You are on page 1of 19

6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface

face chemistry - Nanos…

View PDF Version

DOI: 10.1039/D3NR01261A (Paper) Nanoscale , 2023, Advance Article

Nanoscale friction and wear behavior of a CVD-grown aged WS2 monolayer: the role of

wrinkles and surface chemistry†

Himanshu Rai a, Deepa Thakur b, Aayush Gadalc, Zhijiang Ye*c, Viswanath Balakrishnan *b
and Nitya Nand Gosvami *a
a Department of Materials Science and Engineering, Indian Institute of Technology Delhi, Hauz Khas, New

Delhi 110016, India. E-mail: ngosvami@iitd.ac.in


b School of Engineering, Indian Institute of Technology Mandi, Himachal Pradesh 175075, India. E-mail:

viswa@iitmandi.ac.in
c Department of Mechanical and Manufacturing Engineering, Miami University, Oxford, OH 45056, USA. E-

mail: zye@miamioh.edu

Received 19th March 2023 , Accepted 10th May 2023

First published on 10th May 2023

Abstract
Friction reduction by transition metal dichalcogenide (TMD) monolayers is well documented;
however, wrinkle formation on the surface of TMDs takes place due to strain relaxation over time
and leads to the deterioration of the tribological properties at a small scale. Herein, we report the
role of wrinkles on the wear behavior of a chemical vapor deposition (CVD) grown aged WS2
monolayer and the comparison with wrinkle-free regions. Atomic force microscopy (AFM) was
utilized to perform load-dependent experiments, and we noticed that the wear initiated near
https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBuO… 1/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

wrinkles resulted in the disintegration of the monolayer. In contrast, in the wrinkle-free regions,
wear occurred at significantly higher loads, similar to that of freshly grown WS2, although the
coefficient of friction (COF) was increased due to the changes in surface chemistry as a result of
aging, which was confirmed using X-ray photoelectron spectroscopy (XPS). In the presence of
wrinkles, a ten-fold reduction in the load-carrying capacity was observed compared to the wrinkle-
free regions. Molecular dynamics (MD) simulations were used to corroborate experimental findings,
which demonstrate the role of wrinkles in the initiation of wear due to the stress concentration
under sliding nanocontacts near the wrinkles. In addition, simulations help establish a relationship
between the adsorbed chemical species on the surface and increased COF.

1. Introduction

Understanding nanoscale friction and wear is essential to the fundamental comprehension of energy
dissipation and material failure. 1 For small-scale devices, such as microelectromechanical systems
(MEMS), as well as for flexible electronic devices, surface-related issues, such as friction and wear, are a
serious concern as at a small-scale the surface area-to-volume ratio increases substantially, which results
in stiction and eventually reduces the life of the devices. 2

Recent research has shown that two-dimensional (2D) materials excel as solid lubricants owing to their
remarkable tribological characteristics. 3 However, finding a suitable 2D material to reduce friction and

wear efficiently is still a challenge. The discovery of graphene in 2004 4 and other layered materials in
subsequent years has made a tremendous impact on small-scale tribological research. Among all 2D
materials, graphene is the most researched layered material, and researchers have explored graphene in
both ways: as a thin solid lubricant and as an additive to oils. 5 The tribological behavior of graphene has

been explored extensively from the macro to the nanoscale. 6–11 Nevertheless, better tribological
properties also depend upon the defect-free synthesis of 2D materials; however, it has been realized that
achieving a defect-free flat 2D material when the length in one dimension surpasses a threshold number, is
difficult. 12 One of the defects that destroy the flatness of 2D materials is the self-folding phenomenon

that leads to the formation of wrinkles. 13 A recent study on graphene investigated the frictional behavior
of wrinkled graphene grown by chemical vapor deposition (CVD). 14 Wrinkles on CVD-grown graphene can

be observed quite commonly, and their orientation can play a role in modulating the friction significantly.
A lower coefficient of friction (COF) was observed in a perpendicular sliding direction to the wrinkles than

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBuO… 2/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

that in the parallel direction. An analysis revealed that the puckering mechanism dominates the higher
friction of wrinkled graphene. 14

In addition to graphene, there are numerous 2D materials that are less explored on a small scale in
tribology, such as transition metal dichalcogenides (TMDs). Ky, Dinh Le Cao et al. investigated the frictional
behavior of a CVD-grown MoS2 monolayer at various temperatures and compared it with that of exfoliated
MoS2. 15 800 °C was reported as an optimum temperature for the defect-free CVD growth of MoS2

monolayers. In a very interesting study performed by Cai et al., Mo-, W- and S-based alloys (MoxW1−xS2)
were synthesized, and nanoscale friction was examined using atomic force microscopy (AFM). Moving from
the MoS2 to the WS2 region resulted in a decrease in friction force due to the lower number of excited
phonons with weak interfacial interactions. 16 Huang et al. have investigated the wear behavior of

monolayer MoS2 using a ball-on-disk configuration and reported the failure of MoS2 mainly due to in-plane
stretching. 17 In a recent study, Özoǧul et al. performed AFM experiments on monolayer and multilayer

MoS2 flakes using a diamond nanotip and reported that the damage in the layers is initiated at around 2
μN in both cases, although their wear mechanisms were different. 18 Rapuc et al. studied the nanoscale

frictional behavior of monolayer TMDs and noticed that the TMDs were able to significantly reduce and
postpone wear under high load conditions. Even at a load of 10 μN, no wear was evident on the WS2

monolayer; however, a pre-worn tip was employed to conduct the experiments. 19 Rai et al. explored the
nanoscale wear behavior of a CVD-grown WS2 monolayer and observed that the spatial variation in the

load-carrying capacity between the interior and the edge was attributed to the number of defects. 20
Despite several studies on the tribological behavior of TMDs, the studies on the nanoscale friction and
wear behavior of TMDs are very limited and require significant attention.
In the present work, we have investigated a CVD-grown WS2 monolayer, which has been kept under

ambient conditions (i.e., atmospheric pressure and room temperature) in a sealed container for around
one year (aged). The density of wrinkles increased on the aged sample as observed topographically using
AFM. Load-dependent AFM investigations were carried out to observe the role of wrinkles in the wear
behavior of an aged WS2 monolayer. We report the onset of wear near wrinkles, which proceeds to peel

away the monolayer from the underlying substrate (SiO2/Si). However, in wrinkle-free regions, wear is
initiated at significantly higher loads, similar to that for a freshly grown WS2 monolayer; however, an

increased COF was observed. Our molecular dynamics (MD) simulations complement the experimental
results and demonstrate the involvement of wrinkles in the onset of wear and change in surface chemistry
resulting in increased COF in wrinkle-free regions.

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBuO… 3/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

2. Experimental methodology

2.1 Synthesis of monolayer WS2 via CVD

The WS2 growth was carried out using an atmospheric pressure CVD method similar to that in our recently

reported work. 21 In short, WO3 nanorods were placed in a quartz boat in a tube furnace. At the opening of

the tube furnace, sulfur (S) powder (99.5% Alfa Aesar) was kept, and then 99.9% pure Ar + H2 (95% + 5%)

was introduced at a flow rate of 100 standard cubic centimeters per minute (sccm). The furnace was first
heated to 550 °C in 20 minutes; thereafter, the temperature was raised to 850 °C in 30 minutes, where
growth occurred in 10 minutes. The temperature of the S powder was maintained at around 220 °C during
the growth process. 21

2.2 Atomic force microscopy (AFM)

The tribological behavior of a CVD-grown aged WS2 monolayer under ambient conditions was examined

using a commercial atomic force microscope (AFM, Flex Axiom, and Drive AFM, Nanosurf, Switzerland).
Diamond-like carbon (DLC) coated silicon cantilevers (Multi75DLC, Budgetsensors, Bulgaria) with normal
spring constant of 2.3 N m−1 were used for the measurements. The lateral and normal spring constants of

the cantilevers were calibrated using methods proposed by Sader 22 and Green et al. 23 According to the

beam geometry technique, the lateral force and normal force were calibrated. 24 After performing AFM
experiments, a large area scanning of the experimental region was also carried out to track the changes in
the morphology and tribological properties. The contact imaging mode was used for all AFM
measurements to simultaneously record topography and friction data. It is to be noted that, in this work,
all the experiments were performed using the same AFM probe under identical ambient conditions. The
line profiles of the topography and friction force were plotted using commercially available software,
including Gwyddion and Origin 9.0.

2.3 Raman, photoluminescence and X-ray photoelectron spectroscopies

Raman and photoluminescence (PL) spectra were acquired with a LabRAM HR evolution (Horiba Jobin
Yvon, France) confocal microscope. Raman and PL (514 nm, 50× objective, 0.2 mW LASER power) spectra
were used to characterize the vibrational mode. A Thermo Scientific Nexsa with an Al Kα source gun was
used to perform X-ray photoelectron spectroscopy (XPS) experiments. Throughout the measurement, the
scanning spot size was maintained at 400 μm.

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBuO… 4/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

2.4 Molecular dynamics simulations

Fig. 1 depicts an atomistic model of the apex of a diamond AFM tip placed on a wrinkled WS2 monolayer

surface on an amorphous SiO2 substrate. The WS2 monolayer was 10 × 10 nm2 in size, and the atoms in the

underlying amorphous SiO2 substrate were treated as a rigid body. A diamond tip apex with a

hemispherical shape (radius 2 nm) was used. The height of the wrinkle is also 2 nm. The atoms at both
ends of the WS2 monolayer were fixed to prevent relative sliding. The uppermost atoms of the tip were

modeled as a rigid body exposed to normal loads varying from 0 to 25 nN and attached to a support that
moved laterally at a constant speed of 1 m s−1 through a harmonic spring. The spring's horizontal stiffness

was 8 N m−1; however, it did not resist motion in the vertical direction. A Langevin thermostat was used to
keep the temperature of the unbound atoms at 300 K.

Fig. 1 A snapshot showing the MD simulation setup. The image shows a diamond tip moved over the
wrinkled WS2 monolayer surface, supported by an amorphous SiO2 substrate.

AIRERO 25 and Stillinger–Weber 26 potentials were used to explain the inter-atomic interactions

between the tip and the WS2 monolayer. The Lennard–Jones (LJ) potential was used to simulate the long-
range interactions between the substrate, WS2 and the tip. The mixing rule was used to acquire the LJ

parameters. LAMMPS simulation software 27 was used to carry out the simulations.

3. Results and discussion

3.1 Initial characterization

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBuO… 5/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

To confirm the monolayer nature of the aged WS2 flakes, characterization was performed using Raman and

PL emission spectroscopies, and the results are shown in the ESI Fig. S1. † A vibration is Raman active if it
causes a change in polarizability, where E′ (in-plane vibration mode) near 356 cm−1 and A′1 (out of plane)
near 417 cm−1 are the signature vibration modes of monolayer WS2, while the signal at 520 cm−1 is the

SiO2/Si substrate peak as shown in Fig. S1a (ESI). † 21 WS2 in monolayer form is a semiconductor with an

∼2 eV bandgap, and it provides high PL emission; 21 in contrast, bilayer, trilayer and multilayer WS2 have

indirect bandgaps, resulting in substantially weaker PL emission. 28 ESI Fig. S1b † shows the PL spectrum
of the monolayer WS2 having an emission intensity of ∼12k, which is the signature peak of monolayer WS2.

XPS analysis was performed on the freshly grown WS2 monolayer. The presence of tungsten (W) and S

was observed in the survey spectrum, as shown in Fig. S2 (ESI). † It is common that environmental
hydrocarbons are adsorbed to the material contributing to a carbon (C) peak in the XPS spectrum. It is to
be noted that the signals of oxygen (O) originate from the substrate as the substrate is 300 nm SiO2-coated

Si. The presence of W4+ and S2− proves the formation of WS2 (Fig. S2 † ).

3.2 AFM imaging of the wrinkles

Wrinkles are ubiquitous in 2D materials that form during fabrication. 13 In the present work, the CVD

process is used to grow the WS2 monolayer; therefore, a thermal expansion coefficient (TEC) mismatch

between the SiO2 substrate (TEC = 2.4 × 10−7 K−1) 29 and the WS2 monolayer (TEC = 1.4 × 10−5 K−1) 30 during

cooling is the likely reason for the formation of wrinkles. However, the density of wrinkles was negligible in
the freshly grown sample as we did not observe wrinkles in our earlier work. 20 The formation of wrinkles
increased considerably over time in the aged sample, as observed in the AFM images shown in Fig. 2 . An
increase in the density of wrinkles over time is likely the result of strain relaxation developed in the WS2
monolayer during CVD growth due to the thermal mismatch between the monolayer and the substrate. 31

Fig. 2a shows a low-magnification AFM topographic image of the randomly oriented wrinkles in the aged
WS2 monolayer sample. Most of the wrinkles do not have any specific orientation and height that can be

observed through color contrast as shown in Fig. 2a (wrinkles are brighter in contrast). The corresponding
friction force mapping is shown in Fig. 2b . Wrinkles exhibit higher friction 14 (bright contrast in Fig. 2b )
than the adjacent wrinkle-free regions. Similar to the topographic image ( Fig. 2a ), the friction in the
wrinkled regions is not similar and varies with the wrinkle height (more height, more friction). Fig. 2c
shows high-magnification AFM topography of a wrinkle with a height of ∼5 nm (the line profile across the
wrinkle is shown in the inset). However, each wrinkle is different in height (generally 5–10 nm), but most of

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBuO… 6/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

the wrinkles have a sharp peak, as shown in the line profile of Fig. 2c (inset). The width of the wrinkle is
∼50 nm at the bottom and decreases towards the top; most of the wrinkles are in the range of several
micrometers in length. Fig. 2d illustrates the friction force map of a wrinkle at high magnification showing
higher friction at the wrinkle than that at the adjacent wrinkle-free regions. The friction loop across the
wrinkle (shown in the inset, Fig. 2d ) shows an abrupt increase in friction. We observed that the friction
increased when the AFM tip encountered the wrinkle as the contact area is increased and reached a
maximum and then decreased when the tip slid down the wrinkle due to the reduction in the contact
area. 32 A high-magnification 3D topographic AFM image of a wrinkle revealing height contrast between
the wrinkle and the WS2 monolayer is shown in Fig. S3 (ESI). †

Fig. 2 (a) A low magnification AFM topographic image and (b) a friction force map of the aged WS2
monolayer showing wrinkles. (c) A higher magnification AFM topographic image of a wrinkle showing
the height profile in the inset, and (d) the corresponding friction force map of the wrinkle showing higher
friction on the wrinkle. A friction loop across the wrinkle is shown in the inset.

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBuO… 7/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

3.3 Friction and wear behavior of the wrinkled region

AFM experiments were carried out to investigate the friction and wear of the wrinkled region in the aged
WS2 monolayer. During friction measurement in AFM, the scan direction of the cantilever was

perpendicular to the axial direction. The wrinkled region (image size: 5 μm × 5 μm) was repeatedly
scanned (256-lines) by increasing the normal load in small increments from ∼11 nN to ∼350 nN, using a
DLC-coated silicon AFM tip. Furthermore, the experimental region was imaged with a larger scan size
(zoomed-out image: 10 μm × 10 μm) under a minimal normal load (∼20 nN) to investigate the changes
produced by the sliding probes with increasing loads. Additionally, data repeatability was validated by
conducting repeated experiments on various samples with different tips.
To investigate the effect of wrinkles on the nanoscale wear behavior of a CVD-grown aged WS2
monolayer, load-dependent experiments were performed, as shown in Fig. 3a . A linear relationship
between the friction force and the applied normal load was observed; however, above 300 nN, an abrupt
increase in the friction force can be observed, which is due to the complete wear of the WS2 monolayer.
Wear of the aged WS2 monolayer was initiated at a much lower load (∼100 nN). The initiation of wear near

wrinkles at lower loads can also be seen in the ESI Video, † which shows the evolution of wear with
increasing normal load. Fig. 3b shows the zoomed-out friction force map of the experimental region
(marked as a white dashed square). The line profile across the friction force map ( Fig. 3c ) shows higher
friction force in the experimental region due to the encounter of the AFM tip with the underlying substrate.

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBuO… 8/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

Fig. 3 (a) The effect of normal load on the friction measured on a wrinkled WS2 monolayer. Above 300 nN

normal load, the friction force increased abruptly, likely due to the complete wear of the WS2 monolayer
and the tip encountering the underlying substrate (SiO2/Si). (b) A zoomed-out friction force map of the
experimental region (shown in a white dashed square). (c) The line profile across the friction force map
(white solid line drawn in the middle of (b)) shows an abrupt increase in the friction force in the
experimental region. From the upper left, clockwise around the boundary: the images show wear of the
aged WS2 monolayer with increasing normal load (the corresponding normal load is mentioned in the
inset); wear can be observed to be initiated near wrinkles. A video showing the topographical changes
and wear initiation near wrinkles is available in the ESI. †

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBuO… 9/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

Load-dependent topographical changes during the AFM experiment are shown in Fig. 3 , starting from
the upper left clockwise around the boundary. Wear initiation at lower loads can be observed near
wrinkles and progresses with the increasing normal load. MD simulations were performed to understand
the failure mechanism initiating near wrinkles. Fig. 4a shows an increased friction force in the presence of
a wrinkle on the WS2 monolayer, whereas a perfect monolayer with no wrinkles shows a gradual increase

in the friction force with the increasing normal load. A higher stress concentration leading to severe plastic
deformation events near the wrinkle can be one of the reasons for wear initiation. Eventually, the load-
carrying capacity of the wrinkled monolayer is reduced. A higher stress concentration can be observed as
shown in Fig. 4b (highlighted by the black dashed circle). Initiation of failure near the wrinkle under the tip
can be visualized in the MD simulations snapshot (Fig. S4 † ). Besides Fig. 3 , where more than one wrinkle
was considered (randomly oriented), experiments were also performed by considering individual wrinkles
in different orientations (approximately perpendicular and parallel to the scanning direction) as shown in
Fig. S5 (ESI). † Wear was initiated at a lower normal load for a perpendicular wrinkle as compared to the
parallel wrinkle (Fig. S6 † ). The wrinkle perpendicular to the scanning direction failed at 150 nN, whereas a
parallel wrinkle did not fail within a normal load range applied in the experiment and was flattened with
the increasing normal load (∼300 nN). In a similar study on graphene, Long et al. observed higher friction
when the AFM tip was scanned perpendicular to the graphene wrinkle rather than in the parallel direction;
however, the wear behavior of differently oriented wrinkles was not studied in this work. 14 Therefore,
scanning parallel to the wrinkle direction can be an effective way to delay the failure of the monolayer. In
another study, Huang et al. noticed the folding of wrinkles, resulting in formation of a bilayer and leading
to failure near wrinkles in the graphene samples; 32 however, this has not been observed in the present
study of wrinkled WS2 monolayers. Folding of wrinkles in monolayer WS2 is significantly difficult as the

bending rigidity of monolayer WS2 (∼8–10 eV) 33 is approximately one order of magnitude higher than that
of single-layer graphene (∼1.5 eV). 34

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBu… 10/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

Fig. 4 (a) The effect of normal load on the friction of a wrinkled WS2 monolayer using MD simulations. (b)
A snapshot of the MD simulation showing the stress concentration under the tip near the wrinkle in the
black dashed circle.

3.4 Friction and wear behavior of the wrinkle-free regions

Apart from wrinkles, experiments were also performed in wrinkle-free regions of the aged sample. Load
dependence friction was recorded in a wrinkle-free region (image size: 1 μm × 1 μm), and it increased
linearly with a normal load of up to ∼1000 nN, as shown in Fig. 5a . Above ∼1000 nN, an abrupt increase in
the friction force was observed due to the wear of the WS2 monolayer at higher loads, and the tip
encountered the underlying substrate. A zoomed-out friction force map (image size: 3 μm × 3 μm) ( Fig. 5b )
of the experimental region (shown in the white dashed lines) was recorded at a normal load of ∼20 nN to
observe the changes due to scanning. An experimental region (bright color) showing higher friction can be
observed in the friction force map. The line profile across the experimental region demonstrating an
increase in the friction force is shown in Fig. 5c . A line was drawn in the worn-out area of the experimental
region to observe an increase in the friction force.

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBu… 11/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

Fig. 5 (a) The effect of normal load on the friction of a wrinkled WS2 monolayer. However, above 1000 nN
normal load, the friction force increased abruptly due to the wear of the WS2 monolayer, and the tip
encounters the underlying substrate (SiO2/Si). (b) A zoomed-out friction force map of the experimental
region is shown in the white dashed square. (c) The line profile across the friction force map (white solid
line drawn in (b)) shows an abrupt increase in the friction force in the experimental region.

An increase in the COF was also observed in the wrinkle-free regions of the aged WS2 monolayer
compared to the fresh monolayer sample; however, it is to be noted that the COF was not calculated for
the wrinkled region as friction is non-uniform due to the presence of wrinkles. The COF for the aged
sample (∼0.16) in the wrinkle-free region is noticeably higher than that for the freshly grown WS2
monolayer samples. 19 This increase in COF is due to the chemical changes over the aged monolayer. A
similar phenomenon of high friction on chemically modified graphene has been observed. 35,36 High
friction on chemically modified graphene was attributed to an increase in the out-of-plane stiffness of the
modified surface. High lateral stiffness and low adhesion behavior were also observed compared to the
fresh graphene surface. 35,36 We also observed high lateral stiffness and low adhesion on the aged WS2
monolayer compared to a fresh WS2 sample in the AFM experiments ( Fig. 6 ). The adhesion force was

reduced significantly (∼50%) in the aged sample. For both samples, adhesion measurements were
performed on a 1 μm × 1 μm area using an 8 × 8 grid size (64 points). To confirm the chemical modification

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBu… 12/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

on the aged sample, XPS experiments were performed, and the results are shown in Fig. 7 . XPS analysis
shows that organic species from the environment are adsorbed, and WS2 is gradually oxidized and
possibly passivated with hydroxyl groups over time. Fig. 7a shows the presence of W, S, C and O in the XPS
survey spectrum of the aged WS2 monolayer. The high intensity of C in Fig. 7a shows the adsorption of

hydrocarbons on the surface of the aged WS2 compared to the fresh sample, as shown in Fig. S2. † Fig. 7b
and c also show the presence of hydrocarbons and OH groups adsorbed on the surface. Fig. 7d shows the
XPS spectrum of W. Here, in addition to W4+, an extra peak corresponding to W6+ is present. This higher
binding energy peak arises from the oxidation of WS2 over time. 37 The oxidation of WS2 is still only in
small amounts (calculation of the present amount is shown in the Appendix). The XPS spectrum of S2− is
shown in Fig. 7e . These adsorbed species were loosely bonded with the surface of the aged sample and
started moving when the AFM tip encountered the surface; consequently, friction is increased.
Furthermore, continuous sliding resulted in the gradual removal of the adsorbed particles from the
experimental region; therefore, friction was reduced significantly. To observe this behavior, AFM
experiments were performed in the wrinkle-free region of an aged sample and terminated before the
failure of the monolayer. Furthermore, a large area image (zoomed-out image) was captured to observe
the friction reduction in the experimental region as shown in Fig. S7 (ESI). † Reduced friction in the line
profile of the experimental area is highlighted with a black dashed circle in Fig. S7b. †

Fig. 6 (a) A comparison of the friction loop and (b) adhesion (pull-off force) between the fresh and aged
(wrinkle-free region) WS2 monolayers. The aged sample shows a higher friction and a lower adhesion
force than the fresh WS2 sample.

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBu… 13/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

Fig. 7 XPS plot of an aged WS2 monolayer. (a) The survey scan indicates the presence of W, S, C and O in

the aged sample. The intensity of C is higher in the aged sample due to the adsorption of hydrocarbons
on the monolayer surface over time. (b) C-scan, (c) O-scan, (d) W-scan and (e) S-scan.

To further investigate the effect of oxidization and adsorbed hydrocarbons on the frictional behavior of
WS2, we performed MD simulations with oxidized and hydroxyl group functionalized WS2 and WS2 with
methane molecules on the surface to mimic the experimental conditions of aged WS2, as shown in Fig. 8 .
Our results show the highest friction and wear on the hydroxyl group functionalized WS2 38 ( Fig. 8a ),

which indicates that the friction and wear behavior of the wrinkle-free regions may be dominated by the
functional groups on the surface, especially the oxidized and hydroxyl group regions.

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBu… 14/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

Fig. 8 (a) Load dependence of friction on the oxidized and hydroxyl group functionalized WS2, and WS2

with methane molecules on the surface from MD simulations, and (b), (c), and (d) the corresponding
simulation snapshots. S: blue spheres; W: -silver spheres sandwiched between two S spheres.

4. Conclusions

In conclusion, we studied the friction and wear behavior of a CVD-grown aged WS2 monolayer, and the role
of wrinkles was elucidated using AFM. Experiments were performed in a systematic manner considering
wrinkles and wrinkle-free regions. Load-dependent AFM experiments were performed to observe the effect
of wrinkles on the wear behavior of an aged WS2 monolayer. The wear was observed to be initiated near
wrinkles at a much lower load (∼100 nN) than that for the freshly grown WS2 monolayer. However,
experiments in a wrinkle-free region confirm the wear initiation at a higher load (∼1000 nN). A nearly ten-
fold reduction in load-carrying capacity was observed in the presence of wrinkles. We also observed higher
COF in the aged WS2 monolayer, which can be due to the chemical modifications on the aged surface. Our
MD simulations support the experimental findings and show similar trends for the friction and wear

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBu… 15/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

behavior of the wrinkled WS2 monolayer. In summary, our findings highlight the role of wrinkles in the
wear behavior of aged monolayer WS2 and its significance for attaining the intended tribological

properties. This work will benefit the applications where WS2 is used as a solid lubricant in various
engineering applications.

Data availability

Data will be made available upon request.

Author contributions

Himanshu Rai: data curation, investigation and writing – original draft. Deepa Thakur: formal analysis,
investigation and writing – review and editing. Aayush Gadal: formal analysis and investigation. Zhijiang
Ye: conceptualization, investigation and writing – review and editing. Viswanath Balakrishnan:
conceptualization, investigation, project administration, funding acquisition and writing – review and
editing. Nitya Nand Gosvami: conceptualization, investigation, project administration, funding acquisition
and writing – review and editing.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that
could have appeared to influence the work reported in this paper.

Appendix

From the XPS analysis, we can comment on the presence of oxidation states and their relative amounts.
The areal ratio of the W 4f7/2 peak in W4+ and W6+ oxidation states is 23.85 (W4+ : W6+). This means that W6+ is
present in very small amounts as compared to W4+.
Now, we will calculate the percentage using simple mathematics:

W4+ + W6+ = 100%. (1)

Dividing eqn (1) by W6+, we get

(W4+/W6+) + 1 = (100/W6+)% (2)

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBu… 16/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

23.85 + 1 = (100/W6+)% (3)

24.85 = (100/W6+)% (4)

W6+ = (100/24.85)% = 4.02%. (5)

Hence, the calculated percentage of W6+ is ∼4.02% and that of W4+ is 95.98%.

Acknowledgements

NNG would like to acknowledge the SERB (CRG/2020/002062) and the Indian Institute of Technology Delhi
(MI02369G) for their financial support. DT (IF-180717) would like to acknowledge DST-INSPIRE
(Department of Science and Technology – Innovation in Science Pursuit for Inspired Research) for
providing a Ph.D. fellowship.

References

1. K. S. Kim, H. J. Lee, C. Lee, S. K. Lee, H. Jang, J. H. Ahn, J. H. Kim and H. J. Lee, ACS Nano, 2011, 5,
5107–5114 CrossRef CAS PubMed .
2. D. Boer and T. M. Mayer, MRS Bull., 2001, 26, 302–304 CrossRef .
3. S. Zhang, T. Ma, A. Erdemir and Q. Li, Mater. Today, 2019, 26, 67–86 CrossRef CAS .
4. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A.
Firsov, Science, 2004, 306, 666–669 CrossRef CAS PubMed .
5. D. Berman, A. Erdemir and A. V. Sumant, Mater. Today, 2014, 17, 31–42 CrossRef CAS .
6. A. Morina, A. Neville, M. Priest and J. H. Green, Tribol. Lett., 2006, 24, 243–256 CrossRef CAS

.
7. Y. Huang, Q. Yao, Y. Qi, Y. Cheng, H. Wang, Q. Li and Y. Meng, Carbon, 2017, 115, 600–607 CrossRef

CAS .
8. B. Vasić, A. Matković, R. Gajić and I. Stanković, Carbon, 2016, 107, 723–732 CrossRef .
9. Y. Qi, J. Liu, J. Zhang, Y. Dong and Q. Li, ACS Appl. Mater. Interfaces, 2017, 9, 1099–1106 CrossRef

CAS PubMed .

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBu… 17/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

10. A. Klemenz, L. Pastewka, S. G. Balakrishna, A. Caron, R. Bennewitz and M. Moseler, Nano Lett., 2014,
14, 7145–7152 CrossRef CAS PubMed .
11. Y. Qi, J. Liu, Y. Dong, X. Q. Feng and Q. Li, Carbon, 2018, 139, 59–66 CrossRef CAS .
12. X. Meng, M. Li, Z. Kang, X. Zhang and J. Xiao, J. Phys. D: Appl. Phys., 2013, 46, 055308 CrossRef

.
13. W. Chen, X. Gui, L. Yang, H. Zhu and Z. Tang, Nanoscale Horiz., 2019, 4, 291–320 RSC .
14. F. Long, P. Yasaei, W. Yao, A. Salehi-Khojin and R. Shahbazian-Yassar, ACS Appl. Mater. Interfaces,
2017, 9, 20922–20927 CrossRef CAS PubMed .
15. D. L. C. Ky, B. C. Tran Khac, C. T. Le, Y. S. Kim and K. H. Chung, Friction, 2018, 6, 395–406 CrossRef

CAS .
16. S. Cai, Y. Tao, W. Zhao, S. Huang, C. Sun, X. An, Y. Zhang, Z. Wei, Z. Ni and Y. Chen, Tribol. Int., 2022,
166, 107363 CrossRef CAS .
17. Y. Huang, Q. Yao, Z. Lu, L. Jiao, S. Zhang, Q. Li and Y. Meng, ACS Appl. Nano Mater., 2018, 1, 7092–
7097 CrossRef CAS .
18. A. Özoǧul, F. Trillitzsch, C. Neumann, A. George, A. Turchanin and E. Gnecco, Phys. Rev. Mater., 2020,
4, 2–7 Search PubMed .
19. A. Rapuc, H. Wang and T. Polcar, Appl. Surf. Sci., 2021, 556, 149762 CrossRef CAS .
20. H. Rai, D. Thakur, D. Kumar, A. Pitkar, Z. Ye, V. Balakrishnan and N. N. Gosvami, Appl. Surf. Sci., 2022,
605, 154783 CrossRef CAS .
21. D. Thakur, P. Kumar and V. Balakrishnan, J. Mater. Chem. C, 2020, 8, 10438–10447 RSC .
22. J. E. Sader, I. Larson, P. Mulvaney and L. R. White, Rev. Sci. Instrum., 1995, 66, 3789–3798 CrossRef

CAS .
23. C. P. Green, H. Lioe, J. P. Cleveland, R. Proksch, P. Mulvaney and J. E. Sader, Rev. Sci. Instrum., 2004,
75, 1988–1996 CrossRef CAS .
24. N. N. Gosvami, J. Ma and R. W. Carpick, Tribol. Lett., 2018, 66, 1–10 CrossRef CAS .
25. S. J. Stuart, A. B. Tutein and J. A. Harrison, J. Chem. Phys., 2000, 112, 6472 CrossRef CAS .
26. F. H. Stillinger and T. A. Weber, Phys. Rev. B: Condens. Matter Mater. Phys., 1985, 31, 5262–5271
CrossRef CAS PubMed .
27. S. Plimpton, J. Comput. Phys., 1995, 117, 1–19 CrossRef CAS .
28. D. Thakur, P. Kumar, M. Sabarigresan, R. Ramadurai and V. Balakrishnan, Surf. Interfaces, 2021, 26,
101308 CrossRef CAS .
29. C. Tsou, Y.-S. Huang, H.-C. Li and T.-H. Lai, Sens. Mater., 2005, 17, 441–451 CAS .

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBu… 18/19
6/21/23, 10:12 AM Nanoscale friction and wear behavior of a CVD-grown aged WS 2 monolayer: the role of wrinkles and surface chemistry - Nanos…

30. D. Kumar, B. Singh, P. Kumar, V. Balakrishnan and P. Kumar, J. Phys.: Condens. Matter, 2019, 31,
505403 CrossRef CAS PubMed .
31. Q. Zhang, Z. Chang, G. Xu, Z. Wang, Y. Zhang, Z. Q. Xu, S. Chen, Q. Bao, J. Z. Liu, Y. W. Mai, W. Duan, M.
S. Fuhrer and C. Zheng, Adv. Funct. Mater., 2016, 26, 8707–8714 CrossRef CAS .
32. Z. Huang, S. Chen, Q. Lin, Z. Ji, P. Gong, Z. Sun and B. Shen, Langmuir, 2021, 37, 6776–6782
CrossRef CAS PubMed .
33. K. Lai, W. B. Zhang, F. Zhou, F. Zeng and B. Y. Tang, J. Phys. D: Appl. Phys., 2016, 49, 185301 CrossRef

.
34. Y. Wei, B. Wang, J. Wu, R. Yang and M. L. Dunn, Nano Lett., 2013, 13, 26–30 CrossRef CAS PubMed

.
35. J. H. Ko, S. Kwon, I. S. Byun, J. S. Choi, B. H. Park, Y. H. Kim and J. Y. Park, Tribol. Lett., 2013, 50, 137–
144 CrossRef CAS .
36. S. Kwon, J. H. Ko, K. J. Jeon, Y. H. Kim and J. Y. Park, Nano Lett., 2012, 12, 6043–6048 CrossRef

CAS PubMed .
37. J. Gao, B. Li, J. Tan, P. Chow, T. M. Lu and N. Koratkar, ACS Nano, 2016, 10, 2628–2635 CrossRef

CAS PubMed .
38. R. Li and C. Song, Crystal, 2018, 8, 167 CrossRef .

Footnote

† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3nr01261a

This journal is © The Royal Society of Chemistry 2023

https://pubs.rsc.org/en/content/articlehtml/2023/nr/d3nr01261a?casa_token=2djaO7UcOxoAAAAA:8Q-296UapXAgyqTwiC10mVq2GYt_5P1ZOzBu… 19/19

You might also like