is the first demonstrated use of a hybrid
with graphene-containing solid lubricants. solid state lubricant, combining the excellent tribological proper-
This study presents a dry graphene-based composite film to ties of graphene and unique adhesive properties of zinc and poly-
reduce friction and wear for high load steel-on-steel applications. vinylidene fluoride. In addition, the tertiary composite is afforded
A.A. Alazemi et al. / Carbon 123 (2017) 7e17 15

Fig. 12. Raman spectral map of stationary surface specimen. (a, b) Following tribology testing, intense carbon signal correlates to the visible graphene coating that remains on the
surface of the ball specimen. (c, d) After removing the visible layer, appreciable carbon signal remains on the surface of the ball specimen. This suggests the importance of a zinc-
based binding agent between the carbonaceous graphene and the stainless steel surface. (A colour version of this figure can be viewed online.)

Fig. 13. Raman spectral map of disk substrate surface. (a, b) Following tribology testing, appreciable carbon signal correlates to the scratches that adorn the wear track formed on
the rotating specimen. The intense carbon signal on either side of the wear track are due to residual surface particles from the composite coating film. (A colour version of this figure
can be viewed online.)
16 A.A. Alazemi et al. / Carbon 123 (2017) 7e17

References

[1] B.J. Hamrock, S.R. Schmid, B.O. Jacobson, Fundamentals of Fluid Film Lubri-
cation, CRC press, 2004.
[2] H. Huang, J. Tu, L. Gan, C. Li, An investigation on tribological properties of
graphite nanosheets as oil additive, Wear 261 (2006) 140e144.
[3] M. Kalin, J. Kogovsek, M. Remskar, Mechanisms and improvements in the
friction and wear behavior using MoS2 nanotubes as potential oil additives,
Wear 280 (2012) 36e45.
[4] A.A. Alazemi, V. Etacheri, A.D. Dysart, L.-E. Stacke, V.G. Pol, F. Sadeghi, Ultra-
smooth submicrometer carbon spheres as lubricant additives for friction and
wear reduction, ACS Appl. Mater. Interfaces 7 (2015) 5514e5521.
[5] A.A. Alazemi, A.D. Dysart, X.L. Phuah, V.G. Pol, F. Sadeghi, MoS2 nanolayer
coated carbon spheres as an oil additive for enhanced tribological perfor-
mance, Carbon 110 (2016) 367e377.
[6] M. Sgroi, F. Gili, D. Mangherini, I. Lahouij, F. Dassenoy, I. Garcia, et al., Friction
reduction benefits in valve-train system using IF-MoS2 added engine oil,
Tribol. Trans. 58 (2015) 207e214.
[7] J. Zhou, Z. Wu, Z. Zhang, W. Liu, Q. Xue, Tribological behavior and lubricating
mechanism of Cu nanoparticles in oil, Tribol. Lett. 8 (2000) 213e218.
[8] N. Nemati, M. Emamy, S. Yau, J.-K. Kim, D.-E. Kim, High temperature friction
Fig. 14. Literature comparison of friction reduction from graphene-containing and wear properties of graphene oxide/polytetrafluoroethylene composite
solid lubricants. In the 101 N contact force domain, the tertiary composite demon- coatings deposited on stainless steel, RSC Adv. 6 (2016) 5977e5987.
strates one of the lowest coefficients of friction among solid lubricants containing [9] J.C. Spear, B.W. Ewers, J.D. Batteas, 2D-nanomaterials for controlling friction
graphene (green), graphene oxide (orange), reduced graphene oxide (purple), or other and wear at interfaces, Nano Today 10 (2015) 301e314.
carbon allotropes (blue). (A colour version of this figure can be viewed online.) [10] D. Berman, S.A. Deshmukh, S.K. Sankaranarayanan, A. Erdemir, A.V. Sumant,
Extraordinary macroscale wear resistance of one atom thick graphene layer,
Adv. Funct. Mater 24 (2014) 6640e6646.
[11] A.A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. Teweldebrhan, F. Miao, et al.,
by its simplicity in preparation, requiring only the mixture and Superior thermal conductivity of single-layer graphene, Nano Lett. 8 (2008)
lamination of its base components. While several studies about 902e907.
[12] C. Lee, X. Wei, J.W. Kysar, J. Hone, Measurement of the elastic properties and
graphene as a solid lubricant reported in the literature demonstrate
intrinsic strength of monolayer graphene, Science 321 (2008) 385e388.
improved wear and friction reduction, many are unable to with- [13] A. Klemenz, L. Pastewka, S.G. Balakrishna, A. Caron, R. Bennewitz, M. Moseler,
stand criticism concerning durability and stability under high load Atomic scale mechanisms of friction reduction and wear protection by gra-
conditions [21,22]. Thus, this work has demonstrated that the ter- phene, Nano Lett. 14 (2014) 7145e7152.
[14] D. Berman, A. Erdemir, A.V. Sumant, Graphene: a new emerging lubricant,
tiary composite can achieve very low wear and friction under a high Mater. Today 17 (2014) 31e42.
contact load up to 15 N (1.02 GPa average Hertz contact pressure). [15] D. Marchetto, C. Held, F. Hausen, F. Wa €hlisch, M. Dienwiebel, R. Bennewitz,
Friction and wear on single-layer epitaxial graphene in multi-asperity con-
tacts, Tribol. Lett. 48 (2012) 77e82.
[16] L.-Y. Lin, D.-E. Kim, W.-K. Kim, S.-C. Jun, Friction and wear characteristics of
4. Conclusion
multi-layer graphene films investigated by atomic force microscopy, Surf.
Coat. Technol. 205 (2011) 4864e4869.
The tertiary, graphene-rich composite has proven exceptional [17] C. Lee, X. Wei, Q. Li, R. Carpick, J.W. Kysar, J. Hone, Elastic and frictional
performance as a solid-state lubricant under high contact pressure. properties of graphene, Phys. Status Solidi B 246 (2009) 2562e2567.
[18] T. Filleter, J.L. McChesney, A. Bostwick, E. Rotenberg, K. Emtsev, T. Seyller, et
Tribology testing, under applied loads up to 15 N (Hertz contact al., Friction and dissipation in epitaxial graphene films, Phys. Rev. Lett. 102
pressure 1.02 GPa), demonstrates the composite film retains fric- (2009), 086102.
tion reduction for at least 300 m with stable behavior. Further [19] O. Penkov, H.-J. Kim, H.-J. Kim, D.-E. Kim, Tribology of graphene: a review, Int.
J. Precis. Eng. Man. 15 (2014) 577e585.
testing, under constant load and different sliding distances, evi- [20] F. Wa €hlisch, J. Hoth, C. Held, T. Seyller, R. Bennewitz, Friction and atomic-
dences the significance of zinc oxide to friction reduction and wear layer-scale wear of graphitic lubricants on SiC (0001) in dry sliding, Wear
scar preventing or smoothing. Characterization before tribotesting 300 (2013) 78e81.
[21] M.-S. Won, O.V. Penkov, D.-E. Kim, Durability and degradation mechanism of
suggests that the produced composite film is a physical mixture of graphene coatings deposited on Cu substrates under dry contact sliding,
its precursors, with no chemical or crystallographic modification Carbon 54 (2013) 472e481.
resulting from the preparation process. Spectral mapping after [22] D. Berman, A. Erdemir, A.V. Sumant, Reduced wear and friction enabled by
graphene layers on sliding steel surfaces in dry nitrogen, Carbon 59 (2013)
tribotesting confirms the persistence of the composite film on both
167e175.
the rotating and stationary contact surfaces. The durability and [23] A. Erdemir, Low-friction materials and coatings, in: Multifunctional materials
resilience of the graphene-based coating prove its great potential as for tribological applications, Pan Stanford, 2015, pp. 259e290.
[24] A. Moezzi, A.M. McDonagh, M.B. Cortie, Zinc oxide particles: synthesis,
a solid lubricant for dry sliding and high load-bearing applications.
properties and applications, Chem. Eng. J. 185 (2012) 1e22.
[25] A. Mathiazhagan, R. Joseph, Nanotechnology-a new prospective in organic
coating-review, Int. J. Chem. Eng. Appl. 2 (2011) 225e237.
Acknowledgements [26] C. Wu, F. Li, Y. Zhang, L. Wang, T. Guo, Formation and field emission of
patterned zinc oxide-adhering graphene cathodes, Vacuum 89 (2013) 57e61.
[27] L. Taylor, H. Spikes, H. Camenzind, Film-forming Properties of Zinc-based and
Prof. Vilas Pol acknowledges the Purdue University School of
Ashless Antiwear Additives, SAE Technical Paper, 2000. Paper No. SAE 2000-
Engineering and the School of Chemical Engineering for start-up 01-2030.
funding that supported this work. The authors acknowledge [28] A.M. Barnes, K.D. Bartle, V.R. Thibon, A review of zinc dialkyldithiophosphates
(ZDDPS): characterisation and role in the lubricating oil, Tribol. Int. 34 (2001)
Kyungho Kim for his assistance in performing scanning electron
389e395.
microscopy and energy-dispersive X-ray spectroscopy experi- [29] A.H. Battez, R. Gonza lez, J. Viesca, J. Ferna
ndez, J.D. Fern
andez, A. Machado, et
ments. The authors acknowledge Nick Humphrey for his assistance al., CuO, ZrO2 and ZnO nanoparticles as antiwear additive in oil lubricants,
in fabricating a specimen adapter for the DXR Raman microscope. Wear 265 (2008) 422e428.
[30] M.A. Nicholls, T. Do, P.R. Norton, M. Kasrai, G.M. Bancroft, Review of the
lubrication of metallic surfaces by zinc dialkyl-dithiophosphates, Tribol. Int.
38 (2005) 15e39.
Appendix A. Supplementary data [31] N. Gosvami, J. Bares, F. Mangolini, A. Konicek, D. Yablon, R. Carpick, Mecha-
nisms of antiwear tribofilm growth revealed in situ by single-asperity sliding
contacts, Science 348 (2015) 102e106.
Supplementary data related to this article can be found at http://
[32] E. Ferrari, K. Roberts, M. Sansone, D. Adams, A multi-edge X-ray absorption
dx.doi.org/10.1016/j.carbon.2017.07.030.
A.A. Alazemi et al. / Carbon 123 (2017) 7e17 17

spectroscopy study of the reactivity of zinc di-alkyl-di-thiophosphates anti- J. Mass Spectrom. 39 (2004) 1202e1208.
wear additives: 2. In situ studies of steel/oil interfaces, Wear 236 (1999) [38] R.W.G. Wyckoff, Crystal structures, vol. 2, Interscience, New York, USA, 1963.
259e275. [39] M.I. Saidaminov, N.V. Maksimova, P.V. Zatonskih, A.D. Komarov, M.A. Lutfullin,
[33] M. Nicholls, P. Norton, G. Bancroft, M. Kasrai, G.D. Stasio, L. Wiese, Spatially N.E. Sorokina, et al., Thermal decomposition of graphite nitrate, Carbon 59
resolved nanoscale chemical and mechanical characterization of ZDDP anti- (2013) 337e343.
wear films on aluminumesilicon alloys under cylinder/bore wear conditions, [40] E.H. Abdelhamid, O.D. Jayakumar, V. Kotari, B.P. Mandal, R. Rao, V.M. Naik, et
Tribol. Lett. 18 (2005) 261e278. al., Multiferroic PVDF-Fe3O4 hybrid films with reduced graphene oxide and
[34] G.A. Zickler, B. Smarsly, N. Gierlinger, H. Peterlik, O. Paris, A reconsideration of ZnO nanofillers, RSC Adv. 6 (2016) 20089e20094.
the relationship between the crystallite size La of carbons determined by X- [41] S. Satapathy, S. Pawar, P. Gupta, K. Varma, Effect of annealing on phase
ray diffraction and Raman spectroscopy, Carbon 44 (2006) 3239e3246. transition in poly (vinylidene fluoride) films prepared using polar solvent,
[35] I. Childres, L.A. Jauregui, W. Park, H. Cao, Y.P. Chen, Raman spectroscopy of Bull. Mater. Sci. 34 (2011) 727e733.
graphene and related materials, New Dev. Photon Mater. Res. (2013) 1e20. [42] I. Elashmawi, L. Gaabour, Raman, morphology and electrical behavior of
[36] T. Arii, A. Kishi, The effect of humidity on thermal process of zinc acetate, nanocomposites based on PEO/PVDF with multi-walled carbon nanotubes,
Thermochim. Acta 400 (2003) 175e185. Results Phys. 5 (2015) 105e110.
[37] A.V. Ghule, K. Ghule, C.Y. Chen, W.Y. Chen, S.H. Tzing, H. Chang, et al., In situ [43] T.C. Damen, S. Porto, B. Tell, Raman effect in zinc oxide, Phys. Rev. 142 (1966)
thermo-TOF-SIMS study of thermal decomposition of zinc acetate dihydrate, 570e574.

Você também pode gostar