Next Article in Journal
Muon Irradiation of ZnO Rods: Superparamagnetic Nature Induced by Defects
Next Article in Special Issue
Boxcar Averaging Scanning Nonlinear Dielectric Microscopy
Previous Article in Journal
Incorporation of Superparamagnetic Iron Oxide Nanoparticles into Collagen Formulation for 3D Electrospun Scaffolds
Previous Article in Special Issue
Highly Homogeneous Current Transport in Ultra-Thin Aluminum Nitride (AlN) Epitaxial Films on Gallium Nitride (GaN) Deposited by Plasma Enhanced Atomic Layer Deposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multiscale Investigation of the Structural, Electrical and Photoluminescence Properties of MoS2 Obtained by MoO3 Sulfurization

1
Consiglio Nazionale delle Ricerche—Istituto per la Microelettronica e Microsistemi (CNR-IMM), Strada VIII 5, 95121 Catania, Italy
2
Department of Physics and Astronomy, University of Catania, 95123 Catania, Italy
3
Centre for Energy Research, Institute of Technical Physics and Materials Science, Konkoly-Thege ut 29-33, 1121 Budapest, Hungary
4
Department of Physics and Chemistry Emilio Segrè, University of Palermo, 90123 Palermo, Italy
5
ATEN Center, University of Palermo, 90123 Palermo, Italy
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(2), 182; https://doi.org/10.3390/nano12020182
Submission received: 3 December 2021 / Revised: 1 January 2022 / Accepted: 3 January 2022 / Published: 6 January 2022
(This article belongs to the Special Issue Nanotechnology for Electronic Materials and Devices)

Abstract

:
In this paper, we report a multiscale investigation of the compositional, morphological, structural, electrical, and optical emission properties of 2H-MoS2 obtained by sulfurization at 800 °C of very thin MoO3 films (with thickness ranging from ~2.8 nm to ~4.2 nm) on a SiO2/Si substrate. XPS analyses confirmed that the sulfurization was very effective in the reduction of the oxide to MoS2, with only a small percentage of residual MoO3 present in the final film. High-resolution TEM/STEM analyses revealed the formation of few (i.e., 2–3 layers) of MoS2 nearly aligned with the SiO2 surface in the case of the thinnest (~2.8 nm) MoO3 film, whereas multilayers of MoS2 partially standing up with respect to the substrate were observed for the ~4.2 nm one. Such different configurations indicate the prevalence of different mechanisms (i.e., vapour-solid surface reaction or S diffusion within the film) as a function of the thickness. The uniform thickness distribution of the few-layer and multilayer MoS2 was confirmed by Raman mapping. Furthermore, the correlative plot of the characteristic A1g-E2g Raman modes revealed a compressive strain (ε ≈ −0.78 ± 0.18%) and the coexistence of n- and p-type doped areas in the few-layer MoS2 on SiO2, where the p-type doping is probably due to the presence of residual MoO3. Nanoscale resolution current mapping by C-AFM showed local inhomogeneities in the conductivity of the few-layer MoS2, which are well correlated to the lateral changes in the strain detected by Raman. Finally, characteristic spectroscopic signatures of the defects/disorder in MoS2 films produced by sulfurization were identified by a comparative analysis of Raman and photoluminescence (PL) spectra with CVD grown MoS2 flakes.

1. Introduction

Transition metal dichalcogenides (TMDs) are a wide family of layered van der Waals (vdW) materials with the general chemical formula MX2, M being a transition metal (Ti, Zr, Hf, V, Nb, Ta, Mo, W, Re, Pd, or Pt) and X a chalcogen atom (S, Se, or Te) [1]. Most of them exhibit metallic or semiconducting phases. In particular, semiconducting TMDs have been the object of increasing scientific interest in the last decade, due to their huge potential for applications in several fields, including electronics, optoelectronics, spintronics, valleytronics, chemical/environmental sensing, energy generation, and catalysis [2,3,4,5,6,7,8,9,10]. Molybdenum disulfide (MoS2) is the most investigated among TMDs, due to the natural abundance and good chemical/mechanical stability of its 2H semiconductor phase under ambient conditions. The bandgap tunability as a function of the thickness, with a transition from an indirect bandgap of ~1.2 eV for bulk or few-layer MoS2 to a direct bandgap of ~1.8 eV for monolayer MoS2 [11,12], makes this material appealing for optoelectronic and electronic applications. In fact, the first robust 2D transistor with a large on/off ratio and good field-effect mobility was demonstrated using monolayer 2H-MoS2 flakes as the semiconducting channel [13,14]. This material and other TMDs are currently considered a potential replacement of Si for the next generation of complementary metal oxide semiconductor (CMOS) devices allowing the continuation of Moore’s law [15]. Furthermore, they can represent the basis for new concept (More-than-Moore) devices [16,17].
Due to this wide application potential, scalable and reproducible growth methods for thin films of TMDs are strongly required for their future implementation in manufacturing lines. In this context, research on MoS2 wafer-scale growth and device integration is relatively more mature than for other 2D TMDs.
Top-down synthesis approaches used to separate MoS2 from bulk crystals, such as mechanical exfoliation [18,19], gold-assisted exfoliation [20,21,22,23,24], and liquid exfoliation [25], are not suitable to ensure the reproducibility and thickness control on a wafer scale required for high-end electronic applications. For this reason, bottom-up approaches as Chemical Vapour Deposition (CVD) [26,27], Pulsed Laser Deposition (PLD) [28], Molecular Beam Epitaxy (MBE) [29], and Atomic Layer Deposition (ALD) [30] represent the most promising methods to obtain a reproducible thin film of TMDs on a large area.
In particular, CVD using vapours from S and MoO3 powders has been widely explored by several research groups, since it is a cost-effective method to produce MoS2 domains with good crystalline quality on different substrates [31,32,33]. Although monolayer flakes with a triangular or hexagonal shape and lateral extension from tens to hundreds of micrometres have been obtained under optimized CVD conditions [34], achieving coverage and thickness uniformity on the wafer scale still represents a huge challenge, due to the difficulty of controlling all the parameters involved in the process (including the substrate temperature, the evaporation rates of the S and Mo precursors, the pressure in the chamber, and the carrier gas flow rate) [35,36,37,38,39].
As an alternative to the single-step CVD approach, sulfurization of a Mo (or Mo-oxide) film pre-deposited on a substrate (e.g., by evaporation or sputtering) allows superior control of MoS2 coverage and uniformity by controlling the initial film thickness [40,41,42,43]. Different to CVD (where the Mo–S bonds are mostly formed by vapour phase reaction and the MoS2 lands on the substrate), the sulfurization process is a heterogeneous vapour-solid reaction between the S vapour and the pre-deposited film [44]. The conversion of MoOx to MoS2 by sulfurization has been demonstrated to occur in a wide temperature range, from 500 °C to 1000 °C, although the best quality films are typically obtained at temperatures > 750 °C [44]. Besides the vapour-solid surface reaction, the initial Mo or MoOx film thickness also plays an important role in the process. In fact, with increasing its thickness, the diffusion of S in the film represents the limiting mechanism for the formation of MoS2 layers and determines their alignment with respect to the substrate [45,46]. In particular, at typical sulfurization temperatures of 750–800 °C, single or few-layers of MoS2 horizontally aligned to the substrate plane are obtained for very thin (<3 nm) Mo films, whereas vertically aligned growth occurs for thicker Mo films [47]. This is due to the favoured sulphur diffusion along the vdW gaps between the vertically oriented MoS2 layers [45,47,48]. Besides the initial Mo (or Mo-oxide) thickness, other key factors controlling MoS2 formation include the substrate heating rate, pressure, and local S concentration on the sample surface [49,50,51]. Furthermore, the underlying substrate can play an important role in MoS2 formation during sulfurization of pre-deposited MoO3. In fact, while a higher temperature may enhance the sulfurization degree, on the other hand, it can also result in increased MoO3 evaporation and diffusion of Mo atoms on the substrate surface. This latter phenomenon strongly depends on the adhesion energy and surface diffusivity of Mo atoms on the substrate.
The main disadvantage of the continuous MoS2 films produced by the sulfurization approach is their nanocrystalline structure (with 20–30 nm grain-size) [44], typically resulting in poorer carrier mobility, if compared to the large and isolated monocrystalline MoS2 flakes obtained by the CVD approach. However, the high uniformity and its good compatibility with the fabrication methods used in the semiconductor industry makes this approach appealing for some applications, e.g., MoS2/semiconductor heterojunctions [52] or hydrogen evolution applications [53]. Hence, a detailed characterisation of structural/compositional, vibrational, optical, and electrical properties of MoS2 films produced by Mo sulfurization remains highly desirable.
In this paper, few or multilayer MoS2 on a SiO2/Si substrate have been produced by sulfurization at 800 °C of very thin MoO3 films, from ~2.8 nm to ~4.2 nm (i.e., the critical range for the transition from horizontally to vertically aligned layers). The compositional, morphological, structural, electrical, and optical emission properties of the grown films have been extensively investigated by the combination of several characterisation techniques with macro to nanoscale spatial resolution. This correlative analysis provides deep insight into the potentialities and limitations of this material system for applications.

2. Materials and Methods

The thin molybdenum-oxide films on SiO2 (900 nm)/Si substrates were obtained by DC magnetron sputtering from a Mo-target (using a Quorum Q300-TD system), followed by natural oxidation in air. The sulfurization process, schematically illustrated in Figure 1, was carried out in a two-heating zones furnace (TSH12/38/500, Elite Thermal Systems Ltd., Market Harborough, UK), with the first zone (at a temperature of 150 °C) hosting a crucible with 300 mg sulphur (purity 99.9%, product 28260.234, VWR Chemicals, Radnor, PA, USA), and the second zone (at a temperature of 800 °C) hosting the MoO3/SiO2/Si sample. Starting from a base pressure of 4 × 10−6 bar, the Ar carrier gas (purity 5.0, Messer, Budapest, Hungary) with a flux of 100 sccm transported the S vapours from the first to the second zone. The duration of the sulfurization process was 60 min.
Morphological analyses on the as-deposited MoO3 films and after the sulfurization process were carried out by Tapping mode Atomic Force Microscopy using a DI3100 system by Bruker (Santa Barbara, CA, USA) with Nanoscope V electronics. The compositional properties of the as-deposited metal films and MoS2 formation after the sulfurization process were evaluated by X-ray photoelectron spectroscopy (XPS) using Escalab Xi+ equipment by Thermo Fisher (Waltham, MA, USA), with a monochromatic Al Kα X-ray source (energy = 1486.6 eV). The spectra were collected at a take-off angle of 90° relative to the sample surface and pass energy of 20 eV. The instrument resolution was 0.45 eV (FWHM of the Ag 3d5/2 peak). The spectra were aligned using C1s (285 eV) as reference. High-resolution transmission electron microscopy (HR-TEM), high angle annular dark-field scanning transmission electron microscopy (HAADF-STEM), and energy dispersion spectroscopy (EDS) analyses of the MoS2 thin films were carried out with an aberration-corrected Titan Themis 200 microscope by Thermo Fisher (Waltham, MA USA). To this aim, cross-sectioned samples were prepared by a focused ion beam (FIB). Raman spectroscopy and mapping of MoS2 vibrational peaks were carried out by WiTec Alpha equipment by WiTec (Ulm, Germany), using laser excitation at 532 nm, 1.5 mW power, and 100× objective. Photoluminescence spectra (PL) were collected using a Horiba (Palaiseau, France) system with a laser source of 532 nm. To confirm the uniformity of the MoS2 thin layer across the substrate, the Raman and PL analyses have been performed at different positions on the sample. Finally, nanoscale resolution current mapping of MoS2 on SiO2 was performed by conductive Atomic Force Microscopy (C-AFM) with a DI3100 system by Bruker (Santa Barbara, CA, USA), using Pt-coated Si tips with ~5 nm curvature radius.

3. Results and Discussion

Figure 2a shows a typical AFM morphology of as-deposited MoO3 on the SiO2/Si substrate using the lowest sputtering time (30 s). This analysis indicates a very low root mean square (RMS) surface roughness of 0.35 nm. Similar roughness values have been measured for MoO3 film thicknesses deposited at higher sputtering times. The thickness of the as-deposited films was also evaluated by AFM step height measurements performed on intentionally scratched regions of the films. Figure 2b,c show the morphologies and corresponding line profiles for films deposited with two different sputtering times (30 s and 45 s), resulting in ~2.8 nm and ~4.2 nm thickness, respectively.
XPS compositional analyses performed on the thinnest deposited films revealed that they are predominantly composed of MoO3, with a small (<1%) MoO2 contribution. Recently, Vangelista et al. [44] also reported the complete oxidation (ascribed to air exposure after the deposition) of evaporated Mo films with similar thickness, used for subsequent MoS2 growth by sulfurization. The same authors [44] explained the conversion of MoO3 to MoS2 upon exposure to sulphur according to the following chemical reaction:
2 MoO3(s) + 7 S(g) → 2 MoS2(s) + 3 SO2(g),
which is the result of two intermediate steps:
MoO3 +(x/2) S MoO3−x + (x/2) SO2
MoO3−x + [(7 − x)/2] S MoS2 + [(3 − x)/2] SO2
i.e., the S-induced reduction of the MoO3 to a sub-stoichiometric oxide MoO3−x (2), followed by its conversion to MoS2 (3), with the formation of gaseous SO2 as a by-product.
After the sulfurization process at 800 °C, XPS analyses were performed to evaluate the successful conversion of MoO3 to MoS2. Figure 3a reports an overview spectrum, allowing the quantification of the percentage of elemental concentrations on the sample surface. In particular, molybdenum and sulphur percentages of 3.26% and 6.82%, respectively, were evaluated (besides the large Si and O background), which were close to the stoichiometric [Mo]/[S] ratio for MoS2. More detailed information on the Mo and S bonding was deduced from the Mo3d3/2, Mo3d5/2, and S2s core levels in Figure 3b, and the S2p1/2 and S2p3/2 core levels in Figure 3c. Two doublets were found in the Mo 3d spectrum, and both doublets were fitted with a peak separation of 3.1 eV [44,54,55]. In particular, the deconvolution of the Mo3d peaks shows the predominance of the Mo4+ component, associated with 2H-MoS2, accompanied by a smaller Mo6+ contribution, associated with the presence of residual MoO3. The two S2p1/2 and S2p3/2 peaks [44,54,55] in Figure 3c confirm that sulphur is mainly in the form of sulphide, with a small S-O component.
The structural properties of the MoS2 films were also investigated at nanoscale by transmission electron microscopy on cross-sectioned samples. Figure 4a,b show representative HR-TEM and HAADF-STEM analyses on the few-layers MoS2 sample obtained by sulfurization of the ~2.8 nm MoO3 film. The diffraction contrast in the HR-TEM image Figure 4a demonstrates the presence of two or three crystalline layers embedded between the amorphous SiO2 substrate and amorphous carbon (a–c) protective film. These layers are predominantly oriented parallel to the substrate, with nanometric scale corrugations. Furthermore, an interlayer spacing of ~0.6 nm is directly evaluated from the HRTEM image of a 3L-MoS2 reported in the insert of Figure 4a. The number of MoS2 layers and their nearly parallel orientation with respect to the substrate is confirmed by the HAADF-STEM image in Figure 4b collected on the same sample. On the other hand, a more irregular configuration of the layers can be observed from the HRTEM (Figure 4c) and HAADF-STEM (Figure 4d) analyses performed on the MoS2 multilayer produced by sulfurization of ~4.2 nm film. In fact, in the analysed specimen volume, horizontally oriented MoS2 layers co-exist with layers standing up with respect to the SiO2 surface. This observation is fully consistent with previous reports showing a transition from horizontal to vertically oriented growth for film thickness larger than 3 nm [47].
The layers number uniformity of the grown MoS2 films was also investigated on micro-meter scale areas and with high statistics by Raman spectroscopy. Figure 5 shows two typical Raman spectra of the few-layers (i.e., 2 L–3 L) MoS2 (black line) and of the multilayer MoS2 (red line) grown on SiO2 by the sulfurization process. The two characteristic in-plane (E2g) and out-of-plane (A1g) vibrational modes of MoS2 are clearly identified, and the typical redshift of the E2g peak and blue shift of the A1g with increasing the number of layers [19] is observed. In particular, the difference Δ ω = ω A 1 g ω E 2 g between the wavenumbers of these two main modes is commonly taken as a way to evaluate the number of MoS2 layers, with larger Δ ω values generally associated with a thicker MoS2.
The colour maps in Figure 5b,c illustrate the spatial distribution of the Δ ω values obtained from arrays of 50 × 50 Raman spectra collected on 10 μm × 10 μm scan areas. Figure 5d,e show the histograms of the Δ ω values reported in the two maps, with the indication of the corresponding number of MoS2 layers according to the calibration reported in Ref. [19]. The two distributions are quite uniform and exhibit a ω ≈ 21.8 ± 0.6 cm−1 for the few-layer MoS2 sample and ω ≈ 24.8 ± 0.4 cm−1 for the multilayer MoS2 sample. These Δω values are associated with a 2 L–3 L MoS2 thickness for the first sample, in very good agreement with TEM analyses in Figure 4, and to >4 L MoS2 for the second one.
In the following, we will concentrate our attention on the 2 L–3 L MoS2 sample, since the horizontal configuration of the layers makes it more suitable for electronic applications, similarly to 2H-MoS2 samples produced by CVD or by exfoliation from bulk molybdenite.
The doping type and the biaxial strain (ε) of the thin MoS2 film were also evaluated from the Raman maps by a correlative plot of A1g versus E2g peaks positions, as recently discussed in Ref. [23]. Figure 6a shows as blue circles the ω A 1 g and ω E 2 g values extracted from all the Raman spectra in the array of Figure 5. The red line in Figure 6a represents the ideal ω A 1 g   vs .   ω E 2 g dependence (i.e., the strain line) for a purely strained 3L-MoS2 film. This relation is obtained from the combination of the following two expressions:
ω E 2 g = ω 0 E 2 g 2 γ E 2 g ω 0 E 2 g ε
ω A 1 g = ω 0 A 1 g 2 γ A 1 g ω 0 A 1 g ε
Here, γ E 2 g = 0.39 and γ A 1 g = 0.09 are the Grüneisen parameters for the two vibrational modes of 3L-MoS2, estimated from the literature values of the peaks shift rates as a function of strain percentage (−3 cm−1/% and −0.7 cm−1/% for the E2g and A1g peaks, respectively) [56]. ω E 2 g 0 and ω A 1 g 0 represent the E2g and A1g frequencies for an ideally unstrained and undoped 3L-MoS2. Here, the literature values for a suspended 3L-MoS2 membrane ( ω 0 E 2 g = 382.9   cm 1 and ω 0 A 1 g = 406.4   cm 1 ) [56], not affected by the interaction with the substrate, were taken as the best approximation for these ideal values. This reference point is reported as a red square in Figure 6a, while the two arrows with opposite directions along the strain line indicate the tensile (red-shift) and compressive strain (blue-shift), respectively. Furthermore, the black dashed lines serve as guides to estimate the strain values. The distribution of the experimental points (blue circles) in the plot of Figure 6a clearly indicates that the thin MoS2 film on SiO2 is compressively strained. Figure 6b shows the 2D map of the compressive strain, calculated from the map of ω E 2 g values by applying Equation (4). Furthermore, the corresponding histogram of the ε values is reported in Figure 6c, from which an average strain value ε 0.78 % ± 0.18 % can be deduced.
The strain line separates the n-type and p-type doping regions in the ω A 1 g ω E 2 g diagram in Figure 6a. Noteworthy, the experimental points in Figure 6a are partially located in the n-type region and partially in the p-type one. Unintentional n-type doping is typically reported for MoS2 films produced by different synthesis methods (such as mechanical exfoliation or CVD) and it is commonly ascribed to native defects present in the material [57,58,59,60]. Here, the observed p-type doping in some regions of the MoS2 film produced by sulfurization can be associated with the presence of residual MoO3, as deduced by XPS. In fact, several studies demonstrated how intentionally introducing MoO3 in pristine (n-type) MoS2, e.g., by O2 plasma treatments, results in p-type doping of the material [61,62].
The MoS2 thin layers produced by MoO3 thin films sulfurization exhibit large resistivity values in the range of 10–100 Ω∙cm [63]. This can be ascribed, in part, to the nanocrystalline structure of the films, i.e., the large density of grain boundaries, which are known to introduce resistive contributions in the current path [64]. On the other hand, the local changes in the compressive strain distribution, as well in the carrier density, deduced by Raman mapping is expected to have an effect on the electrical properties of the few-layers of MoS2. To get direct information on the homogeneity of conductivity in this film, local current mapping has been carried out by C-AFM, as schematically depicted in Figure 7a. In this configuration, the current locally injected from the AFM metal tip flows in the MoS2 film and is finally collected from the macroscopic front contact. Due to the nanoscale size of the tip contact, the dominant contributions to the measured resistance are represented by the local tip/MoS2 contact resistance and the spreading resistance in the MoS2 region underneath the tip. Figure 7b shows the contact-mode morphological image on the sample surface, from which an RMS roughness ≈ 0.5 nm slightly higher than the one of the as-deposited MoO3 film (Figure 2a) was deduced. Figure 7c,d report the corresponding C-AFM current map and the histogram of the measured current values. The current map clearly shows submicrometer lateral variations of the conductivity, which are only partially correlated to the morphology, while the histogram shows a Gaussian distribution of these values, resembling the shape of the strain distribution in Figure 7d. From this comparison, we can speculate that these mesoscopic-scale inhomogeneities can be partially ascribed to the lateral changes in the strain and carrier density detected by Raman.
In the last section of this paper, Raman and photoluminescence spectra acquired on the few-layers MoS2 samples produced by sulfurization have been compared with reference spectra acquired on CVD-grown MoS2 samples with a similar thickness.
Figure 8 shows a typical Raman spectrum of 3L-MoS2 on SiO2 produced by MoO3 sulfurization, compared with a spectrum of a 3L-MoS2 sample grown by CVD on SiO2 [65], reported as reference. Some remarkable differences can be clearly observed between MoS2 layers prepared using the two different approaches. In fact, besides a lower E2g/A1g intensity ratio, the two vibrational peaks exhibit a more pronounced asymmetric shape in the 3L-MoS2 produced by sulfurization as compared to the CVD-grown one. The deconvolution analysis of the Raman spectra with four Gaussian contributions, associated with the main E2g and A1g modes and the disorder activated LO(M) and ZO(M) modes [66], is also presented in Figure 8. These LO(M) and ZO(M) components are very small in the Raman spectra of CVD 3L-MoS2, whereas their weight is higher in the 3L-MoS2 produced by sulfurization. In this latter case, they can be ascribed both to the nanocrystalline nature of the film, as well as to the presence of residual MoO3, as deduced from the XPS analyses.
Figure 9 shows the comparison between a PL spectrum measured on the 3L-MoS2 produced by sulfurization with a reference spectrum for CVD grown 3L-MoS2, taken from Ref. [65]. For both spectra, acquired using a 532 nm wavelength laser source, the main emission peak at an energy of 1.86 eV can be observed. However, significant differences in spectral features can be clearly identified from a detailed deconvolution analysis.
The PL spectrum of CVD MoS2 can be fitted by three Gaussian peaks, associated with the two exciton contributions (A0 at 1.86 ± 0.01 eV and B at 1.99 ± 0.01 eV, due to the spin-orbit splitting of the valence band) and the trionic contribution (XT at 1.78 ± 0.01 eV) [65]. On the other hand, the deconvolution analysis of the spectrum for the sulfurization grown sample allowed us to identify a fourth component XD at 1.75 ± 0.01 eV, besides the trion (XT at 1.78 ± 0.01 eV) and exciton peaks (A0 at 1.86 ± 0.01 eV and B at 1.95 ± 0.01 eV). Noteworthy, the presence of this XD contribution is accompanied by a strong decrease in the spectral weight of the exciton peak B, as compared to the case of the CVD sample, as well as its FWHM reduction. The occurrence of a similar feature XD, associated with point defects in the MoS2 lattice, has been recently reported by Chow et al. [67] for the PL spectra of MoS2 flakes subjected to soft Ar-plasma irradiation, and it was also accompanied by a decrease in the exciton peak B with respect to unirradiated flakes. Hence, the observed XD contribution for our samples produced by sulfurization was ascribed to a higher density of point defects with respect to CVD grown samples.

4. Conclusions

In conclusion, we reported a detailed analysis of the compositional, morphological, structural, electrical, and optical emission properties of few or multilayer MoS2 on a SiO2/Si substrate produced by sulfurization of very thin MoO3 films at 800 °C. Both Raman mapping and TEM/STEM analyses showed the formation of 2–3 layers of MoS2 nearly aligned with the SiO2 surface after sulfurization of the thinnest MoO3 film, whereas multilayers of MoS2 (partially standing up) were observed for the thicker MoO3 film. The strain distribution in the few-layer MoS2 on SiO2 was evaluated by the correlative plot of the characteristic A1g-E2g Raman modes, showing the occurrence of a compressive strain ε ≈ −0.78 ± 0.18%. Furthermore, the co-existence of submicrometer areas with n- and p-type doping is detected, with the p-type doping probably due to the presence of residual MoO3, as revealed by XPS analyses. Nanoscale resolution current mapping by C-AFM showed conductivity inhomogeneities in the few-layer MoS2, which are well correlated to the lateral changes in the strain detected by Raman. Finally, the characteristics spectroscopic signatures of the defects/disorder were identified by comparing Raman and PL spectra of sulfurization grown MoS2 with reference analyses of CVD-grown single crystalline MoS2.
The demonstrated MoS2 growth method is quite versatile and can be extended to different substrates, besides SiO2. In particular, the adoption of crystalline substrates (such as sapphire, GaN, and 4H-SiC) with the hexagonal basal plane and good lattice matching with MoS2 is expected to enhance the domain size and electronic quality of the grown films. Furthermore, the homogeneous large area few-layer MoS2 can be transferred to arbitrary substrates (including flexible ones) [68] and find applications in different fields of microelectronics, flexible electronics, and sensing.

Author Contributions

Conceptualization, F.G. and A.K.; methodology, A.K.; software, S.E.P.; formal analysis, S.E.P.; investigation, A.K., S.E.P., E.S., G.G., P.F., A.S., M.N., S.D.F., S.A., M.C. and B.P.; resources, S.D.F.; data curation, S.E.P.; writing—original draft preparation, S.E.P.; writing—review and editing, all authors; supervision, F.G., B.P. and F.M.G.; project administration, F.G.; funding acquisition, F.G., B.P. and F.R. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded, in part, by MUR in the framework of the FlagERA-JTC 2019 project ETMOS. E.S. acknowledges the PON project EleGaNTe (ARS01_01007) funded by MUR for financial support. B.P acknowledges funding from the national project TKP2021-NKTA-05. Furthermore, B.P. acknowledges VEKOP-2.3.3-15-2016-00002 of the European Structural and Investment Funds for providing the microscope facility. Part of the experiments was carried out using the facilities of the Italian Infrastructure Beyond Nano.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

We would like to acknowledge the Thermo Fisher Scientific Applications Laboratory, East Grinstead, UK, for providing XPS spectra on MoS2 layers.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O.V.; Kis, A. 2D transition metal dichalcogenides. Nat. Rev. Mater. 2017, 2, 17033. [Google Scholar] [CrossRef]
  2. Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive Photodetectors Based on Monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497–501. [Google Scholar] [CrossRef]
  3. Yin, Z.; Li, H.; Li, H.; Jiang, L.; Shi, Y.; Sun, Y.; Lu, G.; Zhang, Q.; Chen, X.; Zhang, H. Single-Layer MoS2 Phototransistors. ACS Nano 2012, 6, 74–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Li, H.; Yin, Z.; He, Q.; Li, H.; Huang, X.; Lu, G.; Fam, D.W.H.; Tok, A.l.Y.; Zhang, Q.; Zhang, H. Fabrication of Single-and Multilayer MoS2 Film-Based Field-Effect Transistors for Sensing NO at Room Temperature. Small 2012, 8, 63–67. [Google Scholar] [CrossRef] [PubMed]
  5. Radisavljevic, B.; Whitwick, M.B.; Kis, A. Integrated Circuits and Logic Operations Based on Single-Layer MoS2. ACS Nano 2011, 5, 9934–9938. [Google Scholar] [CrossRef] [PubMed]
  6. Ayari, A.; Cobas, E.; Ogundadegbe, O.; Fuhrer, M.S. Realization and Electrical Characterization of Ultrathin Crystals of Layered Transition-Metal Dichalcogenides. J. Appl. Phys. 2007, 101, 014507. [Google Scholar] [CrossRef] [Green Version]
  7. Luo, Y.K.; Xu, J.; Zhu, T.; Wu, G.; McCormick, E.J.; Zhan, W.; Neupane, M.R.; Kawakami, R.K. Opto-Valleytronic Spin Injection in Monolayer MoS2/Few-Layer Graphene Hybrid Spin Valves. Nano Lett. 2017, 17, 3877–3883. [Google Scholar] [CrossRef] [Green Version]
  8. Jiang, J.; Chen, Z.; Hu, Y.; Xiang, Y.; Zhang, L.; Wang, Y.; Wang, G.-C.; Shi, J. Flexo-photovoltaic effect in MoS2. Nat. Nanotechnol. 2021, 16, 894–901. [Google Scholar] [CrossRef]
  9. Hu, J.; Yu, L.; Deng, J.; Wang, Y.; Cheng, K.; Ma, C.; Zhang, Q.; Wen, W.; Yu, S.; Pan, Y.; et al. Sulfur vacancy-rich MoS2 as a catalyst for the hydrogenation of CO2 to methanol. Nat. Catal. 2021, 4, 242–250. [Google Scholar] [CrossRef]
  10. Li, G.; Chen, Z.; Li, Y.; Zhang, D.; Yang, W.; Liu, Y.; Cao, L. Engineering Substrate Interaction To Improve Hydrogen Evolution Catalysis of Monolayer MoS2 Films beyond Pt. ACS Nano 2020, 14, 1707–1714. [Google Scholar] [CrossRef]
  11. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kuc, A.; Zibouche, N.; Heine, T. Influence of Quantum Confinement on the Electronic Structure of the Transition Metal Sulfide TS2. Phys. Rev. B Condens. Matter Mater. Phys. 2011, 83, 245213. [Google Scholar] [CrossRef] [Green Version]
  13. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-Layer MoS2 Transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, W.; De, D.; Chang, S.C.; Wang, Y.; Peng, H.; Bao, J.; Pei, S.S. High mobility and high on/off ratio field-effect transistors based on chemical vapor deposited single-crystal MoS2 grains. Appl. Phys. Lett. 2013, 102, 142106. [Google Scholar] [CrossRef] [Green Version]
  15. Yoon, Y.; Ganapathi, K.; Salahuddin, S. How Good Can Monolayer MoS2 Transistors Be? Nano Lett. 2011, 11, 3768–3773. [Google Scholar] [CrossRef] [PubMed]
  16. Giannazzo, F.; Greco, G.; Roccaforte, F.; Sonde, S.S. Vertical Transistors Based on 2D Materials: Status and Prospects. Crystals 2018, 8, 70. [Google Scholar] [CrossRef] [Green Version]
  17. Giannazzo, F. Engineering 2D heterojunctions with dielectrics. Nat. Electron. 2019, 2, 54. [Google Scholar] [CrossRef]
  18. Novoselov, K.S.; Jiang, D.; Schedin, F.; Booth, T.J.; Khotkevich, V.V.; Morozov, S.V.; Geim, A.K. Two-dimensional atomic crystals. Proc. Natl. Acad. Sci. USA 2005, 102, 10451–10453. [Google Scholar] [CrossRef] [Green Version]
  19. Lee, C.; Yan, H.; Brus, L.E.; Heinz, T.F.; Hone, J.; Ryu, S. Anomalous lattice vibrations of single-and few-layer MoS2. ACS Nano 2010, 4, 2695–2700. [Google Scholar] [CrossRef] [Green Version]
  20. Velický, M.; Donnelly, G.E.; Hendren, W.R.; McFarland, S.; Scullion, D.; DeBenedetti, W.J.I.; Correa, G.C.; Han, Y.; Wain, A.J.; Hines, M.A.; et al. Mechanism of Gold-Assisted Exfoliation of Centimeter-Sized Transition-Metal Dichalcogenide Monolayers. ACS Nano 2018, 12, 10463–10472. [Google Scholar] [CrossRef] [Green Version]
  21. Desai, S.B.; Madhvapathy, S.R.; Amani, M.; Kiriya, D.; Hettick, M.; Tosun, M.; Zhou, Y.; Dubey, M.; Ager, J.W., III; Chrzan, D.; et al. Gold-Mediated Exfoliation of Ultralarge Optoelectronically-Perfect Monolayers. Adv. Mater. 2016, 28, 4053–4058. [Google Scholar] [CrossRef]
  22. Magda, G.Z.; Pető, J.; Dobrik, G.; Hwang, C.; Biró, L.P.; Tapasztó, L. Exfoliation of Large-Area Transition Metal Chalcogenide Single Layers. Sci. Rep. 2015, 5, 14714. [Google Scholar] [CrossRef] [Green Version]
  23. Panasci, S.E.; Schilirò, E.; Migliore, F.; Cannas, M.; Gelardi, F.M.; Roccaforte, F.; Giannazzo, F.; Agnello, S. Substrate impact on the thickness dependence of vibrational and optical properties of large area MoS2 produced by gold-assisted exfoliation. Appl. Phys. Lett. 2021, 119, 093103. [Google Scholar] [CrossRef]
  24. Panasci, S.E.; Schilirò, E.; Greco, G.; Cannas, M.; Gelardi, F.M.; Agnello, S.; Roccaforte, F.; Giannazzo, F. Strain, Doping, and Electronic Transport of Large Area Monolayer MoS2 Exfoliated on Gold and Transferred to an Insulating Substrate. ACS Appl. Mat. Interf. 2021, 13, 31248–31259. [Google Scholar] [CrossRef] [PubMed]
  25. Coleman, J.N.; Lotya, M.; O’Neill, A.; Bergin, S.D.; King, P.J.; Khan, U.; Young, K.; Gaucher, A.; De, S.; Smith, R.J.; et al. Two-dimensional nanosheets produced by liquid exfoliation of layered materials. Science 2011, 331, 568–571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Lee, Y.H.; Zhang, X.-Q.; Zhang, W.; Chang, M.-T.; Lin, C.-T.; Chang, K.-D.; Yu, Y.-C.; Wang, J.T.-W.; Chang, C.-S.; Li, L.-J.; et al. Synthesis of Large-Area MoS2 Atomic Layers with Chemical Vapor Deposition. Adv. Mater. 2012, 24, 2320–2325. [Google Scholar] [CrossRef] [Green Version]
  27. Zhan, Y.; Liu, Z.; Najmaei, S.; Ajayan, P.M.; Lou, J. Large-Area Vapor-Phase Growth and Characterization of MoS2 Atomic Layers on a SiO2 Substrate. Small 2012, 8, 966–971. [Google Scholar] [CrossRef] [Green Version]
  28. Ho, Y.-T.; Ma, C.-H.; Luong, T.-T.; Wei, L.-L.; Yen, T.-C.; Hsu, W.-T.; Chang, W.-H.; Chu, Y.-C.; Tu, Y.-Y.; Pande, K.P.; et al. Layered MoS2 Grown on c-Sapphire by Pulsed Laser Deposition. Phys. Status Solidi RRL 2015, 9, 187–191. [Google Scholar] [CrossRef]
  29. Fu, D.; Zhao, X.; Zhang, Y.-Y.; Li, L.; Xu, H.; Jang, A.-R.; Yoon, S.I.; Song, P.; Poh, S.M.; Ren, T.; et al. Molecular Beam Epitaxy of Highly Crystalline Monolayer Molybdenum Disulfide on Hexagonal Boron Nitride. J. Am. Chem. Soc. 2017, 139, 9392–9400. [Google Scholar] [CrossRef]
  30. Valdivia, A.; Tweet, D.J.; Conley, J.F., Jr. Atomic layer deposition of two dimensional MoS2 on 150 mm substrates. J. Vac. Sci. Technol. 2016, 34, 21515. [Google Scholar] [CrossRef] [Green Version]
  31. Najmaei, S.; Liu, Z.; Zhou, W.; Zou, X.; Shi, G.; Lei, S.; Yakobson, B.I.; Idrobo, J.-C.; Ajayan, P.M.; Lou, J. Vapour phase growth and grain boundary structure of molybdenum disulphide atomic layers. Nat. Mater. 2013, 12, 754–759. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, H.F.; Wong, S.L.; Chi, D.Z. CVD growth of MoS2-based two-dimensional materials. Chem. Vap. Depos. 2015, 21, 241–259. [Google Scholar] [CrossRef]
  33. Jeon, J.; Jang, S.K.; Jeon, S.M.; Yoo, G.; Jang, Y.H.; Park, J.H.; Lee, S. Layer-controlled CVD growth of large-area two-dimensional MoS2 films. Nanoscale 2015, 7, 1688–1695. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, J.; Tang, W.; Tian, B.; Liu, B.; Zhao, X.; Liu, Y.; Ren, T.; Liu, W.; Geng, D.; Jeong, H.Y.; et al. Chemical Vapor Deposition of High-Quality Large-Sized MoS2 Crystals on Silicon Dioxide Substrates. Adv. Sci. 2016, 3, 1600033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Zhang, Z.; Chen, P.; Duan, X.; Zang, K.; Luo, J.; Duan, X. Robust epitaxial growth of two-dimensional heterostructures, multiheterostructures, and superlattices. Science 2017, 357, 788–792. [Google Scholar] [CrossRef] [Green Version]
  36. Liu, B.; Fathi, M.; Chen, L.; Abbas, A.; Ma, Y.; Zhou, C. Chemical vapor deposition growth of monolayer WSe2 with tunable device characteristics and growth mechanism study. ACS Nano 2015, 9, 6119–6127. [Google Scholar] [CrossRef]
  37. Tang, L.; Tan, J.; Nong, H.; Liu, B.; Cheng, H.M. Chemical Vapor Deposition Growth of Two-Dimensional Compound Materials: Controllability, Material Quality, and Growth Mechanism. Acc. Mater. Res. 2020, 2, 36–47. [Google Scholar] [CrossRef]
  38. Wang, S.; Rong, Y.; Fan, Y.; Pacios, M.; Bhaskaran, H.; He, K.; Warner, J.H. Shape evolution of monolayer MoS2 crystals grown by chemical vapor deposition. Chem. Mater. 2014, 26, 6371–6379. [Google Scholar] [CrossRef]
  39. Yang, S.Y.; Shim, G.W.; Seo, S.B.; Choi, S.Y. Effective shape-controlled growth of monolayer MoS2 flakes by powder-based chemical vapor deposition. Nano Res. 2017, 10, 255–262. [Google Scholar] [CrossRef]
  40. Wu, C.R.; Chang, X.R.; Wu, C.H.; Lin, S.Y. The growth mechanism of transition metal dichalcogenides by using sulfurization of pre-deposited transition metals and the 2D crystal hetero-structure establishment. Sci. Rep. 2017, 7, 42146. [Google Scholar] [CrossRef] [Green Version]
  41. Li, D.; Xiao, Z.; Mu, S.; Wang, F.; Liu, Y.; Song, J.; Huang, X.; Jiang, L.; Xiao, J.; Liu, L.; et al. A facile space-confined solid-phase sulfurization strategy for growth of high-quality ultrathin molybdenum disulfide single crystals. Nano Lett. 2018, 18, 2021–2032. [Google Scholar] [CrossRef] [PubMed]
  42. Taheri, P.; Wang, J.; Xing, H.; Destino, J.F.; Arik, M.M.; Zhao, C.; Kang, K.; Blizzard, B.; Zhang, L.; Zhao, P. Growth mechanism of largescale MoS2 monolayer by sulfurization of MoO3 film. Mater. Res. Expr. 2016, 3, 075009. [Google Scholar] [CrossRef]
  43. Hutar, P.; Spankova, M.; Sojkova, M.; Dobrocka, E.; Vegso, K.; Hagara, J.; Halahovets, Y.; Majkova, E.; Siffalovic, P.; Hulman, M. Highly crystalline MoS2 thin films fabricated by sulfurization. Phys. Status Solidi (B) 2019, 256, 1900342. [Google Scholar] [CrossRef]
  44. Vangelista, S.; Cinquanta, E.; Martella, C.; Alia, M.; Longo, M.; Lamperti, A.; Mantovan, R.; Basso Basset, F.; Pezzoli, F.; Molle, A. Towards a uniform and large-scale deposition of MoS2 nanosheets via sulfurization of ultra-thin Mo based solid films. Nanotechnology 2016, 27, 175703. [Google Scholar] [CrossRef]
  45. Kong, D.; Wang, H.; Cha, J.J.; Pasta, M.; Koski, K.J.; Yao, J.; Cui, Y. Synthesis of MoS2 and MoSe2 Films with Vertically Aligned Layers. Nano Lett. 2013, 13, 1341–1347. [Google Scholar] [CrossRef] [PubMed]
  46. Cho, S.-Y.; Kim, S.J.; Lee, Y.; Kim, J.-S.; Jung, W.-B.; Yoo, H.-W.; Kim, J.; Jung, H.-T. Highly Enhanced Gas Adsorption Properties in Vertically Aligned MoS2 Layers. ACS Nano 2015, 9, 9314–9321. [Google Scholar] [CrossRef]
  47. Jung, Y.; Shen, J.; Liu, Y.; Woods, J.M.; Sun, Y.; Cha, J.J. Metal Seed Layer Thickness-Induced Transition from Vertical to Horizontal Growth of MoS2 and WS2. Nano Lett. 2014, 14, 6842–6849. [Google Scholar] [CrossRef]
  48. Stern, C.; Grinvald, S.; Kirshner, M.; Sinai, O.; Oksman, M.; Alon, H.; Meiron, O.E.; Bar-Sadan, M.; Houben, L.; Naveh, D. Growth Mechanisms and Electronic Properties of Vertically Aligned MoS2. Sci. Rep. 2018, 8, 16480. [Google Scholar] [CrossRef]
  49. Shang, S.-L.; Lindwall, G.; Wang, Y.; Redwing, J.M.; Anderson, T.; Liu, Z.-K. Lateral Versus Vertical Growth of Two-Dimensional Layered Transition-Metal Dichalcogenides: Thermodynamic Insight into MoS2. Nano Lett. 2016, 16, 5742–5750. [Google Scholar] [CrossRef]
  50. Sojková, M.; Vegso, K.; Mrkyvkova, N.; Hagara, J.; Hutár, P.; Rosová, A.; Čaplovičová, M.; Ludacka, V.; Majková, E.; Siffalovic, P.; et al. Tuning the orientation of few-layer MoS2 films using one-zone sulfurization. RSC Adv. 2019, 9, 29645–29651. [Google Scholar] [CrossRef] [Green Version]
  51. Shahzad, R.; Kim, T.; Kang, S.W. Effects of temperature and pressure on sulfurization of molybdenum nano-sheets for MoS2 synthesis. Thin Solid Film. 2017, 641, 79–86. [Google Scholar] [CrossRef]
  52. Lee II, E.W.; Ma, L.; Nath, D.N.; Lee, C.H.; Arehart, A.; Wu, Y.; Rajan, S. Growth and electrical characterization of two-dimensional layered MoS2/SiC heterojunctions. Appl. Phys. Lett. 2014, 105, 203504. [Google Scholar] [CrossRef]
  53. Wang, H.; Zhang, Q.; Yao, H.; Liang, Z.; Lee, H.-W.; Hsu, P.-C.; Zheng, G.; Cui, Y. High electrochemical selectivity of edge versus terrace sites in two-dimensional layered MoS2 materials. Nano Lett. 2014, 14, 7138–7144. [Google Scholar] [CrossRef] [PubMed]
  54. Hadouda, H.; Pouzet, J.; Bernede, J.C.; Barreau, A. MoS2 thin film synthesis by soft sulfurization of a molybdenum layer. Mater. Chem. Phys. 1995, 42, 291–297. [Google Scholar] [CrossRef]
  55. Naujokaitis, A.; Gaigalas, P.; Bittencourt, C.; Mickevičius, S.; Jagminas, A. 1T/2H MoS2/MoO3 hybrid assembles with glycine as highly efficient and stable electrocatalyst for water splitting. Int. J. Hydrog. Energy 2019, 44, 24237–24245. [Google Scholar] [CrossRef]
  56. Lloyd, D.; Liu, X.; Christopher, J.S.; Cantley, L.; Wadehra, A.; Kim, B.L.; Goldberg, B.B.; Swan, A.K.; Bunch, J.S. Band Gap Engineering with Ultralarge Biaxial Strains in Suspended Monolayer MoS2. Nano Lett. 2016, 16, 5836–5841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Bampoulis, P.; van Bremen, R.; Yao, Q.; Poelsema, B.; Zandvliet, H.J.W.; Sotthewes, K. Defect Dominated Charge Transport and Fermi Level Pinning in MoS2/Metal Contacts. ACS Appl. Mater. Interfaces 2017, 9, 19278–19286. [Google Scholar] [CrossRef] [Green Version]
  58. Sotthewes, K.; van Bremen, R.; Dollekamp, E.; Boulogne, T.; Nowakowski, K.; Kas, D.; Zandvliet Harold, J.W.; Bampoulis, P. Universal Fermi-Level Pinning in Transition-Metal Dichalcogenides. J. Phys. Chem. C 2019, 123, 5411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Giannazzo, F.; Schilirò, E.; Greco, G.; Roccaforte, F. Conductive Atomic Force Microscopy of Semiconducting Transition Metal Dichalcogenides and Heterostructures. Nanomaterials 2020, 10, 803. [Google Scholar] [CrossRef] [PubMed]
  60. Giannazzo, F.; Fisichella, G.; Piazza, A.; Agnello, S.; Roccaforte, F. Nanoscale Inhomogeneity of the Schottky Barrier and Resistivity in MoS2 Multilayers. Phys. Rev. B 2015, 92, 081307. [Google Scholar] [CrossRef]
  61. Zhu, H.; Qin, X.; Cheng, L.; Azcatl, A.; Kim, J.; Wallace, R.M. Remote Plasma Oxidation and Atomic Layer Etching of MoS2. ACS Appl. Mater. Interfaces 2016, 8, 19119–19126. [Google Scholar] [CrossRef] [PubMed]
  62. Giannazzo, F.; Fisichella, G.; Greco, G.; Di Franco, S.; Deretzis, I.; La Magna, A.; Bongiorno, C.; Nicotra, G.; Spinella, C.; Scopelliti, M.; et al. Ambipolar MoS2 Transistors by Nanoscale Tailoring of Schottky Barrier Using Oxygen Plasma Functionalization. ACS Appl. Mater. Interfaces 2017, 9, 23164–23174. [Google Scholar] [CrossRef]
  63. Hamada, T.; Tomiya, S.; Tatsumi, T.; Hamada, M.; Horiguchi, T.; Kakushima, K.; Tsutsui, K.; Wakabayashi, H. Sheet Resistance Reduction of MoS2 Film Using Sputtering and Chlorine Plasma Treatment Followed by Sulfur Vapor Annealing. IEEE J. Electron Devices Soc. 2021, 9, 278–285. [Google Scholar] [CrossRef]
  64. Giannazzo, F.; Bosi, M.; Fabbri, F.; Schilirò, E.; Greco, G.; Roccaforte, F. Direct Probing of Grain Boundary Resistance in Chemical Vapor Deposition-Grown Monolayer MoS2 by Conductive Atomic Force Microscopy. Phys. Status Solidi RRL 2020, 14, 1900393. [Google Scholar] [CrossRef]
  65. Golovynskyi, S.; Irfan, I.; Bosi, M.; Seravalli, L.; Datsenko, O.I.; Golovynska, I.; Li, B.; Lin, D.; Qu, J. Exciton and trion in few-layer MoS2: Thickness- and temperature-dependent photoluminescence. Appl. Surf. Sci. 2020, 515, 146033. [Google Scholar] [CrossRef]
  66. Mignuzzi, S.; Pollard, A.J.; Bonini, N.; Brennan, B.; Gilmore, I.S.; Pimenta, M.A.; Richards, D.; Roy, D. Effect of disorder on Raman scattering of single-layer MoS2. Phys. Rev. B. 2015, 91, 195411. [Google Scholar] [CrossRef] [Green Version]
  67. Chow, P.K.; Jacobs-Gedrim, R.B.; Gao, J.; Lu, T.-M.; Yu, B.; Terrones, H.; Koratkar, N. Defect-Induced Photoluminescence in Monolayer Semiconducting Transition Metal Dichalcogenides. ACS Nano 2015, 9, 1520–1527. [Google Scholar] [CrossRef]
  68. Watson, A.J.; Lu, W.; Guimaraes, M.H.D.; Stöhr, M. Transfer of large-scale two-dimensional semiconductors: Challenges and developments. 2D Mater. 2021, 8, 032001. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the sulfurization process of the thin MoO3 films on the SiO2/Si substrates.
Figure 1. Schematic illustration of the sulfurization process of the thin MoO3 films on the SiO2/Si substrates.
Nanomaterials 12 00182 g001
Figure 2. (a) Typical AFM morphology of as-deposited MoO3 thin films on SiO2, with the indication of the root mean square (RMS) roughness. (b,c) Determination of the thickness of films deposited with two different sputtering times by measurement of the step heights (~2.8 nm and ~4.2 nm) with respect to SiO2 on scratched regions.
Figure 2. (a) Typical AFM morphology of as-deposited MoO3 thin films on SiO2, with the indication of the root mean square (RMS) roughness. (b,c) Determination of the thickness of films deposited with two different sputtering times by measurement of the step heights (~2.8 nm and ~4.2 nm) with respect to SiO2 on scratched regions.
Nanomaterials 12 00182 g002
Figure 3. (a) Survey XPS spectrum of MoS2 on SiO2 produced by sulfurization of the 2.8 nm MoO3 film, with the indication of the evaluated surface elemental composition. (b) XPS spectra of the Mo 3d and S 2s core levels, with the deconvolution of the Mo4+ contribution (related to MoS2) and the Mo6+ contribution (related to residual MoO3). (c) S 2p core levels spectra, indicating the predominance of the sulphide contribution, with a small S-O component.
Figure 3. (a) Survey XPS spectrum of MoS2 on SiO2 produced by sulfurization of the 2.8 nm MoO3 film, with the indication of the evaluated surface elemental composition. (b) XPS spectra of the Mo 3d and S 2s core levels, with the deconvolution of the Mo4+ contribution (related to MoS2) and the Mo6+ contribution (related to residual MoO3). (c) S 2p core levels spectra, indicating the predominance of the sulphide contribution, with a small S-O component.
Nanomaterials 12 00182 g003
Figure 4. Cross sectional HR-TEM (a) and HAADF-STEM (b) images of few-layers MoS2 obtained by sulfurization of the ~2.8 nm MoO3 film on the SiO2 substrate. MoS2 is composed by nearly horizontally aligned 2–3 layers. The interlayer spacing in a 3-layers region is evaluated from the HR-TEM in the insert of panel (a). Cross sectional HR-TEM (c) and HAADF-STEM (d) of multilayers MoS2 obtained by sulfurization of the ~4.2 nm MoO3 film.
Figure 4. Cross sectional HR-TEM (a) and HAADF-STEM (b) images of few-layers MoS2 obtained by sulfurization of the ~2.8 nm MoO3 film on the SiO2 substrate. MoS2 is composed by nearly horizontally aligned 2–3 layers. The interlayer spacing in a 3-layers region is evaluated from the HR-TEM in the insert of panel (a). Cross sectional HR-TEM (c) and HAADF-STEM (d) of multilayers MoS2 obtained by sulfurization of the ~4.2 nm MoO3 film.
Nanomaterials 12 00182 g004
Figure 5. (a) Representative Raman spectra of the few-layers (FL) MoS2 (black-line) and multilayer (ML) MoS2 samples obtained by sulfurization of the 2.8 and 4.2 nm MoO3 films on SiO2. Colour maps of the A1g-E2g wavenumber difference Δω obtained from arrays of Raman spectra collected on 10 μm × 10 μm scan areas on the FL-MoS2 (b) and on the ML-MoS2 (c) samples. Histogram of Δω values showing a distribution with a peak at ω ≈ 21.8 ± 0.6 cm−1 for the FL-MoS2 sample associated to 2 L–3 L MoS2 (d) and ω ≈ 24.8 ± 0.4 cm−1 for the ML-MoS2 sample, corresponding to >4 L MoS2 thickness (e).
Figure 5. (a) Representative Raman spectra of the few-layers (FL) MoS2 (black-line) and multilayer (ML) MoS2 samples obtained by sulfurization of the 2.8 and 4.2 nm MoO3 films on SiO2. Colour maps of the A1g-E2g wavenumber difference Δω obtained from arrays of Raman spectra collected on 10 μm × 10 μm scan areas on the FL-MoS2 (b) and on the ML-MoS2 (c) samples. Histogram of Δω values showing a distribution with a peak at ω ≈ 21.8 ± 0.6 cm−1 for the FL-MoS2 sample associated to 2 L–3 L MoS2 (d) and ω ≈ 24.8 ± 0.4 cm−1 for the ML-MoS2 sample, corresponding to >4 L MoS2 thickness (e).
Nanomaterials 12 00182 g005
Figure 6. (a) Correlative plot of the ωA1g and ωE2g values (blue circles) extracted from all the Raman spectra in the array of Figure 5. The red line represents the ideal ωA1g vs. ωE2g dependence (i.e., the strain line) for a purely strained 3L-MoS2 film. The red square corresponds to the frequencies ω E 2 g 0 and ω A 1 g 0 for an ideally unstrained and undoped 3L-MoS2, while the two red arrows with opposite directions along the strain line indicate the tensile (red-shift) and compressive strain (blue-shift), respectively. (b) Map and (c) corresponding histogram of the compressive strain on a 10 μm × 10 μm area.
Figure 6. (a) Correlative plot of the ωA1g and ωE2g values (blue circles) extracted from all the Raman spectra in the array of Figure 5. The red line represents the ideal ωA1g vs. ωE2g dependence (i.e., the strain line) for a purely strained 3L-MoS2 film. The red square corresponds to the frequencies ω E 2 g 0 and ω A 1 g 0 for an ideally unstrained and undoped 3L-MoS2, while the two red arrows with opposite directions along the strain line indicate the tensile (red-shift) and compressive strain (blue-shift), respectively. (b) Map and (c) corresponding histogram of the compressive strain on a 10 μm × 10 μm area.
Nanomaterials 12 00182 g006
Figure 7. (a) Schematic of the C-AFM setup for local conductivity mapping of few-layers MoS2 on SiO2. (b) Morphology and (c) current map simultaneously measured with tip-to-sample bias of 5 V. (d) Histogram of current values from the C-AFM map.
Figure 7. (a) Schematic of the C-AFM setup for local conductivity mapping of few-layers MoS2 on SiO2. (b) Morphology and (c) current map simultaneously measured with tip-to-sample bias of 5 V. (d) Histogram of current values from the C-AFM map.
Nanomaterials 12 00182 g007
Figure 8. Raman spectrum for 3L-MoS2 produced by sulfurization (red), compared with a reference spectrum for CVD grown 3L-MoS2 (blue). Data for CVD 3L-MoS2 were adapted with permission from [65], copyright Elsevier 2020.
Figure 8. Raman spectrum for 3L-MoS2 produced by sulfurization (red), compared with a reference spectrum for CVD grown 3L-MoS2 (blue). Data for CVD 3L-MoS2 were adapted with permission from [65], copyright Elsevier 2020.
Nanomaterials 12 00182 g008
Figure 9. Photoluminescence (PL) spectra for 3L-MoS2 produced by sulfurization, compared with a reference spectrum for CVD grown 3L-MoS2. The deconvolution analysis indicated the presence of the excitonic contributions A0, B, and of the trionic contribution XT (grey lines) for the CVD grown sample. In addition, the defect-related peak XD (orange line) is identified in the sulfurization grown sample. Data for CVD 3L-MoS2 were adapted with permission from [65], copyright Elsevier 2020.
Figure 9. Photoluminescence (PL) spectra for 3L-MoS2 produced by sulfurization, compared with a reference spectrum for CVD grown 3L-MoS2. The deconvolution analysis indicated the presence of the excitonic contributions A0, B, and of the trionic contribution XT (grey lines) for the CVD grown sample. In addition, the defect-related peak XD (orange line) is identified in the sulfurization grown sample. Data for CVD 3L-MoS2 were adapted with permission from [65], copyright Elsevier 2020.
Nanomaterials 12 00182 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Panasci, S.E.; Koos, A.; Schilirò, E.; Di Franco, S.; Greco, G.; Fiorenza, P.; Roccaforte, F.; Agnello, S.; Cannas, M.; Gelardi, F.M.; et al. Multiscale Investigation of the Structural, Electrical and Photoluminescence Properties of MoS2 Obtained by MoO3 Sulfurization. Nanomaterials 2022, 12, 182. https://doi.org/10.3390/nano12020182

AMA Style

Panasci SE, Koos A, Schilirò E, Di Franco S, Greco G, Fiorenza P, Roccaforte F, Agnello S, Cannas M, Gelardi FM, et al. Multiscale Investigation of the Structural, Electrical and Photoluminescence Properties of MoS2 Obtained by MoO3 Sulfurization. Nanomaterials. 2022; 12(2):182. https://doi.org/10.3390/nano12020182

Chicago/Turabian Style

Panasci, Salvatore E., Antal Koos, Emanuela Schilirò, Salvatore Di Franco, Giuseppe Greco, Patrick Fiorenza, Fabrizio Roccaforte, Simonpietro Agnello, Marco Cannas, Franco M. Gelardi, and et al. 2022. "Multiscale Investigation of the Structural, Electrical and Photoluminescence Properties of MoS2 Obtained by MoO3 Sulfurization" Nanomaterials 12, no. 2: 182. https://doi.org/10.3390/nano12020182

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop