You are on page 1of 6

PHYSICAL REVIEW MATERIALS 4, 033603 (2020)

Plowing-induced nanoexfoliation of mono- and multilayer MoS2 surfaces

Alper Özoğul ,1 Felix Trillitzsch,1 Christof Neumann,2,3 Antony George,2,3 Andrey Turchanin,2,3 and Enrico Gnecco 1
1
Otto Schott Institute of Materials Research (OSIM), Friedrich Schiller University Jena, 07743 Jena, Germany
2
Institute of Physical Chemistry, Friedrich Schiller University Jena, Lessingstrasse 10, 07743 Jena, Germany
3
Jena Center for Soft Matter (JCSM), Philosophenweg 7, 07743 Jena, Germany

(Received 12 November 2019; accepted 26 February 2020; published 10 March 2020)

The surface structures resulting from scratching monolayer MoS2 on SiO2 and multilayer MoS2 using diamond
nanotips with normal forces up to 3 μN and scan velocities up to 10 μm/s in ambient conditions are compared. In
both cases the damage process initiates with normal forces of about 2 μN. As shown by postmortem AFM images
the monolayer is peeled off along the zigzag direction of MoS2 and folded in the form of flat triangular flakes.
Wrinkling and multiple folding can be associated to debris accumulation originated from the SiO2 substrate,
which is also scratched by the tip. On the multilayer MoS2 thicker chips with linear size of about 100 nm
and height of few tens of nm are rolled out. The chipping is associated to a distinct stick-slip motion of the
AFM tip with repetition period logarithmically increasing with the velocity but definitely below the chip length,
suggesting that the process occurs through a series of exfoliation events. Simple energetic considerations allow
to explain why peeled monolayers tend to fold over the untreated surface while thicker layers tend to bend and
form chips.

DOI: 10.1103/PhysRevMaterials.4.033603

I. INTRODUCTION by its chemical inertness, strength and easy shear capability


[4]. MoS2 performs well in a broad range of temperatures
The ultimate goal of lubrication is reducing energy con-
and, oppositely to graphene, also in dry environments [5].
sumption with obvious advantages in terms of profitability
While the frictional properties of MoS2 have been intensively
and, even more important, in reducing greenhouse gases in
investigated on the nanoscale [6–8], wear properties have
Earth’s atmosphere. Well-known applications include bearing, surprisingly attracted less interest, as compared to other 2D
shafts, blades, slitters, gears, pistons, chains and several com- materials. The use of AFM allowed us to use the same tip to
ponents in automotive, aerospace and marine industry. Tra- scratch the surfaces while recording topography and lateral
ditionally, lubrication is performed using oily and fat fluids, force variations accompanying the scratch process, and to
which unavoidably leads to serious environmental problems image the resulting nanostructures with much lower normal
such as the disposal of used substances. In some cases, forces. Admittedly, a source of inspiration of the present
these lubricants may be also inapplicable, e.g., when they investigation has been the study of nanoindentation and nano-
evaporate in a vacuum environment, which is required for scratches on graphene on platinum presented by Klemenz
aerospace applications. It is thus of primary importance to et al. in 2014 [9]. In that work, the authors characterized
develop alternative lubrication techniques avoiding the use of the tribological response of the coated metal as a function
fluids and, before testing them, to characterize and possibly of the load applied by a diamond tip up to the rupture of
understand their functioning under model conditions such as the graphene layer. The measurements were accompanied
repeated scratching of the lubricated surface by means of e.g. by atomistic simulations showing that the friction coefficient
an indenter of well-defined geometry. In this context the use abruptly jumps when graphene breaks up and the tip starts
of solid lubricants, and specifically of 2D materials, appears sliding on the bare Pt surface. Here we move in a different
as promising, feasible and scalable. Graphene, molybdenum direction and try to correlate the friction variations to the
disulphide (MoS2 ), tungsten disulphide (WS2 ), and hexagonal structures formed on the scratched MoS2 surfaces. Although
boron nitride (hBN), in the form of solid powders, already the stress required to initiate damage is practically the same
find commercial application as additives to traditional fluid in on monolayers and bulk surfaces, the evolution and the results
engine oil and diesel fuel [1]. The same powders can be also of the wear process are fundamentally different.
efficiently used to “lubricate” a dry contact, i.e., to prevent the
occurrence of wear by coating the surfaces they are aimed to
protect [2]. II. MATERIALS AND METHODS
In this work, we compare the mechanical performance of MoS2 triangles were grown by chemical vapor deposition
mono- and multilayer MoS2 under the action of a diamond on a Si wafer with a 300-nm-thick oxide layer. They presented
nanotip scratching the surfaces in a controlled way under a variable size distribution (up to few tens ofmicrometers) and
ambient conditions. The most important lubricant properties were randomly oriented on the substrate, as shown in Ref. [10]
of MoS2 have been recently reviewed by Vazirisereshk et al. (Fig. S1). The edges of the islands correspond to the zigzag
[3] The frictional response of this material is ultimately ruled direction of MoS2 [11]. On the top of larger islands secondary

2475-9953/2020/4(3)/033603(6) 033603-1 ©2020 American Physical Society


ALPER ÖZOĞUL et al. PHYSICAL REVIEW MATERIALS 4, 033603 (2020)

FIG. 1. (a) Tapping-mode AFM image of a MoS2 monolayer (with one or two overlayers in the colored areas) after scratching with FN =
3 μN and v = 100 nm/s forth and back (two bottom lines) or FN = 3 μN and v = 1 μm/s (two top lines). (b) Lock-in topography image
corresponding to the rectangular area in (a). The inset shows a high resolution (tapping-mode) image corresponding to the rectangle. The
yellow and blue arrows indicate wrinkling and folding effects, respectively. (c) Cross-section along the white line in (b).

layers may be formed, which is a well-known outcome of III. RESULTS AND DISCUSSION
the CVD growth process [12]. A detailed description of the
A. Nanoscratching of MoS2 monolayers on SiO2
preparation process is given in [13]. The bulk sample is
commercially available natural MoS2 . Representative results of plowing wear on MoS2 islands
The optical microscopy image was taken with a Zeiss Axio on SiO2 (AFM topography images recorded after damaging)
Imager Z1.m microscope equipped with a 5 megapixel CCD are shown in Figs. 1(a) and 1(b). Note that the surface was
camera (AxioCam ICc5) in bright field operation. For optimal scratched along a direction forming an angle of about 45o
absorption contrast of the 2D material, Si substrates with with respect to the closest edge of the triangle [Ref. [10] (Fig.
300 nm of oxide layer were used. This contrast is due to S3)], which corresponds to the zigzag orientations of the Mo
interference, with the oxide substrate acting as a spacer, as and S atoms [11]. A normal force FN ≈ 2 μN was required to
explained for instance in Ref. [14]. initiate damage, independently of the scan velocity. The MoS2
The scanning electron microscopy images were obtained is peeled off during the scratch test in the form of flakes which
by a Zeiss Sigma VP at a beam energy of 15 kV using the are multiply folded (as discussed below) but not completely
in-lens detector of the system. detached from the substrate. The bare SiO2 regions remaining
Scratch tests were performed with a NanoWizard 4 AFM after peeling [Fig. 1(b)] have a triangular shape with cutoff
(JPK Instruments) by using AIST-NT D300 AFM cantilevers line initially following the zigzag direction of MoS2 (as seen
with single crystal diamond tips (spring constant k = 35 N/m, also from comparison with the orientation of the island) but
nominal tip radius r ≈ 20 nm) glued on them [Ref. [10] (Fig. suddenly turning by 90o along the armchair direction and
S2)]. Normal force calibration of the cantilevers was done by ending back up in the scratch line. The lateral force (friction)
the thermal noise deflection method, as described in Ref. [15]. profile acquired while scratching is rather irregular [as seen
Lateral force calibration of the cantilevers were done by the in the example given in Ref. [10] (Fig. S4)] as compared to
wedge calibration method, as described in Ref. [16]. Tip the profiles on the bulk material surface presented in the next
velocity and normal force were set constant during scratching. section. The fluctuations of the friction force are in the order
Both samples were scratched with different tip velocities and of 500 nN and bear no clear resemblance with the morphology
normal forces sufficient for scratching MoS2 , in the range of of the scratched surface.
1–3 μN. Lateral deflection of the cantilever and tip height As seen in Fig. 1(b) the material displaced by the diamond
were recorded while scratching in real time. Samples were tip is arranged asymmetrically with respect to the scratch line,
scanned before and after scratching in tapping mode with which is possibly due to the chosen direction of motion (far
the same tip. Scratches were made parallel to the step edges from the zigzag and armchair crystallographic orientations)
on MoS2 triangles (zigzag direction) [11], and in arbitrary and/or to the structure of the very end of the tip. More details
directions on the bulk sample. To minimize the effect of on the morphology of the displaced material are given in the
build-up on the tip, scratches in each set were made in an inset. On the lower side of the scratch the peeled layers show
irregular order. evidence of wrinkling (yellow arrows) and multiple folding

033603-2
PLOWING-INDUCED NANOEXFOLIATION OF MONO- AND … PHYSICAL REVIEW MATERIALS 4, 033603 (2020)

accumulated onto the tip when scratching and released when


imaging in tapping-mode.
On the upper side of the scratch line a very different
pattern is observed. It consists in a series of extrusions with
well-defined shapes and orientations. The pattern has a reg-
ular repetition distance of about 60 nm, which is possibly
due the formation of surface ripples on the SiO2 substrate
FIG. 2. Possible scenarios (not in scale) of (a) single layer peel- during scratching. This hypothesis is supported by a recent
ing in the presence of an obstacle, as deduced from Fig. 1(b), and investigation on the early stages of abrasive wear on a bare
(b) multiple layers chipping induced by the diamond tip sliding on a SiO2 surface showing that regular ripple structures are indeed
MoS2 monolayer on SiO2 and on bulk MoS2 , respectively. formed using larger diamond tips but comparable shear stress
values [17]. Interestingly, in Ref. [17], the ripples appear only
as a little remnant of the SiO2 scratched away by the tip, the
(blue arrows). Note also two aggregates (black stars) pinning remaining part being mostly pulverized. In the present case,
the folded flakes to the groove. The cross-section in Fig. 1(c) they are visible only on one side of the scratch groove, which
suggests that the MoS2 is reshaped when it is peeled off is possibly due to the uneven stress distribution resulting from
according to the tentative scheme in Fig. 2(a). The primary the tip asymmetry.
folding is trapped between the tip and the secondary folding Our observations are consistent with recent experiments by
pinned by the agglomerate, and turned into a wrinkle, which Tran Khac et al. [18], who also observed the detachment of
is finally overcome by the tip elastically driven by the AFM MoS2 from scratch grooves made by AFM, but did not resolve
cantilever. The internal structure of the aggregates could not the shape and orientation of the resulting surface structures.
be resolved. Since their height (up to ∼100 nm) is well above Folding and wrinkling of MoS2 , as induced by local shear
that of a MoS2 layer (0.65 nm) they are possibly formed stress, were hypothesized by Barboza et al. [19] but, to the
by SiO2 wear particles built up by the tip during scratching. best of our knowledge, never observed simultaneously in
Interestingly, the shape of the bare SiO2 and the MoS2 flakes nanoscratch tests. Experimental and theoretical studies on
are well visible also in SEM images of the damaged areas fracture mechanics of monolayer MoS2 by Wang et al. are also
(Fig. 3, corresponding to a different scratch test), but this is in line with our observations [20]. In Ref. [20], in situ TEM
not the case for the aggregates. It is thus not excluded that images show indeed that crack propagation can move from the
the aggregates are simply formed by material, including SiO2 , zigzag direction to the armchair direction (and vice versa) all
of a sudden, which is similar to the behavior of the monolayer
in Fig. 1(b). Accompanying molecular dynamics (MD) simu-
lations also suggested that the turning points correspond to S
vacancies in the monolayer. Assuming a radius of curvature R
in the order of 1 nm, as estimated from the height of the folded
structures in Fig. 1(c), and a bending stiffness D = 9.6 eV
[21], the strain energy density required to fold a monolayer is
Efold = D/(2R2 ) ≈ 0.8 J/m2 . Additional energy (0.22 J/m2 )
[22] is required to separate MoS2 from SiO2 (0.22 J/m2 ). The
sum of these values is much lower than the work (per unit
area) done by the friction force: W = F/w ≈ 8 J/m2 , where
F ≈ 500 nN is the order of the friction fluctuations and w ≈
60 nm is the width of the scratch groove. It means that most of
the energy dissipation accompanying the scratch process went
into rupture of Si-O bonds in the substrate. A simple calcula-
tion confirms this finding. Since the area of the cross-section
of a scratch groove A ≈ 150 nm2 , the energy dissipation per
unit volume Ediss = F/A = 3.3 × 109 J/m3 . Considering that
SiO2 has a density ρ = 2650 kg/m3 , a molar mass m =
60.08 g and an average bond energy Ebond = 621.7 kJ/mol
[23], one gets Emelt = ρ × Ebond /m = 2.71010 J/m3 for the
energy required for melting the substrate. Since Ediss is one
order of magnitude less than Emelt , it is therefore apparent that,
similar to microscratch tests in Ref. [17], SiO2 is pulverized
in the form of tiny nanoparticles but not fluidized during the
scratch process.

B. Nanoscratching of multilayer MoS2


FIG. 3. SEM images of a MoS2 monolayer scratched under Figure 4 gives an overview of the structures observed
similar conditions to those in Fig. 1. on a bulk material surface scratched under similar loading

033603-3
ALPER ÖZOĞUL et al. PHYSICAL REVIEW MATERIALS 4, 033603 (2020)

FIG. 4. AFM topography image of the bulk MoS2 surface after scratching. The set points are 1, 1.5, 2, 2.5 and 3 μN from left to right. Tip
velocities during scratch are 100 nm/s, 320 nm/s, 1 μm/s, 3.2 μm/s, and 10 μm/s from bottom to top. The arrow corresponds to the scratch
direction.

conditions as the monolayer. In this case, we performed section. Every time the friction force reaches a critical value
25 scratch tests with five different scan velocities and five corresponding to the peaks in Fig. 5(b), the chip detachment
different values of the normal force FN . The minimum force advances (with possible sliding of inner layers past each other
required to damage the surface and build up MoS2 structures [28]) until the process ends and a new chip starts to be formed
is 2 μN, i.e., in the same range observed for the MoS2 at a moment not clearly defined in the friction profiles.
monolayer on SiO2 . Above this threshold, a series of chips The average slope k of the lateral force curves in the
with variable length and height (up to 150 nm and ∼75 nm, stick phase, which corresponds to the lateral stiffness of the
respectively) are formed for all values of the scan velocities. tip-substrate contact [29], has been estimated for all values
At the threshold value of FN = 2 μN the series is present only
along limited sections of the scratched line (if v < 10 μm/s).
Note that we use the term “chips” to identify these structures
since the comparison between their height and the depth of
the scratch groove (two to five monolayers) implies that the
structures must contain voids. Due to the high curvature (R ∼
5−20 nm) the chips could not be resolved in details as the
flakes in Sec. III A. On the other hand, important information
could be inferred from the friction force variations recorded
while scratching.
In Fig. 5, the friction variations are overlapped to cor-
responding cross-sections of the chips, as recorded by the
same tip after scratching. The friction fluctuates irregularly
(with rms of up to 2.2 μN) on the undamaged part of the
MoS2 surface, whereas a clear saw-tooth profile is observed
when the chips are built up. These profiles are reminiscent
of the stick-slip motion accompanying the rippling of brittle
[24] or compliant surfaces [25] scratched by a silicon tip.
Differently from those cases, where the stick-slip repetition
d corresponds to the distance between two ripples, d is less
than the linear size of the chips, although still in the tens of
nm range (see below). It is also noticeable that the baseline
of the friction force follows the topography of the scratch
by increasing in the areas of build-up, whereas the stick-slip
motion does not show any noticeable correlation with it. The
stick slip is thus indicating discontinuous advancement of the
tip accompanied by detachment and bending but no folding
of a chip as schematically shown in Fig. 2(b). It is indeed
well-known that the bending stiffness D of MoS2 layers grows
rapidly with their thicknesses and the strain energy even more.
For instance, the bending stiffness of a MoS2 flake formed by
three monolayers is one order of magnitude larger than the
value of D = 9.6 eV corresponding to one monolayer only
[26]. At the same time the strain energy becomes two orders of FIG. 5. (a) Topography profiles after scratching (thin curves) and
magnitude higher than the interlayer binding energy between overlaid lateral deflection signals during scratches (thick curves) cor-
monolayers, 0.32 J/m2 [27]. As a result, the chips simply responding to the scratch lines marked with asterisks in Fig. 4. (b) In-
bend with radii of curvature of few tens of nm but do not sets corresponding to the rectangle in (a). Vertical bars correspond to
fold on the substrate as the monolayers in the previous sub- 2 nm (continuous lines) and 1 μN (dashed lines), respectively.

033603-4
PLOWING-INDUCED NANOEXFOLIATION OF MONO- AND … PHYSICAL REVIEW MATERIALS 4, 033603 (2020)

dependency is ultimately ruled by the detailed dynamics of the


bending and exfoliation processes leading to chip formation,
and could be better understood by multimillion atom MD
simulations. In this context the lateral force signals recorded
while scratching allow us to estimate that the average energy
dissipated in a slip event is Ediss = F d, where F = kd
is the friction force drop. With k = 40 N/m and d = 15 nm,
Ediss = 9 × 1015 J. Dividing this value by the interlayer bind-
ing energy between monolayers, 0.32 J/m2 (see above) we
conclude that the detached area is A = 0.028 μm2 at most,
corresponding to a width A/d = 1.9 μm of the exfoliated
chips. This is one order of magnitude larger than the width
values observed in Fig. 3, suggesting that most of the energy
FIG. 6. Lateral stiffness k and stick-slip repetition distance d released in a slip was used for bending the chips rather than
measured while scratching the multilayered MoS2 with different set for exfoliating them.
points and scan velocities v. Dashed lines: exponential fits (for k) and
logarithmic fits (for d) for different set points.
IV. CONCLUSIONS
of FN and v leading to chip formation. As shown by Mazo
Summing up, we have characterized the early stages of
et al. [30], k is an important indicator for possible ageing
plowing wear on the nanoscale on MoS2 in both mono-
mechanisms in the frictional contact between a nano tip and a
and multilayer forms. The large normal force values lead-
crystal surface. If the area of contact Acon grows with time, k
ing to surface failure (∼2 μN) confirm the high quality of
increases according to a relation determined by the contact
this material in both forms for applications as solid surface
geometry [31]. In Fig. 6(a), the stiffness k is plotted as a
coating. Depending on the number of removed layers two
function of the scan velocity. k is found to decrease by a factor
different scenarios are observed. In the case of the monolayer,
3 up to v = 1 μm/s, and more modestly above this value,
a regular peeling mechanism and folding/wrinkling processes
whereas the buildup is rather independent of v, as seen from
are observed. Most of the energy dissipated while sliding is
the results in Fig. 4. On the other hand, an increase in FN
used for pulverizing the SiO2 substrate, and not for cutting
does not influence k significantly, although it leads to more
and reshaping the MoS2 flakes. On multilayered MoS2 , the
damage and larger buildup. Altogether these results indicate
top surface layers are removed in the form of chips which are
that, while contact ageing must occur at low speeds, it cannot
progressively exfoliated as the scratch process goes on. This
be attributed to the direct contact between the diamond tip and
hypothesis is substantiated by the stick-slip character of the
the evolving MoS2 surface alone. Considering that the scratch
lateral force variations, which has been observed with velocity
tests were performed under ambient conditions, a possible
values ranging on two orders of magnitude and, in ambient
explanation is the formation of a water meniscus in the rough
conditions, is possibly influenced by capillary condensation.
contact area formed in the scratch process. An increase of
The discussion originated by the experimental results in
friction caused by capillary condensation at the step edges
Secs. III A and III B suggest several follow-ups of our
of graphite has been demonstrated by MD simulations of
work. For instance, the role of atom vacancies in the peeling
AFM scanning performed by Egberts et al. [32]. Simple
mechanism could be further investigated by increasing their
adsorption of water molecules at step edges of MoS2 has
number e.g. using Ar + ion bombardment [35]. It would be
been recently demonstrated in MD simulations, which also
also interesting to repeat the scratch tests in different liquids
showed that oxidation is less likely to occur [33]. Assuming
to study the influence of the environment on the exfoliation
that capillary condensation takes place between the AFM tip
process. Last but not least, repeating the measurements on
and the edges of the exfoliated MoS2 , the longer is the contact
MoS2 islands grown on a SiO2 layer with thickness below 100
time the thicker is the meniscus. This leads to a logarithmic
nm and/or imaging the resulting structures with different tips
decrease in the velocity dependence of friction [34] and to
than those used for scratching may help to shed light on the
a comparable effect in the lateral stiffness which, similarly to
nature of the aggregates similar to those observed in Fig. 1.
friction, increases with the contact area. To test this possibility,
we have repeated the scratch test in dry nitrogen. As shown
in Ref. [10] (Fig. S5) k is velocity independent in this case,
ACKNOWLEDGMENTS
supporting the plausibility of our assumption.
The average value of the repetition distance d, on the This work was partially supported by funding through
other hand, increases logarithmically with the scan velocity the European Union’s Horizon 2020 and DFG research and
v [Fig. 6(b)]. This is very different from the behavior that we innovation programme FLAG-ERA under a grant TU149/9-
recently observed on SiO2 , where d was found to increase lin- 1, DFG Collaborative Research Center SFB 1375 “NOA”
early with v [17]. As discussed also in [17], the relation d (v) Project B2 and the Thuringian Ministry for Economy, Science
depends on the velocity dependence of the kinetic friction in and Digital Societies via FGR 0088 “2D-Sens.” We thank
the slip phase, a direct measurement of which is beyond the Stephanie Höppener and Ulrich S. Schubert for enabling the
time resolution of any standard AFM. In the present case, this SEM characterization.

033603-5
ALPER ÖZOĞUL et al. PHYSICAL REVIEW MATERIALS 4, 033603 (2020)

[1] T. A. Saleh, Nanotechnology in Oil and Gas Industries (Springer [17] E. Gnecco, J. Hennig, E. Moayedi, and L. Wondraczek, Phys.
International Publishing, Cham, Switzerland, 2017). Rev. Mater. 2, 115601 (2018).
[2] T. W. Scharf and S. V. Prasad, J. Mater. Sci. 48, 511 (2013). [18] B. C. Tran Khac, F. W. DelRio, and K. H. Chung, ACS Appl.
[3] M. R. Vazirisereshk, A. Martini, D. A. Strubbe, and M. Z. Mater. Interfaces 10, 9164 (2018).
Baykara, Lubricants 7, 57 (2019). [19] A. P. M. Barboza, H. Chacham, C. K. Oliveira, T. F. D.
[4] J. C. Spear, B. W. Ewers, and J. D. Batteas, Nano Today 10, 301 Fernandes, E. H. M. Ferreira, B. S. Archanjo, R. J. C. Batista,
(2015). A. B. De Oliveira, and B. R. A. Neves, Nano Lett. 12, 2313
[5] D.-H. Cho, J. Jung, C. Kim, J. Lee, S.-D. Oh, K.-S. Kim, and (2012).
C. Lee, Nanomaterials 9, 293 (2019). [20] S. Wang, Z. Qin, G. S. Jung, F. J. Martin-Martinez, K. Zhang,
[6] C. Lee, Q. Li, W. Kalb, X. Z. Liu, H. Berger, R. W. Carpick, M. J. Buehler, and J. H. Warner, ACS Nano 10, 9831 (2016).
and J. Hone, Science 328, 76 (2010). [21] J. W. Jiang, Z. Qi, H. S. Park, and T. Rabczuk, Nanotechnology
[7] T. Filleter, J. L. McChesney, A. Bostwick, E. Rotenberg, K. V. 24, 435705 (2013).
Emtsev, T. Seyller, K. Horn, and R. Bennewitz, Phys. Rev. Lett. [22] D. Lloyd, X. Liu, N. Boddeti, L. Cantley, R. Long, M. L. Dunn,
102, 086102 (2009). and J. S. Bunch, Nano Lett. 17, 5329 (2017).
[8] Z. Ye, A. Balkanci, A. Martini, and M. Z. Baykara, Phys. Rev. [23] T. Miura, Y. Benino, R. Sato, and T. Komatsu, J. Eur. Ceram.
B 96, 115401 (2017). Soc. 23, 409 (2003).
[9] A. Klemenz, L. Pastewka, S. G. Balakrishna, A. Caron, R. [24] A. Socoliuc, E. Gnecco, R. Bennewitz, and E. Meyer, Phys.
Bennewitz, and M. Moseler, Nano Lett. 14, 7145 (2014). Rev. B 68, 115416 (2003).
[10] See Supplemental Material at http://link.aps.org/supplemental/ [25] E. Gnecco, P. Pedraz, P. Nita, F. Dinelli, S. Napolitano, and P.
10.1103/PhysRevMaterials.4.033603 for an optical microscope Pingue, New J. Phys. 17, 032001 (2015).
image of the MoS2 triangles on SiO2 , an overview of a scratched [26] G. Casillas, U. Santiago, H. Barrón, D. Alducin, A. Ponce, and
island used to identify the scan direction, a series of friction M. José-Yacamán, J. Phys. Chem. C 119, 710 (2015).
force profiles while scratching a MoS2 monolayer and the [27] T. Björkman, A. Gulans, A. V. Krasheninnikov, and R. M.
lateral stiffness estimated when scratching bulk MoS2 in dry Nieminen, Phys. Rev. Lett. 108, 235502 (2012).
nitrogen. [28] D. M. Tang, D. G. Kvashnin, S. Najmei, Y. Bando, K. Kimoto,
[11] A. M. Van Der Zande, P. Y. Huang, D. A. Chenet, P. Koskinen, P. M. Ajayan, B. I. Yakobson, P. V. Sorokin, J. Lou,
T. C. Berkelbach, Y. You, G. H. Lee, T. F. Heinz, D. R. and D. Golberg, Nat. Commun. 5, 3631 (2014).
Reichman, D. A. Muller, and J. C. Hone, Nat. Mater. 12, 554 [29] R. W. Carpick, D. F. Ogletree, and M. Salmeron, Appl. Phys.
(2013). Lett. 70, 1548 (1997).
[12] S. Wang, Y. Rong, Y. Fan, M. Pacios, H. Bhaskaran, K. He, and [30] J. J. Mazo, D. Dietzel, A. Schirmeisen, J. G. Vilhena, and E.
J. H. Warner, Chem. Mater. 26, 6371 (2014). Gnecco, Phys. Rev. Lett. 118, 246101 (2017).
[13] A. George, C. Neumann, D. Kaiser, R. Mupparapu, T. Lehnert, [31] B. Luan and M. O. Robbins, Nature (London) 435, 929 (2005).
U. Hübner, Z. Tang, A. Winter, U. Kaiser, I. Staude, and A. [32] P. Egberts, Z. Ye, X. Z. Liu, Y. Dong, A. Martini, and R. W.
Turchanin, J. Phys. Mater. 2, 016001 (2019) Carpick, Phys. Rev. B 88, 035409 (2013).
[14] C. T. Nottbohm, A. Turchanin, A. Beyer, R. Stosch, and A. [33] G. Levita and M. C. Righi, ChemPhysChem 18, 1475
Gölzhäuser, Small 7, 874 (2011). (2017).
[15] J. E. Sader, I. Larson, P. Mulvaney, and L. R. White, Rev. Sci. [34] E. Riedo, F. Levy, and H. Brune, Phys. Rev. Lett. 88, 185505
Instrum. 66, 3789 (1995). (2002).
[16] D. F. Ogletree, R. W. Carpick, and M. Salmeron, Rev. Sci. [35] S. Bertolazzi, S. Bonacchi, G. Nan, A. Pershin, D. Beljonne,
Instrum. 67, 3298 (1996). and P. Samorì, Adv. Mater. 29, 1606760 (2017).

033603-6

You might also like