Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Tribology International 134 (2019) 87–92

Contents lists available at ScienceDirect

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

Atomistic mechanism of the weakened wear resistance of few-layer T


graphene induced by point defects
Fang Zheng, Fangli Duan∗
State Key Laboratory of Mechanical Transmissions, Chongqing University, Chongqing 400030, China

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

Keywords: Effect of various point defects on the wear properties of few-layer graphene was investigated by molecular
Graphene dynamics simulations. Our results show that the presence of Stone-Wales defect, double vacancies and single
Wear vacancy reduces the critical normal load for onset of adhesive wear to 85, 15 and 11% of that of the intact
Molecular dynamics graphene, respectively. The analysis of the potential energy surfaces of defective graphene sheets indicates that it
Point defect
is the highest chemical reactivity induced by point defects that determines the onset of adhesive wear. This study
provides an atomic-level understanding for the weakening effects of point defects on the wear resistance of few-
layer graphene.

1. Introduction induced atomistic interlocking at the interface, and the consequent


lattice deformation and C–C bond breaking [18]. Moreover, wear re-
Graphene, a single-atom-thick film of carbon, offers a promising sistance of few-layer graphene is sensitively dependent on the nature of
solution to control friction and adhesion in various applications, espe- interaction between counterfaces. Wang and Duan found that only a
cially in micro- and nanoscale devices with moving contact [1–4]. few chemical bonds formed at the interface, usually occurred in ad-
However, many tribological experiments at the micro- and macroscale hesive wear, could lead to tearing and peeling off of graphene, resulting
have shown that graphene is fragile and easily worn out [5–8], which in a much reduced critical normal load for adhesive wear compared
seems to contradict its extreme mechanical strength and superior lu- with that for abrasive wear [19].
brication properties exhibited in the nanoscale tests [9–11]. Structural defects, such as point and line defects, are introduced
A few factors have been found to contribute to the deterioration of unavoidably into graphene lattice during the manufacturing or transfer
wear resistance of graphene at the micro- and macroscale. The step process of graphene. Structural defects will weaken the mechanical
edge of a graphene sheet supported by a substrate, when scratched by strength and increase the chemical reactivity of graphene in the local
an atomic force microscopy (AFM) probe, exhibits a remarkably re- region around themselves [20,21], which could further affect the wear
duced critical normal load for wear failure, up to ∼2 orders of mag- properties of graphene. However, the effect of structural defects on the
nitude smaller than that of intact interior region. This reduction of wear wear properties of graphene has not been fully understood [22–24]. In
resistance is resulted from two failure modes at the step edge, i.e., this paper, effect of various point defects on the wear properties of
adhesive wear via atom-by-atom removal and peel induced folding and graphene was investigated by MD simulations of scratching a diamond
rupturing [12–14]. In an AFM experiment on chemical vapor deposition tip over a few-layer graphene sheet with point defects on the first
graphene, Vasić et al. observed that tearing and peeling off of graphene graphene layer. Our results show that the presence of single vacancy
initiate exactly from wrinkles, caused probably by out-of-plane de- and double vacancies defects could reduce the critical normal load for
formations that could induce larger contact area and interaction force onset of adhesive wear to 11–15% of that of the intact graphene.
between the graphene and AFM probe [15]. Moreover, the change of chemical reactivity induced by point defects
Surface roughness of both the substrate for supporting graphene and was analyzed to understand the weakening effect of point defects on
the counterface is also considered to be a basic reason for degradation wear resistance of graphene.
of wear resistance of graphene [16,17]. By using molecular dynamics
(MD) simulation, Xu et al. found that even atomic roughness of the
surfaces could promote the rupture of graphene, resulted from the


Corresponding author.
E-mail address: flduan@cqu.edu.cn (F. Duan).

https://doi.org/10.1016/j.triboint.2019.01.035
Received 18 October 2018; Received in revised form 25 January 2019; Accepted 25 January 2019
Available online 26 January 2019
0301-679X/ © 2019 Elsevier Ltd. All rights reserved.
F. Zheng, F. Duan Tribology International 134 (2019) 87–92

Fig. 1. (a) The atomistic model of a diamond tip


scratching on a four-layer defective graphene sheet.
(b) Distribution of the four pairs of point defects
lying along the sliding path of the tip on the first
graphene layer. (c) Atomistic structures of the re-
constructed Stone-Wales (SW), double vacancies
(DV) and single vacancy (SV) defects.

2. Simulation methods graphene sheet. Then, these atomistic models were well relaxed for
40 ps at 300 K under the corresponding normal loads added on the rigid
The modeling system comprised a hemispherical diamond tip with a layer of the tip. After the relaxation, the diamond tip slid on the gra-
radius of 25 Å and a four-layer graphene sheet (Fig. 1(a)). The diamond phene sheet along the x direction with a constant speed of 10 m/s under
tip was carved from a (111)-oriented diamond crystal bulk. The surface constant normal loads [13]. All the simulations were performed under a
of the tip did not be passivated and had the dangling bonds left on it. microcanonical (NVE) ensemble. The system temperature was con-
Each graphene layer was composed of 5280 atoms with a dimension of trolled at 300 K by applying the Langevin thermostat to the three
170 Å × 81 Å along the x and y directions. Four pairs of point defects atomic layers adjacent to the rigid layers of the tip and the third atomic
were placed on the first graphene layer lying equally spaced along the layer of graphene. All the MD simulations were conducted by using
sliding path of the tip, as shown in Fig. 1(b). Three types of point de- LAMMPS software [29].
fects, Stone-Wales (SW), double vacancies (DV) and single vacancy (SV)
defects [25], were investigated in this study (Fig. 1(c)). For simplicity's
sake, we named the defective graphene sheets with SW, DV or SV de- 3. Results and discussions
fects as SW-GS, DV-GS and SV-GS, respectively, and denoted the intact
graphene sheet as intact GS. 3.1. Effect of point defects on adhesive wear
Two different types of wear, adhesive wear and abrasive wear, were
simulated in this study. For the abrasive wear simulations, the intera- To reveal the effect of point defects on adhesive wear of graphene,
tomic interactions within the diamond tip and the graphene sheet were we observed the adhesive wear processes of defective graphene sheets.
described via the AIREBO potential [26]. To model the mechanical Fig. 2 shows an adhesive wear process of the DV-GS under a normal
interaction at the interface, we used the 6-12 Lennard-Jones potential, load of 200 nN (28.9 GPa) (Video 1). During the sliding process, we
with the same parameters as in AIREBO potential, to describe the in- observed the formation of interfacial bonds at the contact surface,
teractions between the diamond tip and the graphene sheet [19,23]. For which is a prerequisite for the occurrence of adhesive wear. Due to the
the adhesive wear simulations, AIREBO potential was employed to strong adhesive interaction at the interface induced by interfacial
describe all the interatomic interactions within the whole system bonds, the further sliding of the tip leads to the rupture of in-plane C–C
[13,19,24], which means the formation and breakage of chemical bonds and the consequent formation of cavities on the topmost gra-
bonds could occur at the interface between the diamond tip and the phene layer. This adhesive wear process is consistent with that ob-
graphene sheet. served in previous MD simulation [19]. The promoting effect of point
In the adhesive wear simulations, the topmost two atomic layers of defects on adhesive wear of graphene is obvious in this example. As
the diamond tip were held rigid to apply normal loads and sliding (or shown in Fig. 2, the two cavities formed on the first graphene layer start
indentation) velocity. For the abrasive wear simulations, the whole respectively from the locations of the first and second defects. And the
diamond tip was rigid to avoid heavy plastic deformation caused by third defects increase the expanding and fracture of the second cavity,
large normal loads, which should be used to initiate abrasive wear. The which initiates from the second defects. This enhancement effect on
bottommost atomic layer of the graphene sheet was held fixed to sup- adhesive wear of graphene is resulted from the increased chemical re-
port the whole system in all simulations. To prevent relative sliding activity caused by point defects, which could lead to the formation of
between adjacent graphene layers, the left edge atoms of each graphene chemical bonds between the diamond tip and the carbon atoms of
layer were held fixed during the sliding simulations [19]. Free graphene around the point defects [13], as we will discuss in Section
boundary conditions were adopted in the x and z directions, and peri- 3.4.
odic boundary condition was employed in the y direction. Supplementary video related to this article can be found at https://
We first performed the loading processes by indenting the diamond doi.org/10.1016/j.triboint.2019.01.035
tip with a constant speed of 10 m/s along the -z direction [27,28] to To clarify the role of interfacial bonds on adhesive wear of gra-
obtain the atomistic models with various normal loads acted on the phene, we calculated the number of in-plane broken bonds on the

88
F. Zheng, F. Duan Tribology International 134 (2019) 87–92

Fig. 2. Adhesive wear process of the DV-GS under a normal load of 200 nN. The sliding distance is given in each frame. The front view shows the diamond tip and all
graphene layers, while the top view only shows the top three graphene layers. The carbon atoms of graphene with at least one original in-plane C–C bonds broken are
highlighted by bigger size balls.

graphene layers, interfacial bonds between the tip and graphene and
interlayer bonds between adjacent graphene layers during the sliding
process. A cutoff distance of 1.7 Å was used to determine the formation
and breakage of C–C bonds [30,31]. As shown in Fig. 3, the initial
formation of interfacial bonds occurs around the first defects, which
simultaneously leads to the initial breakage of in-plane C–C bonds.
When the tip slides across the second and third defects, there is also
abruptly increasing formation of the interfacial bonds, which finally
leads to a severe fracture of the first graphene layer. These results show
that the point defects of graphene could induce the formation of in-
terfacial bonds, thus may start a new or worsen an existed adhesive
wear process of the graphene sheets. Because of its excellent flexibility,
graphene could readjust its atomic configuration to adapt to surface
morphology of the tip during the sliding process, leading to a closer
contact and stronger interfacial interaction at the interface [19,32,33].
Thus, the third point defects in this example exhibit a much more severe
Fig. 3. Variation of the number of in-plane broken bonds, interfacial bonds and effect on the adhesive wear of graphene than the previous two defects.
interlayer bonds during the sliding process of the example shown in Fig. 2. It also should be noted that the interlayer bonds, due to its limited
Numbers in circle and the dashed lines indicate the locations of the four pairs of
number formed in this example, have no obvious influence on the wear
point defects.
process, therefore, this example indicates a close correlation between
wear process of graphene and the formation of interfacial bonds.

Fig. 4. Abrasive wear process of the DV-GS under a normal load of 875 nN. The sliding distance is given in each frame. The front view shows the diamond tip and all
graphene layers, while the top view only shows the top three graphene layers. The carbon atoms of graphene with at least one original in-plane C–C bonds broken are
highlighted by bigger size balls.

89
F. Zheng, F. Duan Tribology International 134 (2019) 87–92

critical normal loads and the corresponding average contact pressures


for onset of adhesive wear or abrasive wear on the intact GS, SW-GS,
DV-GS and SV-GS. The average contact pressure was estimated by di-
viding the critical normal load by the contact area, which was calcu-
lated by multiplying the total number of contact atoms by the atomic
area of graphene (2.62 Å2 per atom) [19].
For the intact GS, the critical normal loads for onset of adhesive
wear and abrasive wear are 325 and 1025 nN, respectively. The critical
normal load for adhesive wear (325 nN) is much smaller than that for
abrasive wear (1025 nN), in good agreement with the results in pre-
vious studies [13,19]. The presence of point defects exhibits an obvious
weakening effect on wear resistance of graphene. For the onset of ad-
hesive wear, the critical normal loads are 275, 50 and 35 nN for SW-GS,
DV-GS and SV-GS, respectively. Although SW defects show a relatively
small weakening effect on wear resistance of graphene, the existence of
DV and SV defects reduces the critical normal load to only 15.3 and
10.7% of that of the intact GS. For the onset of abrasive wear, the
critical normal loads are 875, 825 and 800 nN, respectively, for SW-GS,
DV-GS and SV-GS, falling in the range of 78–85% of that of the intact
Fig. 5. Variation of the number of in-plane broken bonds and interlayer bonds GS. Obviously, the weakening effect of point defects on the resistance of
during the sliding process of the example shown in Fig. 4. Numbers in circle and adhesive wear is much more remarkable compared with their effect on
the dashed lines indicate the locations of the four pairs of point defects. the resistance of abrasive wear.
We also carried out MD simulations of sliding a diamond-like carbon
3.2. Effect of point defects on abrasive wear (DLC) tip, with the same shape and size as the diamond tip, on the
defective graphene sheets (see Fig. S3). The comparison of the critical
Fig. 4 shows the abrasive wear process of the DV-GS under a normal normal loads for onset of adhesive wear between these two types of tips
load of 875 nN (109.5 GPa) (Video 2). Different from adhesive wear, is shown in Table S1. For the DLC tip systems, the critical normal loads
abrasive wear of graphene is caused by the rupture of the graphene for onset of adhesive wear are 250, 200, 30 and 15 nN, respectively, for
layers when applied load exceeds the mechanical strength of graphene. intact GS, SW-GS, DV-GS and SV-GS, exhibiting a weakening effect of
In this example, the initial rupture of the first graphene layer starts point defects similar to that for the diamond tip systems. However, the
exactly at the location of the third defects. After this initial rupture, the critical normal loads of the DLC tip systems are a little smaller than that
wear zone expands along with the sliding of the tip and finally develops of the diamond tip systems. The basic reason may be that the DLC tip
into a wear track, in which dispersed atoms from different layers are has more dangling bonds on its surface compared with the diamond tip,
mixed together, resulting in a disordered carbon structure. This kind of which will facilitate the formation of more chemical bonds between the
wear track is consistent with that observed in previous MD simulations DLC tip and the graphene sheet.
[31,34,35]. The enhancement effect of point defects on abrasive wear is
caused by the weakened mechanical strength of the first graphene layer 3.4. Higher chemical reactivity induced by point defects
around the location of the defects, which facilitates the rupture of the
graphene sheet. The formation of interfacial bonds is closely associated with the
The formation of interlayer bonds is a typical phenomenon observed chemical reactivity of defective graphene sheets. The word “chemical
in the abrasive wear process of few-layer graphene [31,34,35]. To reactivity” used here indicates the ability of graphene atoms to form
clarify the role of interlayer bonds on abrasive wear of graphene, we chemical bonds with atoms of the tip. To characterize the change of
calculated the number of in-plane broken bonds and interlayer bonds chemical reactivity induced by point defects, we calculated the poten-
during the sliding process. As shown in Fig. 5, when the tip slides across tial energy surfaces of defective graphene sheets by scanning a carbon
the second and third defects, the formation of interlayer bonds is cor- atom probe on the defective graphene sheets. At each scanning hor-
respondingly accelerated. And the onset of abrasive wear occurs exactly izontal position, the atom probe was applied a constant normal load,
at the location of the third defects, where dozens of interlayer bonds then the whole system was relaxed to its equilibrium state. The po-
have already formed beneath the diamond tip. These results indicate tential energy experienced by the carbon atom probe was calculated to
that the point defects could facilitate the formation of interlayer bonds, describe the chemical reactivity of the site where the atom probe lo-
because their weakening effect on mechanical strength easily induces cated on the defective graphene sheets.
out-of-plane deformation. We also observed that the rupture of gra- Fig. 6 shows the potential energy surfaces of the intact GS, SW-GS,
phene in abrasive wear only occurs after a certain number of interlayer DV-GS and SV-GS under a contact pressure of 38.1 GPa, which is a little
bonds have formed beneath the diamond tip. Therefore, the increased smaller than the critical contact pressure of intact GS for adhesive wear.
formation of interlayer bonds also could further facilitate the initial We used two separated color-bars in Fig. 6. The lowest potential energy
rupture and onset of abrasive wear of graphene. It should be noted that experienced by the probe atom is -0.034 eV when only physical inter-
the formation of interlayer bonds not only is an indicator of the extent action exists between the probe atom and graphene. Once the probe
of out-of-plane deformation of the graphene layers, but also contributes atom forms a chemical bond with the graphene sheet, its potential
to decrease the mechanical strength of graphene because of the sp2-to- energy will abruptly drop to below -1.8 eV. By using these separated
sp3 bonding transition [36,37]. color-bars, the chemical bonding area on the defective graphene sheet
is clearly marked by blue. As shown in Fig. 6, the point defects induce
3.3. Critical normal load for the onset of initial wear an increase of chemical reactivity in the local region around them-
selves. Under this particular contact pressure (38.1 GPa), the increased
To quantitatively characterize the effect of point defects on wear chemical reactivity of the defective region gives rise to the formation of
resistance of graphene, we carried out a series of sliding simulations chemical bond with the probe atom, in contrast to the situation of the
under various normal loads and searched out the critical normal load intact GS in which no chemical bond is formed.
for the onset of initial wear (see Figs. S1 and S2). Table 1 lists the We further extracted the maximum interaction energy and the size

90
F. Zheng, F. Duan Tribology International 134 (2019) 87–92

Table 1
Critical normal loads and the corresponding average contact pressures for onset of adhesive wear or abrasive wear on the intact GS, SW-GS, DV-GS and SV-GS.
Critical normal load/Average contact pressure (nN/GPa) Intact GS SW-GS DV-GS SV-GS

Adhesive wear 325/39.4 275/37.3 50/9.9 35/8.1


Abrasive wear 1025/124.1 875/111.3 825/106 800/103.5

of the chemical bonding area respectively from the potential energy


surfaces in Fig. 6. As shown in Fig. 7, the maximum interaction energy
is 0.034, 2.13, 6.7 and 7.8 eV, respectively, for the intact GS, SW-GS,
DV-GS and SV-GS. The order of maximum interaction energy, intact
GS < SW-GS < DV-GS < SV-GS, is exactly contrary to that of the
critical normal loads for adhesive wear (Table 1). Moreover, the DV and
SV defects exhibit chemical reactivity much higher than the SW defect,
also consistent with the remarkable reduction of critical normal load for
adhesive wear caused by DV and SV defects. It seems that it is the
highest chemical reactivity that determines the onset of adhesive wear
of graphene, even though it occupies only a very limited area. The
possible reason may be that once interfacial bonds form at the site of
the highest chemical reactivity, they will very possibly lead to the onset
of adhesive wear of defective graphene sheets.
The size of chemical bonding area is 12.92, 45.22 and 24.06 Å2,
respectively, for SW-GS, DV-GS and SV-GS. Obviously, the DV-GS ex-
hibits a much larger chemical bonding area than the SV-GS. We also
observed that the DV-GS displays larger friction and more severe wear Fig. 7. Maximum interaction energy and the size of chemical bonding area for
than the SV-GS in the sliding simulations under applied loads larger the intact GS, SW-GS, DV-GS and SV-GS extracted from the potential energy
than 100 nN (see Fig. S4). It seems that, under a heavy load, it is the size surfaces in Fig. 6.
of chemical bonding area, not the highest chemical reactivity, that
dominates the extent of adhesive wear of defective graphene.
We also calculated the potential energy surfaces of defective

Fig. 6. Potential energy surfaces detected by scanning a carbon atom probe on the (a) intact GS, (b) SW-GS, (c) DV-GS and (d) SV-GS under a contact pressure of
38.1 GPa, which is a little smaller than the critical contact pressure for adhesive wear of intact GS. Black lines show the atomic lattice of the graphene sheet.

91
F. Zheng, F. Duan Tribology International 134 (2019) 87–92

graphene sheets under a contact pressure of 9.8 GPa, which is very close on sliding steel surfaces. Carbon 2013;54:454–9.
to the critical contact pressure of DV-GS for adhesive wear (see Fig. S5). [8] Berman D, Erdemir A, Sumant AV. Reduced wear and friction enabled by graphene
layers on sliding steel surfaces in dry nitrogen. Carbon 2013;59:167–75.
The maximum interaction energy and the size of chemical bonding area [9] Lee C, Wei X, Kysar JW, Hone J. Measurement of the elastic properties and intrinsic
were extracted from the potential energy surfaces (see Fig. S6). As strength of monolayer graphene. Science 2008;321:385–8.
shown in Fig. S6, the effect of point defects on these two parameters [10] Marchetto D, Held C, Hausen F, Wahlisch F, Dienwiebel M, Bennewitz R. Friction
and wear on single-layer epitaxial graphene in multi-asperity contacts. Tribol Lett
under 9.8 GPa is similar to that under 38.1 GPa. 2012;48:77–82.
[11] Feng X, Kwon S, Park JY, Salmeron M. Superlubric sliding of graphene nanoflakes
4. Conclusions on graphene. ACS Nano 2013;7:1718–24.
[12] Vasić B, Matković A, Gajić R, Stanković I. Wear properties of graphene edges probed
by atomic force microscopy based lateral manipulation. Carbon 2016;107:723–32.
Effect of various point defects on the wear properties of few-layer [13] Qi Y, Liu J, Zhang J, Dong Y, Li QY. Wear resistance limited by step edge failure: the
graphene was investigated by MD simulations. The effect of point de- rise and fall of graphene as an atomically thin lubricating material. ACS Appl Mater
Interfaces 2017;9:1099–106.
fects on the wear process was studied by monitoring the variations of
[14] Qi Y, Liu J, Dong Y, Feng XQ, Li Q. Impacts of environments on nanoscale wear
the interfacial bonds and the interlayer bonds during sliding process. behavior of graphene: edge passivation vs. substrate pinning. Carbon
We searched out the critical normal loads for onset of initial wear on 2018;139:59–66.
the defective graphene sheets. The presence of SW defect, DV and SV [15] Vasić B, Zurutuza A, Gajić R. Spatial variation of wear and electrical properties
across wrinkles in chemical vapour deposition graphene. Carbon 2016;102:304–10.
defects reduces the critical normal load for onset of adhesive wear to [16] Huang Y, Yao Q, Qi Y, Cheng Y, Wang H, Li Q, et al. Wear evolution of monolayer
85, 15 and 11% of that of the intact graphene, respectively, while re- graphene at the macroscale. Carbon 2017;115:600–7.
duces the critical normal load to 85, 80 and 78% of that of the intact [17] Marchetto D, Held C, Hausen F, Wahlisch F, Dienwiebel M, Bennewitz R. Friction
and wear on single-layer epitaxial graphene in multi-asperity contacts. Tribol Lett
graphene, respectively, for onset of abrasive wear. Moreover, the 2012;48:77–82.
change of chemical reactivity induced by point defects was character- [18] Xu Q, Li X, Zhang J, Hu Y, Wang H, Ma T. Suppressing nanoscale wear by graphene/
ized by the potential energy surfaces of defective graphene sheets. graphene interfacial contact architecture: a molecular dynamics study. ACS Appl
Mater Interfaces 2017;9:40959–68.
These results indicate that it is the highest chemical reactivity, even [19] Wang LF, Duan FL. Nanoscale wear mechanisms of few-layer graphene sheets in-
though it occupies only a very limited area, that determines the onset of duced by interfacial adhesion. Tribol Int 2018;123:266–72.
adhesive wear of defective graphene sheets. This study provides an [20] Mortazavi B, Ahzi S. Thermal conductivity and tensile response of defective gra-
phene: a molecular dynamics study. Carbon 2013;63:460–70.
atomic-level understanding for the weakening effects of point defects [21] Sun S, Wang C, Chen M, Zheng J. A novel method to control atomic defects in
on wear properties of few-layer graphene. graphene sheets, by selective surface reactions. Appl Surf Sci 2013;283:566–70.
[22] Peng YT, Wang ZG, Zou K. Friction and wear properties of different types of gra-
phene nanosheets as effective solid lubricants. Langmuir 2015;31:7782–91.
Acknowledgments
[23] Sun XY, Wu R, Xia R, Chu XH, Xu YJ. Effects of Stone-Wales and vacancy defects in
atomic-scale friction on defective graphite. Appl Phys Lett 2014;104:183109.
The project is supported by the National Natural Science Foundation [24] Bai L, Srikanth N, Zhao B, Liu B, Liu Z, Zhou K. Lubrication mechanisms of graphene
of China (Grant No. 51775066) and the Fundamental Research Funds for DLC films scratched by a diamond tip. J Phys Appl Phys 2016;49:485302.
[25] Banhart F, Kotakoski J, Krasheninnikov AV. Structural defects in graphene. ACS
for the Central Universities of China (Grant No. Nano 2011;5:26–41.
106112017CDJPT280002). [26] Stuart SJ, Tutein AB, Harrison JA. A reactive potential for hydrocarbons with in-
termolecular interactions. J Comput Phys 2000;112:6472–86.
[27] Kilymis DA, Delaye JM. Nanoindentation of pristine and disordered silica: mole-
Appendix A. Supplementary data cular dynamics simulations. J Non-Cryst Solids 2013;382:87–94.
[28] Tavazza F, Senftle TP, Zou C, Becker CA, van Duin ACT. Molecular dynamics in-
Supplementary data to this article can be found online at https:// vestigation of the effects of tip–substrate interactions during nanoindentation. J
Phys Chem C 2015;119:13580–9.
doi.org/10.1016/j.triboint.2019.01.035. [29] Plimpton S. Fast parallel algorithms for short-range molecular dynamics. J Comput
Phys 1995;117:1–19.
References [30] Sandoz-Rosado EJ, Tertuliano OA, Terrell EJ. An atomistic study of the abrasive
wear and failure of graphene sheets when used as a solid lubricant and a compar-
ison to diamond-like-carbon coatings. Carbon 2012;50:4078–84.
[1] Nilsson L, Andersen M, Balog R, Laegsgaard E, Hofmann P, Besenbacher F, Hammer [31] Yang L, Zhang Q, Diao DF. Cross-linking-induced frictional behavior of multilayer
B, Stensgaard I, Hornekaer L. Graphene coatings: probing the limits of the one atom graphene: origin of friction. Tribol Lett 2016;62:33.
thick protection layer. ACS Nano 2012;6:10258–66. [32] Liu XZ, Li Q, Egberts P, Carpick RW. Nanoscale adhesive properties of graphene: the
[2] Nayak PK, Hsu CJ, Wang SC, Sung JC, Huang JL. Graphene coated Ni films: a effect of sliding history. Adv Mater Interfaces 2014;1:1300053.
protective coating. Thin Solid Films 2013;529:312–6. [33] Li S, Li Q, Carpick RW, Gumbsch P, Liu XZ, Ding X, et al. The evolving quality of
[3] Pu JB, Mo YF, Wan SH, Wang LP. Fabrication of novel graphene-fullerene hybrid frictional contact with graphene. Nature 2016;539:541–5.
lubricating films based on self-assembly for MEMS applications. Chem Commun [34] Zhang Q, Diao DF, Kubo M. Nanoscratching of multi-layer graphene by molecular
2014;50:469–71. dynamics simulations. Tribol Int 2015;88:85–8.
[4] Penkov O, Kim HJ, Kim DE. Tribology of graphene: a review. Int J Precis Eng Manuf [35] Zhang Q, Diao DF, Yang L. Dangling bond induced cross-linking model in nano-
2014;15:577–85. scratched graphene layers. Surf Coating Technol 2013;237:230–3.
[5] Kim KS, Lee HJ, Lee C, Lee SK, Jang H, Ahn JH, Kim JH, Lee HJ. Chemical vapor [36] Pei QX, Zhang YW, Shenoy VB. A molecular dynamics study of the mechanical
deposition-grown graphene: the thinnest solid lubricant. ACS Nano properties of hydrogen functionalized graphene. Carbon 2010;48:898–904.
2011;5:5107–14. [37] Verma A, Parashar A. The effect of STW defects on the mechanical properties and
[6] Won MS, Penkov OV, Kim DE. Durability and degradation mechanism of graphene fracture toughness of pristine and hydrogenated graphene. Phys Chem Chem Phys
coatings deposited on Cu substrates under dry contact sliding. Carbon 2017;19:16023–37.
2013;54:472–81.
[7] Berman D, Erdemir A, Sumant AV. Few layer graphene to reduce wear and friction

92

You might also like