Interactions between Atmospheric Oxygen and MoS2 Crystals: Comparison
Please note this is a comparison between Version 1 by Robert Szoszkiewicz and Version 2 by Amina Yu.

MoS2 belongs to a class of transition metal dichalcogenides (TMDCs). TMDCs share a common formula MeX2, where Me is a transition metal element from group four (Ti, Zr, Hf), five (V, Nb or Ta) or six (Mo, W), and X is a chalcogen (S, Se or Te). Their crystalline structure comprises an inner Me layer sandwiched by two X layers. Herein, thermal oxidation of the microscopic MoS2 flakes is reviewed. An impact of relative humidity is also mentioned.

  • MoS crystals
  • surface oxidation
  • surface etching
  • α-MoO

1. Introduction

Initially, microscopic and naturally occurring MoS2 crystals, aka molybdenites, were used as solid lubricants in various encapsulated devices, such as low Earth orbit (LEO) satellites [1][2]. This is thanks to their low friction and propensity to peel off easily, because of any two adjacent layers interacting weakly via the van der Waals forces [2][3][4][3,4,5]. Later in time, microscopic MoS2 crystallites and their composites found usage in batteries, photovoltaic devices and catalysts for the hydrogen evolution reaction (HER) and other catalytic processes [5][6][7][8][9][6,7,8,9,10]. In recent years, interest in single MoS2 crystals has surged in the context of nanoelectronics, particularly on flexible substrates. According to the Web of Science, a number of published papers with the “MoS2” keyword increased slowly from 164 papers in 2000 to 220 papers in 2010, but then it started to grow exponentially from 288 papers in 2011 to reach 5549 papers in 2019 and level off to 5661 manuscripts published in 2020. Coincidently, year 2011 marked appearance of the first published reports about single-layer MoS  transistors [10][11].

Most of the current research with MoS2 crystals has been performed on the 2H MoS2 polymorph. Moreover, 2H MoS2 has been shown to be a 2D semiconductor with high electron mobilities of up to 200 cm^{2} V^{−1} s^{−1} and a bulk 1.2 eV indirect bandgap, which increases and changes its character with decreasing number of the MoS2 layers, so that the 2H MoS 2 monolayer has a 1.8 eV direct bandgap [2][3]. These properties make single 2H MoS 2 flakes of variable thickness an ideal material for tightly packed nanoscale transistors and devices with variable bandgap. Recently, single-layer MoS2 based transistors, as well as their assemblies in the form of logic circuits[11][12][13] [12,13,14] and in-memory computing devices with room-temperature current on/off ratios of 1 × 10 8 and ultralow standby power dissipation in single layers of MoS2, have been developed [14][15][15,16]. At the same time, great progress has been made into synthesis of wafer-scale polycrystalline MoS2 monolayers[16] [17] and large-domain single-crystalline MoS2 monolayers by chemical vapor deposition [13][17][14,18], as well as the wafer-scale transfer and stacking of monolayer MoS2 for heterogeneous integrations [18][19][19,20]. Due to discovery of large photoluminescence quantum efficiency of the chemically treated MoS2 crystals [20][21], next-generation flexible nanoelectronics devices based on single MoS2 flakes appear already in sensing, optoelectronics and energy harvesting [2][21][22][3,22,23]. Furthermore, the presence of various conduction mechanisms beyond electronic currents, such as spin currents, polariton currents, valley and other topological currents within single MoS2 layers and/or their heterostructures with other 2D materials, has opened new vistas in the MoS2 based nanoelectronics[2][23][24] [3,24,25].

Each microelectronic device heats up while working, mostly due to Joule heating, and some MoS2 based transistors have been measured recently to locally reach 250 °C and more [25][26]. In addition, exposure to ambient conditions has been shown to reversibly reduce the on-state current in back-gated bilayer MoS2 based FETs by up to two orders of magnitude [26][27]. However, other studies on thin MoS2 FET transistors have shown up to three-fold drops in carrier mobilities in air with respect to vacuum, which did not recover upon vacuum annealing[27] [28]. Therefore, it is important to investigate the chemical surface reactivity of single MoS2 flakes in ambient conditions above room temperature. Furthermore, the impact of relative humidity cannot be neglected due to capillary condensation at local lengths[28][29] [29,30].

From the time of seminal works of Ross et al. [30][31] on Mo polycrystalline powders in the 1950s till about the 2010s MoS  crystals have been considered inert to oxidation till 600 °C in air [31][32]. Substantial progress has been made in this field, particularly after 2013, when several seminal papers about local oxidative etching have been published[32][33][34] [33,34,35]. A whole zoo of the surface reactions of 2H MoS2 crystals with oxygen, much beyond oxidative etching, has been discovered by using high-resolution tools, such as AFMs, SEMs and TEMs. These reactions involve physio and chemical oxygen adsorption, oxygen dissociative reactions, and other direct and non-direct oxidation mechanisms, as well as oxygen penetration between the stacked MoS2 layers [32][33][34][35][36][37][38][39][40][41][42][33,35,36,37,38,39,40,41,42,43]. Below, we present a brief survey of the results obtained on various MoS2 as a function of temperature and relative humidity. Despite substantial efforts interactions of oxygen with single MoS2 flakes and crystals are still not completely understood, particularly in terms of competing thermal etching mechanisms. 

2. Experimental Observations of Thermally-induced MoS2 Oxidation in Air and in Water

To start with, one must acknowledge major differences between MoS2 powders, bulk MoS2 crystals and single MoS2 flakes. Pulverized MoS2 powders comprise mostly the 2H MoS2 polymorph, and their physicochemical properties have been studied macroscopically. Bulk MoS2 crystals have been studied on the macro- and micro-scales, but mostly computationally. Single microscopic and exclusively 2H MoS2 crystalline flakes of various thickness from one monolayer (ML) to much thicker structures have been studied computationally, as well as experimentally. Research in these different MoS2 forms and such different length scales have provided certain amount of information at each length scale, which is often not directly transferable to a different length scale.

Early experimental studies of polycrystalline MoS2 powders have shown that Mo oxide layers present on the surface form slowly and act as passivating layers till at least 100 °C [30][34][31,35]. The appearance of Mo oxides at room temperature, mostly in the form of MoO3 (with other oxide forms below a detection threshold), has been confirmed as well on the CVD-grown defective MoS2 monolayers[43] [56]. However, in contrast to the MoS2 powders and CVD-grown MoS2 samples, freshly exfoliated single MoS2 flakes did not show any direct manifestations of the protective oxide layers [44]. The lack of the passivating Mo oxide layers on single mechanically exfoliated MoS2 flakes has been confirmed even in the samples left in air for a year. Single MoS2 flakes, however, have been shown to gradually increase their number of sulfur defects when left in air, which, in turn, became oxidized towards substitutional Mo species, but not MoO3[45] [60]. Nevertheless, some variations in the MoS2 stoichiometry between even the same kind of mechanically exfoliated samples also take their toll and produce detectable differences across the same samples [46][69].

As the temperature has been increased, Yamamoto et al. [34][35] showed—via local micro-Raman studies—that thin MoS2 flakes start to show any electronic density changes above 200 °C. In the light of the aforementioned studies, this might mean that accumulation of defects and oxygenated S vacancies start to become detectable microscopically only above 200 °C. At temperatures between 320 and 400 °C oxidative thermal etching regime has been observed in the case of single MoS2 flakes [32][33][34][39][33–35,40]. Such conditions have produced characteristic microscopic triangular etch pits within the MoS2 basal planes, as has been observed mostly via AFM. The etch pits were typically only 1 ML deep and have been observed on thin and thick single MoS2 flakes. Furthermore, they changed orientations in between the subsequent layers, which has been explained based on the crystal structure of the 2H MoS2 polymorph [39][40]. When single MoS2 crystals were heated for a short time, the etch pits on basal planes were associated with virtually no change in thickness and lateral dimension of the studied MoS2 crystals. Variability of the growth speeds for the triangular etch pits has been observed between MoS2 flakes, which suggested thermally activated processes being at the origin of oxidative etching[34][39] [35,40]. Ukegbu et al. have exploited this observation further to calculate activation energy of the oxidative etching process from the growth rate of the triangular etch pits[39] [40]. The triangular shape of the pits was related to the hexagonal symmetry of the Mo planes within the MoS2 crystal lattice, and several mechanisms for their creation have been proposed[32][34][39][47] [33,35,40,71]. Thanks to the S-TEM studies, a zigzag Mo (ZZ-Mo) edge was assigned as the dominant termination of the triangular etch pits [47][71]. Further details on triangular etch pits formation have been found using local Raman studies of the signature MoS2 Raman bands at 384 cm−1 (E12g mode) and 408 cm−1 (A1g mode). The observed results, particularly in the case of thin MoS2 flakes, pointed out an electron density withdrawal from the MoS2 layer and consequently its p-type doping upon progressive oxidative etching. In addition to etching two other kinds of interactions with molecular oxygen, which can also follow up oxidative etching, have been observed. In particular, the presence of small MoOx clusters, identified as mostly the MoO3 clusters, as well as presence of small MoO3 patches [33][41][34,42]. Furthermore, other processes have been spotted, such as oxygen diffusion into the freshly exposed etch pits[48] [73] and oxygen incorporation in between the MoS2 sheets, which based on the XRD results swells the MoS2 crystals.

Above 400–410 °C substantial oxidation involving decrease of the flakes’ volume has been observed in air [40][49][41,45]. It confirmed predictions about particularly fast oxidation along the crystalline edges[40][41][50][51] [41,42,46,47]. Oxidation and oxidative etching in some instances produced visible MoO3 deposits (in various forms) appearing on the single MoS2 flakes/nanosheets [41][52][53][42,48,49] or even full MoS2 transformation [35] [34]into single MoO3 crystals. Oxidation at temperatures of more than 500 °C has been shown to be detrimental even to the thick MoS2 flakes and led to formation of deep indents and likely Mo oxide patches with messy topographies[40][49] [41,45]. Those were surface morphologies without any clear pit pattern (observed previously in oxidative etching), but rather with many surface pits of various depths and various forms of surface islands with frayed edges. Often, some dendritic structures were observed. The resulting Mo oxides have been shown to leave the substrate surface, but often not entirely if oxidation was not too rapid[34][35][39] [35,36,40].

The aforementioned phenomenological aspects related to the MoOx formation on globally heated basal planes within the MoS2 flakes in dry air are summarized in Figure 1 below.

Figure 1. Phenomenological aspects of the MoOx formation onto MoS2 basal planes during their heating in dry air. Above ca. 320 °C, triangular 1 ML deep etch pits were observed[32][33][34][39] [33–35,40]. Nucleated growth of the etch pits was supplemented by appearance of small sub-nm size MoOx clusters detected already at ca. 350 °C via a combination of various experimental techniques[52] [48]. At temperatures between 370 and 390 °C, the largest amounts of Mo oxides and their derivatives accumulate on the MoS2 surface [41][42]. Some of it forms irregular patches (marked in violet)[33][41] [34,42]. Above 410 °C, less oxide has been detected indirectly and locally via AFM techniques, as well as chemically and globally via XPS studies. In addition, in this regime, substantial surface defects have been observed. 

 

In addition to the temperature related oxidation results in dry air, presented in Fig. 1, water vapors have been shown to change everything and produce larger amounts of Mo oxides and generally quicker basal plane passivation by complete and thin MoO3 layers then oxidation in air[30][35][41] [31,36,42]. To complement the aforementioned findings, the studies in water at ambient conditions have provided the following conclusions. Defective MoS2 nanosheets have been found to partially dissolve in water, particularly above pH = 2 and at concentrations of total Mo of less than 2 mM[54] [62]. Their dissolution products were various kinds of molybdate ions, with monomolybdates prevailing above pH of 6–8, depending on the report. Single non-defective MoS2 flakes have been found by some researchers to be stable in water, while other reports showed needlelike protrusions on MoS2 crystals left in water for more than 1 h. Thus, more research is needed to resolve this issue.

3. Mechanistic Details of the MoS2 Oxidative Processes

A great majority of mechanistic approaches used atomistic calculations based on the local DFT approaches, which were a primary tool for investigating defect formation energies, reaction energy barriers, reaction mechanisms and the electronic structure properties of the considered MoS2 atomistic structures.

Regarding engaging into a chemical reaction, the O2 molecules, in contrast to atomic oxygen, have been computationally shown to be non-reactive towards the S atoms. A high kinetic activation energy barrier of 1.59 eV was obtained for dissociative adsorption of the O2 molecule on 1 ML thick MoS2 . Dissociative oxygen adsorption leads to adsorbed oxygen atoms in the form of two stable oxygen-terminated sulfurs. Not only is the kinetic barrier for dissociative O2 adsorption large, but the process is thermodynamically not efficient, because its binding energy of about 0.8 eV is much less than the kinetic barrier of 1.59 eV [42][55][43,57]. Therefore, it is not expected to propagate, and single O2 dissociative events are rare. 

The findings of Santosh et al. [42][43] contributed well to explaining the exceptional stability of the pristine, non-defective MoS2 flakes against oxidation. Direct oxygen binding to Mo atoms is impossible on pristine basal planes due to steric reasons. However, other oxidative processes have been found to exist on pristine and non-defective MoS2 basal planes. For example, oxygen-induced single sulfur vacancy (SSV) creation was observed by Peto et al. [56][60], who—due to advances in local scale imaging—supplemented their DFT calculations with experimental observations.

Recently, a more comprehensive scenario for initial events associated with oxidative etching has been proposed by Farigliano et al. [72][57] for the non-defective MoS2 basal plane. They used NEB calculations at 0 K temperature to find out crucial intermediates within the processes and then proceeded with ab initio molecular-dynamics simulations at higher temperatures to show decomposition pathways of the key intermediate in the processes. The key initial intermediate consisted of an O atom adsorbed on top of an S atom (Oads) with a second O atom inserted (Oin) into the S-Mo bond, i.e., the OSOMo moiety. Such an intermediate was obtained via two pathways. The first pathway proceeded via a dissociative oxygen adsorption pathway on neighboring sulfur atoms and its further surface reorganization. Second pathway was a direct O2 adsorption on the same sulfur atom and its further surface reorganization. However, calculated activation energies for both pathways are at least 0.4 eV more than obtained via Santosh et al. [42][43] for dissociative oxygen adsorption onto pristine MoS2, which suggests that the trajectory proposed by Farigliano et al. [57][72] is less probable or that there are differences in the DFT implementations between these studies. Nevertheless, the results of Farigliano et al. provide an important link between the dissociative oxygen adsorption and oxidative etching. This is because, in the subsequent steps, Farigliano et al. [57][72] have found out that the OSOMo intermediate decomposes either via direct SO2 desorption to generate a single sulfur vacancy, or via SO desorption, leaving substitutional oxygen on the surface.

It has been established that single sulfur vacancies in the basal sulfur planes are the most frequent and important from the point of view of an initial oxygen attack [58][59][60][70,81,82], since their activation energies are much less than in the case of non-defected basal planes. Along this line of research Martincová et al. have studied oxidative etching directly on “simplified edges” of the MoS2 crystals[50][61] [46,58]. The initial state was taken to be the one in which the O2 molecule is in the edge vicinity. The final state of the dissociative splitting was taken to be the one with two separate oxygen atoms adsorbed on adjacent sulfur atoms at an edge. Several dissociation pathways were considered, differing mainly in an initial position of the O2 molecule. The most favorable pathway yielded a very low O2 dissociation barrier of only 0.31 eV and several other pathways yielded similar barriers, all not exceeding 0.5 eV. Such results still do not suggest immediate dissociative O2 splitting onto MoS2 edges; however, they clearly produce much lower Ea values than 0.8 eV obtained by Santosh et al. for the same process on a defective MoS2 basal plane. However, the link between dissociative oxygen adsorption and oxidative etching is not straightforward due to other various steps necessary.

Overall, there are several successes of the presented above theoretical approaches. First, chemical inertness for reactions with oxygen of the non-defective MoS2 basal planes at room temperature has been explained in terms of high activation barrier for a dissociative oxygen adsorption onto such planes with an energy barrier of at least 1.6 eV. Second, according to experimental results, oxygen-induced single sulfur vacancy (SSV) creation and its later oxidation have been shown to slowly introduce defects within the MoS2 flakes, even at room temperatures, thus making them prone to oxidation via decreasing the energy barrier down to 0.8 eV on SSV and to 0.3 eV on edges/grain boundaries. These results agree qualitatively with experimentally obtained activation energies for oxidative etching which range from 0.6 to ca. 1 eV for pristine MoS2 samples, where single oxidation events according to the transition state theory would occur on the time scales of months.

Last, but not least, a unified picture for oxidative etching is still elusive. Some models and theoretical simulations refer to dissociative oxygen adsorption on sulfur atoms as an initial limiting step, i.e., studies by Martincová et al. [62][46], as well as by Farigliano et al. [57][72], while others refer to direct reaction of oxygen with Mo atoms exposed due to omnipresent single sulfur vacancies, i.e., Lv et al.[58] [70], Santosh et al. [43] and Zhou et al. [33]. In addition, the role of initially physically adsorbed oxygen in thermal etching is still not clear [41], since its diffusion to the reaction centers can compete with direct reactions with atmospheric oxygen.

REFERENCES:

1. Kadantsev, E.S.; Hawrylak, P. Electronic structure of a single MoS2 monolayer. Solid State Commun. 2012, 152, 909–913. [CrossRef]

2. Tagawa, M.; Muromoto, M.; Hachiue, S.; Yokota, K.; Ohmae, N.; Matsumoto, K.; Suzuki, M. Hyperthermal atomic oxygen

interaction with MoS2 lubricants and relevance to space environmental effects in low earth orbit—Effects on friction coefficient

and wear-life. Tribol. Lett. 2005, 18, 437–443. [CrossRef]

3. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.; Strano, M.S. Electronics and optoelectronics of two-dimensional transition

metal dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. [CrossRef] [PubMed]

4. Fleischauer, P.D.; Lince, J. A comparison of oxidation and oxygen substitution in MoS2 solid film lubricants. Tribol. Int. 1999, 32,

627–636. [CrossRef]

5. Lince, J.; Frantz, P.P. Anisotropic oxidation of MoS2 crystallites studied by angle-resolved X-ray photoelectron spectroscopy.

Tribol. Lett. 2001, 9, 211–218. [CrossRef]

6. Imanishi, N.; Kanamura, K.; Takehara, Z. Synthesis of MoS2 thin film by chemical vapor deposition method and discharge

characteristics as a cathode of the lithium secondary battery. J. Electrochem. Soc. 1992, 139, 2082–2087. [CrossRef]

7. Jäger-Waldau, A.; Lux-Steiner, M.; Bucher, E. MoS2, MoSe2, WS2 and WSe2 thin films for photovoltaics. Solid State Phenom. 1994,

37–38, 479–484. [CrossRef]

8. Jaramillo, T.F.; Jorgensen, K.P.; Bonde, J.; Nilsen, J.H.; Horch, S.; Chorkendorff, I. Identification of active edge sites for electrochemical

H2 evolution from MoS2 nanocatalysts. Science 2007, 317, 100–102. [CrossRef]

9. Guo, J.; Li, F.; Sun, Y.; Zhang, X.; Tang, L. Oxygen-incorporated MoS2 ultrathin nanosheets grown on graphene for efficient

electrochemical hydrogen evolution. J. Power Sources 2015, 291, 195–200. [CrossRef]

10. Xie, J.; Zhang, J.; Li, S.; Grote, F.; Zhang, X.; Wang, R.; Lei, Y.; Pan, B.; Xie, Y. Controllable Disordered engineering in oxygenincorporated

MoS2 ultrathin nanosheets for efficient hydrogen evolution. J. Am. Chem. Soc. 2013, 135, 17881–17888. [CrossRef]

11. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150.

[CrossRef] [PubMed]

12. Radisavljevic, B.; Whitwick, M.B.; Kis, A. Integrated circuits and logic operations based on single-layer MoS2. ACS Nano 2011, 5,

9934–9938. [CrossRef] [PubMed]

13. Wachter, S.; Polyushkin, D.; Bethge, O.; Mueller, T. A microprocessor based on a two-dimensional semiconductor. Nat. Commun.

2017, 8, 14948. [CrossRef]

Materials 2021, 14, 5979 29 of 31

14. Lee, J.; Pak, S.; Giraud, P.; Lee, Y.-W.; Cho, Y.; Hong, J.; Jang, A.R.; Chung, H.-S.; Hong, W.-K.; Jeong, H.Y.; et al. Thermodynamically

stable synthesis of large-scale and highly crystalline transition metal dichalcogenide monolayers and their unipolar n-n

heterojunction devices. Adv. Mater. 2017, 29, 1702206. [CrossRef] [PubMed]

15. Tan, C.; Liu, Z.; Huang, W.; Zhang, H. Non-volatile resistive memory devices based on solution-processed ultrathin twodimensional

nanomaterials. Chem. Soc. Rev. 2015, 44, 2615–2628. [CrossRef]

16. Marega, G.M.; Zhao, Y.; Avsar, A.;Wang, Z.; Tripathi, M.; Radenovic, A.; Kis, A. Logic-in-memory based on an atomically thin

semiconductor. Nature 2020, 587, 72–77. [CrossRef]

17. Kang, K.; Xie, S.; Huang, L.; Han, Y.; Huang, P.Y.; Mak, K.F.; Kim, C.J.; Muller, D.; Park, J. High-mobility three-atom-thick

semiconducting films with wafer-scale homogeneity. Nature 2015, 520, 656–660. [CrossRef] [PubMed]

18. Chang, M.C.; Ho, P.H.; Tseng, M.F.; Lin, F.Y.; Hou, C.H.; Lin, I.K.; Wang, H.; Huang, P.P.; Chiang, C.H.; Yang, Y.C.; et al. Fast

growth of large-grain and continuous MoS2 films through a self-capping vapor-liquid-solid method. Nat. Commun. 2020, 11, 3682.

[CrossRef]

19. Shim, J.; Bae, S.-H.; Kong, W.; Lee, D.; Qiao, K.; Nezich, D.; Park, Y.J.; Zhao, R.; Sundaram, S.; Li, X.; et al. Controlled crack

propagation for atomic precision handling of wafer-scale two-dimensional materials. Science 2018, 362, 665–670. [CrossRef]

20. Kang, K.; Lee, K.H.; Han, Y.; Gao, H.; Xie, S.; Muller, D.A.; Park, J. Layer-by-layer assembly of two-dimensional materials into

wafer-scale heterostructures. Nature 2017, 550, 229–233. [CrossRef]

21. Amani, M.; Lien, D.H.; Kiriya, D.; Xiao, J.; Azcatl, A.; Noh, J.; Madhvapathy, S.R.; Addou, R.; Kc, S.; Dubey, M.; et al. Near-unity

photoluminescence quantum yield in MoS2. Science 2015, 350, 1065–1068. [CrossRef] [PubMed]

22. Gonzalez Marin, J.F.; Unuchek, D.;Watanabe, K.; Taniguchi, T.; Kis, A. MoS2 photodetectors integrated with photonic circuits.

NPJ 2D Mater. Appl. 2019, 3, 14. [CrossRef]

23. Gong, C.; Zhang, Y.; Chen, W.; Chu, J.; Lei, T.; Pu, J.; Dai, L.; Wu, C.; Cheng, Y.; Zhai, T.; et al. Electronic and optoelectronic

applications based on 2D novel anisotropic transition metal dichalcogenides. Adv. Sci. 2017, 4, 1700231. [CrossRef] [PubMed]

24. Unuchek, D.; Ciarrocchi, A.; Avsar, A.; Watanabe, K.; Taniguchi, T.; Kis, A. Room-temperature electrical control of exciton flux in

a van derWaals heterostructure. Nature 2018, 560, 340–344. [CrossRef] [PubMed]

25. Mak, K.F.; He, K.; Shan, J.; Heinz, T.F. Control of valley polarization in monolayer MoS2 by optical helicity. Nat. Nanotechnol. 2012,

7, 494–498. [CrossRef]

26. Yalon, E.; McClellan, C.J.; Smithe, K.K.H.; Muñoz Rojo, M.; Xu, R.L.; Suryavanshi, S.V.; Gabourie, A.; Neumann, C.M.; Xiong, F.;

Farimani, A.B.; et al. Energy dissipation in monolayer MoS2 electronics. Nano Lett. 2017, 17, 3429–3433. [CrossRef]

27. Qiu, H.; Pan, L.; Yao, Z.; Li, J.; Shi, Y.; Wang, X.; Park, W.; Park, J.; Jang, J.; Lee, H.; et al. Electrical characterization of back-gated

bi-layer MoS2 field-effect transistors and the effect of ambient on their performances. Appl. Phys. Lett. 2012, 100, 123104.

[CrossRef]

28. Park, W.; Park, J.; Jang, J.; Lee, H.; Jeong, H.; Cho, K.; Hong, S.; Lee, T. Oxygen environmental and passivation effects on

molybdenum disulfide field effect transistors. Nanotechnology 2013, 24, 095202. [CrossRef]

29. Li, T.-D.; Gao, J.; Szoszkiewicz, R.; Landman, U.; Riedo, E. Structured and viscous water in subnanometer gaps. Phys. Rev. B 2007,

75, 115415. [CrossRef]

30. Szoszkiewicz, R.; Riedo, E. Nucleation time of nanoscale water bridges. Phys. Rev. Lett. 2005, 95, 135502. [CrossRef]

31. Ross, S. Surface oxidation of molybdenum disulfide. J. Phys. Chem. 1955, 59, 889–892. [CrossRef]

32. Ebrahimi Kahrizsangi, R.; Abbasi, M.H.; Saidi, A. Model-Fitting Approach to Kinetic Analysis of Non-Isothermal Oxidation of

Molybdenite. Iran. J. Chem. Chem. Eng. 2007, 26, 119–123.

33. Zhou, H.; Yu, F.; Liu, Y.; Zou, X.; Cong, C.; Qiu, C.; Yu, T.; Yan, Z.; Shen, X.; Sun, L.; et al. Thickness-dependent patterning of

MoS2 sheets with well-oriented triangular pits by heating in air. Nano Res. 2013, 6, 703–711. [CrossRef]

34. Wu, J.; Li, H.; Yin, Z.; Li, H.; Liu, J.; Cao, X.; Zhang, Q.; Zhang, H. Layer thinning and etching of mechanically exfoliated MoS2

nanosheets by thermal annealing in air. Small 2013, 9, 3314–3319. [CrossRef]

35. Yamamoto, M.; Einstein, T.L.; Fuhrer, M.S.; Cullen, W.G. Anisotropic etching of atomically thin MoS2. J. Phys. Chem. C 2013, 117,

25643–25649. [CrossRef]

36. Walter, T.N.; Kwok, F.; Simchi, H.; Aldosari, H.M.; Mohney, S.E. Oxidation and oxidative vapor-phase etching of few-layer MoS2.

J. Vac. Sci. Technol. B 2017, 35, 21203. [CrossRef]

37. Rao, R.; E Islam, A.; Campbell, P.M.; Vogel, E.M.; Maruyama, B. In situ thermal oxidation kinetics in few layer MoS2. 2D Mater.

2017, 4, 025058. [CrossRef]

38. Tang, J.; Wei, Z.; Wang, Q.; Wang, Y.; Han, B.; Li, X.; Huang, B.; Liao, M.; Liu, J.; Li, N.; et al. In situ oxygen doping of monolayer

MoS2 for novel electronics. Small 2020, 16, 2–9. [CrossRef]

39. Wang, G.; Pandey, R.; Karna, S.P. Physics and chemistry of oxidation of two-dimensional nanomaterials by molecular oxygen.

Wiley Interdiscip. Rev. Comput. Mol. Sci. 2017, 7, e1280. [CrossRef]

40. Ukegbu, U.; Szoszkiewicz, R. Microscopic kinetics of heat-induced oxidative etching of thick MoS2 crystals. J. Phys. Chem. C 2019,

123, 22123–22129. [CrossRef]

41. Spychalski, W.L.; Pisarek, M.; Szoszkiewicz, R. Microscale Insight into Oxidation of Single MoS2 Crystals in Air. J. Phys. Chem. C

2017, 121, 26027–26033. [CrossRef]

42. Szoszkiewicz, R.; Rogala, M.; Dabrowski, P. Surface-bound and volatile mo oxides produced during oxidation of single MoS2

crystals in air and high relative humidity. Materials 2020, 13, 3067. [CrossRef] [PubMed]

Materials 2021, 14, 5979 30 of 31

43. KC, S.; Longo, R.C.; Wallace, R.M.; Cho, K. Surface oxidation energetics and kinetics on MoS2 monolayer. J. Appl. Phys. 2015,

117, 135301. [CrossRef]

44. Krawczyk, M.; Pisarek, M.; Szoszkiewicz, R.; Jablonski, A. Surface characterization of MoS2 atomic layers mechanically exfoliated

on a Si substrate. Materials 2020, 13, 3595. [CrossRef] [PubMed]

45. Jia, F.; Liu, C.; Yang, B.; Song, S. Microscale control of edge defect and oxidation on molybdenum disulfide through thermal

treatment in air and nitrogen atmospheres. Appl. Surf. Sci. 2018, 462, 471–479. [CrossRef]

46. Martincova, J.; Otyepka, M.; Lazar, P. Oxidation of metallic two-dimensional transition metal dichalcogenides: 1T-MoS2 and

1T-TaS2. 2D Mater. 2020, 7, 045005. [CrossRef]

47. Park, S.; Garcia-Esparza, A.T.; Abroshan, H.; Abraham, B.; Vinson, J.; Gallo, A.; Nordlund, D.; Park, J.; Kim, T.R.; Vallez, L.; et al.

Operando study of thermal oxidation of monolayer MoS2. Adv. Sci. 2021, 8, 2002768. [CrossRef]

48. Rogala, M.; Sokołowski, S.; Ukegbu, U.; Mierzwantos, A.; Szoszkiewicz, R. Direct identification of surface bound MoO3 on single

MoS2 flakes heated etin dry and humid air. Adv. Mater. Interfaces 2021, 2100328. [CrossRef]

49. Bal.llou, E.V.; [42]Ross, S. The adsorption of benzene and water vapors by molybdenum disulfide. J. Phys. Chem. 1953, 57, 653–657.

[CrossRef]

50. De Castro, I.A.; Datta, R.S.; Ou, J.Z.; Castellanos-Gomez, A.; Sriram, S.; Daeneke, T.; Kalan-tar-Zadehou e, K. Molybdenum oxides—

From fundamentals al. [to functionality. Adv. Mater. 2017, 29, 1–31. [CrossRef]

51. Åsbrink, S.; Kihlborg, L.; Malinowski, M. High-pressure single-crystal X-ray diffraction studies of MoO3. I. Lattice parameters up

to 74 GPa. J. Appl. Crystallogr. 1988, 21, 960–962. [CrossRef]

52. Bihn, J.H.; Park, J.; Kang, Y.C. Synthesis and characterization of mo films deposited by RF sputtering at various oxygen ratios. J.

Korean Phys. Soc. 2011, 58, 509–514. [CrossRef]

53. Smolik, G.R.; Petti, D.A.; Schuetz, S.T. Oxidation and volatilization of TZM alloy in air. J. Nucl. Mater. 2000, 283-287, 1458–1462.

[CrossRef][32]

54. Mirabelli, G.; McGeough, C.; Schmidt, M.; McCarthy, E.K.; Monaghan, S.; Povey, I.M.; McCarthy, M.M.; Gity, F.; Nagle, R.E.;

Hughes, G.; et al. Air sensitivity of MoS2, MoSe2, MoTe2, HfS2, and HfSe2. J. Appl. Phys. 2016, 120, 125102. [CrossRef]

55. Jo, S.S.; Singh, A.; Yang, L.; Tiwari, S.C.; Hong, S.; Krishnamoorthy, A.; Sales, M.G.; Oliver, S.M.; Fox, J.; Cavalero, R.L.; et al.

Growth kinetics and atomistic mechanisms of native oxidation of ZrSxSe2–x and MoS2 crystals. Nano Lett. 2020, 20, 8592–8599.

[CrossRef] [PubMed]

56. Gao, J.; Li, B.; Tan, J.; Chow, P.; Lu, T.-M.; Koratkar, N. Aging of transition metal dichalcogenide monolayers. ACS Nano 2016, 10,

2628–2635. [CrossRef]

57. Longo, R.C.; Addou, R.; Santosh, K.C.; Noh, J.Y.; Smyth, C.M.; Barrera, D.; Zhang, C.; Hsu, J.W.P.; Wallace, R.M.; Cho, K. Intrinsic

air stadditionbility mechanisms of two-dimensional transition metal dichalcogenide surfaces: Basal versus edge oxidation. 2D Mater.

2017, 4, 025050. [CrossRef]

58. Martincová, J.; Otyepka, M.; Lazar, P. Is single layer MoS2 stable in the roair? Chem. A Eur. J. 2017, 23, 13233–13239. [CrossRef]

59. Sen, H.S.; Sahin, H.; Peeters, F.M.; Durgun, E. Monolaye of initiallyrs of MoS2 as an oxidation protective nanocoating material. J. Appl. Phys.

2014, 116, 083508. [CrossRef]

60. Pet˝o, J.; Ollár, T.; Vancsó, P.; Popov, Z.I.; Magda, G.Z.; Dobrik, G.; Hwang, C.; Sorokin, P.B.; Tapasztó, L. Spontaneous doping of

the basal plane of MoS2 single laysicaers through oxygen substitution under ambient conditions. Nat. Chem. 2018, 10, 1246–1251.

[CrossRef]

61. Wang, Z.; Bussche, A.V.D.; Qiu, Y.; Valentin, T.M.; Gion, K.; Kane, A.B.; Hurt, R.H. Chemicaly dissolution pathways of MoS2

nanosheets in biological ands environmental media. Environ. Sci. Technol. 2016, 50, 7208–7217. [CrossRef]

62. Zhang, X.; Jia, F.; Yang, B.; Song, S. Oxidation of molybdenum disulfide sheet in water under in situ atomic force microscopy

observation. J. Phys. Chem. C 2017, 121, 9938–9943. [CrossRef]

63. Seguin, L.; Figlarz, M.; Cavagnat, R.; Lassègues, J.C. Infrared and Raman spectra oxf MoO3 molybdenum trioxides and MoO3

xH2O molybdenum trioxide hydrates. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 1995, 51, 1323–1344. [CrossRef]

64. Cruywagen, in theJ.J.; Heyns, J.B.B. Solubility of yellow molybdenum(VI) oxide dihydrate (Mo03.2H20) in 3,0M-sodium perchlorate at

25 C. S. Afr. J. Chem. 1981, 34, 118–120.

65. Oyerinde, O.F.; Weeks, C.L.; Anbar, A.D.; Spiro, T.G. Sol etcution structure of molybdic acid from Raman spectroscopy and DFT

analysis. Inorganica Chim. Acta 2008, 361, 1000–1007. [CrossRef]

66. Piquemal, J.-Y.; Manoli, J.-M.; Beaunier, P.; Ensuque, A.; Toug isne, P.; Legrand, A.-P.; Brégeault, J.-M. Using inorganic silicate

precursor/molybdenum peroxo complexestil/onium salt interfaces in aqueous acidic media to design mesoporous silica with high

molybdenum not content and high dispersion. Microporous Mesoporous Mater. 1999, 29, 291–304. [CrossRef]

67. Espinosa, F.M.; Ryu, Y.K.; Marinov, K.; Dumcenco, D.; Kis, A.; Garcia, R. Direct fabrication of thin leaayer MoS2 field-effect nanoscale

tr[40]ansistors by oxidation scanning probe lithography. Appl. Phys. Lett. 2015, 106, 103503. [CrossRef]

68. Fang, L.; Shu, Y.; Wang, A.A.; Zhang, T. Green synthesis and characterization of anisotropic uniform single-crystal  -MoO3

nanostructures. J. Phys. Chem. C 2007, 111, 2401–2408. [CrossRef]

69. Addou, R.; McDonnell, S.; Barrera, D.; Guo, Z.; Azcatl, A.;Wang, J.; Zhu, H.; Hinkle, C.L.; Quevedo-Lopez, M.; Alshareef, H.N.;

et al. Impurities and electronic property variations of natural MoS2 crystal surfaces. ACS Nano 2015, 9, 9124–9133. [CrossRef]

70. Lv, D.;Wang, H.; Zhu, D.; Lin, J.; Yin, G.; Lin, F.; Zhang, Z.; Jin, C. Atomic process of oxidative etching in monolayer molybdenum

disulfide. Sci. Bull. 2017, 62, 846–851. [CrossRef]

Materials 2021, 14, 5979 31 of 31

71. Chakraborty, B.; Bera, A.; Muthu, D.V.S.; Bhowmick, S.; Waghmare, U.V.; Sood, A.K. Symmetry-dependent phonon renormalization

in monolayer MoS2 transistor. Phys. Rev. B 2012, 85, 2–5. [CrossRef]

72. Farigliano, L.M.; Paredes-Olivera, P.A.; Patrito, E.M. Initial steps of oxidative etching of MoS2 basal plane induced by O2. J. Phys.

Chem. C 2020, 124, 13177–13186. [CrossRef]

73. Huang, Y.; Wu, J.; Xu, X.; Ho, Y.; Ni, G.; Zou, Q.; Koon, G.K.W.; Zhao, W.; Casintro Neto, A.H.; Eda, G.; et al. An innovative way of

etching MoS2: Characte its rization and mechanistic investigation. Nano Res. 2013, 6, 200–207. [CrossRef]

74. Maguire, P.; Jadwifszczak, J.; O’Brien, M.; Keane, D.; Duesberg, G.S.; McEvoy, N.; Zhang, H. Defect-moderated oxidative etching

of MoS2. J. Appl. Phys. 2019, 126, 164301. [CrossRef]

75. Castellanos-Gomez, A.; Barkelid, M.; Goossens, A.M.; Calado, V.E.; van der Zant, H.S.J.; Steele, G.A. Laser-thinning of MoS2: On

demand generation of a single-layer semiconductor. Nano Lett. 2012, 12, 3187–3192. [CrossRef] [PubMed]

76. Sunamura, K.; Page, T.R.; Yoshida, K.; Yanon to the, T.A.; Hayamizu, Y. Laser-induced electrochemical thinning of MoS2. J. Mater. Chem.

C 2016, 4, 3268–3273. [CreossRef]

77. Ukegbu, U.K. Microscale Etction centhing/Oxidation of Thick MoS2 Flakes. Master’s Thesis, University of Warsaw, Warsaw, Poland, 2018.

78. Henkelman, G.; Ubers can couaga, B.P.; Jónsson, H. A climbing image nudged elastic band method for finding saddle points and

minimum energy peaths. J. Chem. Phys. 2000, 113, 9901–9904. [CrossRef]

79. Nan, H.; Wang, Z.; Wang, W.; Liang, Z.; Lu, Y.; Chen, Q.; He, D.; Tan, P.; Miao, F.; Wang, X.; et al. Strong photoluminescence

enhancement of MoS2 through defect engineering and oxygen bonding. ACS Nano 2014, 8, 5738–5745. [CrossRef] [PubMed]

80. Grønborg, S.S.; Thorarinsdottir, K.; Kyhl, L.; Rodríguez-Fernández, J.; E Sanders, C.; Bianchi, M.; Hofmann, P.; Miwa, J.; Ulstrup,

S.; Lauritsen, J.V. Basal plane oxygen exch direange of epitaxial MoS2 without edge oxidation. 2D Mater. 2019, 6, 045013. [CrossRef]

81. Hong, J.; Hu, Z.; Probert, M.; Li, K.; Lv, D.; Yang, X.; Gu, L.; Mao, N.; Feng, Q.; Xie, L.; et al. Exploring atomic defect s in

molybdenum disulphide monolayers. Nat. Commun. 2015, 6, 6293. [CrossRef]

82. Zhou, W.; Zou, X.; Najmaei, S.; Liu, Z.; Shi, Y.; Kong, J.; Lou, J.; Ajayan, P.M.; Yakobson, B.I.; Idrobo, J.C. Intrinsict structural

defects in mons olayer molybdenum disulfide. Nano Lett. 2013, 13, 2615–2622. [CrossRef]

83. Ko, T.Y.; Jeong, A.; Kim, W.; Lee, J.; Kim, Y.; Lee, J.E.; Ryu, G.H.; Park, K.; Kim, D.; Lee, Z.; et al. On-stack two-dimensional

conversion of MoS2 into MoO3. 2D Mater. 2016, 4, 014003. [CrossRef]

84. Kuzmin, A.; Purans, J. Dehydration of the molybdenum trioxide athydrates MoO3nH2O: In situ x-ray absorption spectroscopy

study at the Mo K edge. J. Phys. Condens. Matter 2000, 12, 1959–1970. [CrossRef]

85. Rice, R.H.; Mokarian-Tabari, P.; King,W.P.; Szoszkiewicz, R. Local thermomechanical analysispheri of a microphase-separated thin

lamellar PS-b-PEO film. Langmuir 2012, 28, 13503–13511. [CrossRef] [PubMed]

86. Zhu, H.; Qin, X.; Cheng, L.; Azcatl, oxygA.; Kim, J.;Wallace, R.M. Remote plasma oxidation and atomic layer etching of MoS2. ACS

Appl. Mater. Interfaces 2016, 8, 19119–19126. [CrossRef] [PubMed]

87. Li, H.; Qi, X.;Wu, J.; Zeng, Z.;Wei, J.; Zhang, H. Investigation of MoS2 and graphene nanosheets by magnetic force microscopy.

ACS Nano 2013, 7, 2842–2849. [CrossRef]

88. Lavini, F.; Calò, A.; Gao, Y.; Albisetti, E.; Li, T.D.; Cao, T.; Li, G.; Cao, L.; Aruta, C.; Riedo, E. Friction and work function oscillatory

behavior for an even and odd number of layers in polycrystalline MoS2. Nanoscale 2018, 10, 8304–8312. [CrossRef]

89. Kim, C.; Moon, I.; Lee, D.; Choi, M.S.; Ahmed, F.; Nam, S.; Cho, Y.; Shin, H.-J.; Park, S.; Yoo, W.J. Fermi level pinning at electrical

metal contacts of monolayer molybdenum dichalcogenides. ACS Nano 2017, 11, 1588–1596. [CrossRef]

90. Choi, M.S.; Qu, D.; Lee, D.; Liu, X.; Watanabe, K.; Taniguchi, T.; Yoo, W.J. Lateral MoS2 p–n junction formed by chemical doping

for use in high-performance optoelectronics. ACS Nano 2014, 8, 9332–9340. [CrossRef]

91. Inzani, K.; Nematollahi, M.; Vullum-Bruer, F.; Grande, T.; Reenaas, T.W.; Selbach, S.M. Electronic properties of reduced

molybdenum oxides. Phys. Chem. Chem. Phys. 2017, 19, 9232–9245. [CrossRef]

92. Chuang, S.; Battaglia, C.; Azcatl, A.; McDonnell, S.; Kang, J.S.; Yin, X.; Tosun, M.; Kapadia, R.; Fang, H.; Wallace, R.M.; et al.

MoS2 P-type transistors and diodes enabled by high work function MoOx contacts. Nano Lett. 2014, 14, 1337–1342. [CrossRef]

[PubMed]

93. Xing, K.; Xiang, Y.; Jiang, M.; Creedon, D.L.; Akhgar, G.; Yianni, S.A.; Xiao, H.; Ley, L.; Stacey, A.; McCallum, J.C.; et al. MoO3

induces p-type surface conductivity by surface transfer doping in diamond. Appl. Surf. Sci. 2020, 509, 144890. [CrossRef]

94. Kessler, M.A.; Ohrdes, T.; Wolpensinger, B.; Harder, N.-P. Charge carrier lifetime degradation in Cz silicon through the formation

of a boron-rich layer during BBr3diffusion processes. Semicond. Sci. Technol. 2010, 25, 055001. [CrossRef]

95. Szoszkiewicz, R.; Okada, T.; Jones, S.C.; Li, T.D.; King, W.P.; Marder, S.R.; Riedo, E. high-speed, sub-15 nm Feature size

thermochemical nanolithography. Nano Lett. 2007, 7, 1064–1069. [CrossRef] [PubMed]

96. Martínez, R.V.; Martínez, J.; Garcia, R. Silicon nanowire circuits fabricated by AFM oxidation nanolithography. Nanotechnology

2010, 21, 245301. [CrossRef] [PubMed]

97. Chen, L.; Wen, J.; Zhang, P.; Yu, B.; Chen, C.; Ma, T.; Lu, X.; Kim, S.H.; Qian, L. Nanomanufacturing of silicon surface with a

single atomic layer precision via mechanochemical reactions. Nat. Commun. 2018, 9, 1542. [CrossRef]

98. Garcia, R.; Knoll, A.; Riedo, E. Advanced scanning probe lithography. Nat. Nanotechnol. 2014, 9, 577–587. [CrossRef]