Next Article in Journal
Manifesting Epoxide and Hydroxyl Groups in XPS Spectra and Valence Band of Graphene Derivatives
Next Article in Special Issue
Interface Engineering Modulated Valley Polarization in MoS2/hBN Heterostructure
Previous Article in Journal
piRNA and miRNA Can Suppress the Expression of Multiple Sclerosis Candidate Genes
Previous Article in Special Issue
Study of Defects and Nano-patterned Substrate Regulation Mechanism in AlN Epilayers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photodetection Properties of MoS2, WS2 and MoxW1-xS2 Heterostructure: A Comparative Study

1
Laboratory of Physics of Condensed Mater, University of Picardie Jules Verne, 80039 Amiens, France
2
Technology Innovation Institute, Abu Dhabi P.O. Box 9639, United Arab Emirates
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(1), 24; https://doi.org/10.3390/nano13010024
Submission received: 24 November 2022 / Revised: 4 December 2022 / Accepted: 16 December 2022 / Published: 21 December 2022
(This article belongs to the Special Issue Advances in Nanostructured Semiconductors and Heterojunctions)

Abstract

:
Layered transition metals dichalcogenides such as MoS2 and WS2 have shown a tunable bandgap, making them highly desirable for optoelectronic applications. Here, we report on one-step chemical vapor deposited MoS2, WS2 and MoxW1-xS2 heterostructures incorporated into photoconductive devices to be examined and compared in view of their use as potential photodetectors. Vertically aligned MoS2 nanosheets and horizontally stacked WS2 layers, and their heterostructure form MoxW1-xS2, exhibit direct and indirect bandgap, respectively. To analyze these structures, various characterization methods were used to elucidate their properties including Raman spectroscopy, X-ray diffraction, X-ray photoelectron spectrometry and high-resolution transmission electron microscopy. While all the investigated samples show a photoresponse in a broad wavelength range between 400 nm and 700 nm, the vertical MoS2 nanosheets sample exhibits the highest performances at a low bias voltage of 5 V. Our findings demonstrate a responsivity and a specific detectivity of 47.4 mA W−1 and 1.4 × 1011 Jones, respectively, achieved by MoxW1-xS2. This study offers insights into the use of a facile elaboration technique for tuning the performance of MoxW1-xS2 heterostructure-based photodetectors.

1. Introduction

Semiconductor photodetectors, namely photodiodes, are the most common types of detectors used in optical communication systems owing to their compact size, fast detection speed and high detection efficiency. Practical photodiodes can have a variety of device structures, beyond the basic PN junction construction, to improve their efficiency [1,2]. High-performance photodetectors have been used in a wide range of applications, including electro-optical displays [3], imaging [4], environmental monitoring [5], optical communication [6], military applications and security checks [7]. In these domains, two-dimensional (2D) materials, especially transition-metal dichalcogenides (TMDs), are becoming more attractive for designing photodetectors [8,9,10] due to their unique properties such as their ability to operate in the full range of visible light while having high photodetection polarization sensitivity, a fast photoresponse and high spatially resolved imaging [9]. This class of materials exhibits a layer-dependent electronic band structure in terms of unique physical characteristics and detection mechanisms [11,12]. Photodetectors based on 2D-TMDs materials are more sensitive throughout a wide range of the electromagnetic spectrum compared to photodetectors based on conventional bulk semiconductors [8]. However, an enhanced absorption coefficient and a longer lifespan of photoexcited carriers are preferred for optimal photocurrent generation and photodetector operation. For instance, MoS2 thin films have a high light absorption coefficient of 107 m−1, a high absorption of 95% of total light [13] and a direct bandgap of 1.8 eV. Moreover, photodetectors based on 2D-TMDs possess a good current on/off ratio, efficiency, higher chemical and mechanical stability and a stronger light–matter interaction compared to conventional photodetectors. Furthermore, mono- and few-layers MoS2 present other desirable electronic properties that make them suitable for optoelectronic applications, such as their high carrier mobility and electrostatic integrity [14,15,16]. Owing to these properties, a MoS2 single-layered transistor was found to achieve a responsivity up to 880 × 103 mA W−1 under a 560 nm excitation [17]. Moreover, an ultrabroadband multilayer MoS2 photodetector was reported to operate in 445–2717 nm range achieving a responsivity and a specific detectivity of 50.7 mA W−1 and 1.6 × 109 Jones, respectively [18]. It was also shown that the photoresponse of the MoS2 photodetector could be enhanced by chemical doping to improve its responsivity and specific detectivity up to 105 mA W−1 and 9.4 × 1012 Jones, respectively [19]. The reported responsivity and specific detectivity values are approximately 15 and 5 times higher relative to those of pristine photodetector. In addition to that, WS2 is considered as another promising candidate for photodetection owing to a range of outstanding properties, such as its bandgap tunability, its high carrier mobility and its efficient optical absorption [20,21]. WS2-based photodetectors have been reported to exhibit a responsivity of 4 mA W−1 at operating wavelengths ranging from the visible to the near infrared (IR) range [22]. Once combined with other 2D-materials such as graphene, the resulting heterostructure has shown a higher responsivity and specific detectivity of 3.5 × 103 mA W−1 and 1012 Jones, respectively [23]. First attempts to use a MoxW1-xS2 heterostructure as a photodetector showed promising performances such as a responsivity of 2.3 × 103 mA W−1 obtained under 450 nm excitation [24], and a responsivity and a specific detectivity of 6.7 × 106 mA W−1 and 3.1 × 1013 Jones under 457 nm laser light, respectively [25]. Other interesting works on alloying and ternary 2D-TMD materials can be found elsewhere [26,27,28].
In this work, a systematic comparative study is conducted on MoS2, WS2 and MoxW1-xS2 heterostructure to emphasize their photodetection performances while using identical fabrication and analysis routes. All samples were fabricated in one single step chemical vapor deposition (CVD) process and underwent extensive characterization investigations.

2. Materials and Methods

Among several techniques used to fabricate 2D materials, CVD is the commonly employed technique to control defects, crystallinity, and morphology of this class of materials. In particular, CVD sulfurization process is a facile one-step processing route that allows the fabrication of several sulfur based 2D-materials. The fabrication control is often ensured by monitoring several processing parameters such as gas flow, temperature, heating rate, the distance between precursors, and the position and height of the collecting substrate [29]. The synthesis of all samples was obtained in a one-step CVD process using a single-heating zone furnace at atmospheric pressure, as shown in Figure 1. The CVD system mainly consists of a quartz tube connected to high-purity (99.999%) argon cylinder streaming at flow rates of 70 sccm and 50 sccm for MoS2 fabrication and for WS2 and MoS2/WS2 respectively. The SiO2/Si (1 × 1 cm2) substrates were rinsed successively in deionized water, acetone, and ethanol in an ultrasonic bath for 10 min each.
A powder consisting of WO3 (≥99.8%), MoO3 (≥99.8%), or mixed WO3/MoO3 with ratio 1:1 was mixed with Sulphur ≥99.98% using a ball-milling machine. All chemicals were purchased from Sigma Aldrich (Saint Louis, MO, USA). A diluted suspension solution with a concentration of 100 mg/mL was prepared using either the WO3/S, the MoO3/S, or the MoO3-WO3/S with ethanol and subsequently sonicated to enhance the homogeneity of the solution. Prior to the CVD process, a drop of 10 μL of the suspension solution was directly dropped onto the cleaned substrate using a pipette, as shown in Figure 1. Subsequently a 250 mg of sulfur powder was introduced at the edge of the sealed end of the quartz tube. Initially, the furnace was heated from room temperature up to 400 °C at a 20 °C/min heating rate, then to 850 °C at 5 °C/min for MoS2 fabrication and 950 °C at 5 °C/min for WS2 and MoS2/WS2. During the growth of either the MoS2, the WS2, or the MoxW1-xS2, the temperature was maintained at 850 °C or 950 °C for 30 min. Then, the furnace was allowed to cool down naturally to room temperature.
The morphology of the fabricated specimens was analyzed using a dual beam focused-ion beam and scanning electron microscope (FIB-SEM) Scios 2 ThermoFisher Scientific (Waltham, MA, USA) microscope. The same tool was also used for the preparation of thin lamella for the transmission electron microscopy (TEM) study. TEM study was carried out using Tecnai and Titan systems from ThermoFisher Scientific (Waltham, MA, USA). TEM samples were prepared on a thin carbon coated Cu mesh grid by transferring the grown TMD samples through a gentle physical exfoliation. The vibrational modes of the processed samples were examined with a micro-Raman spectrometer Renishaw (Wotton-under-Edge, UK), using a laser excitation of 532 nm. The crystalline structure was investigated by X-ray diffraction (XRD) using a D8 Discover diffractometer Bruker (Billerica, MA, USA); KaCu = 1.54 Å. The optical properties were investigated using a UV-Vis-near IR spectrometer JASCO V-670. An X-ray photoelectron spectroscopy (XPS) study was carried out using a PHI VersaProbe III scanning XPS microprobe Physical Electronics (Chanhassen, MN, USA), equipped with a monochromatic and microfocused Al K-Alpha X-ray source (1486.6 eV). During the experiment, an E-neutralizer (1 V), was implemented. CasaXPS processing software 2.3. was used for the calibration and the curve fitting. Finally, electrical measurements were performed using Palmsens-4 electrochemical workstation under ambient conditions.

3. Results and Discussion

3.1. Material Characterization

Figure 2a presents the Raman spectra of the MoS2 sample. The observed main Raman vibrational modes indicate the presence of hexagonal 2H-MoS2 such as E12g (382 cm−1) and A1g (409.8 cm−1), which correspond to the in-plane and out-of-plane atomic vibrations, respectively [30].
The main Raman vibration modes recorded for WS2 correspond to a 2H-WS2 structure such as 2LA(M), E12g and A1g as shown in Figure 2b. The strongest peak at 350 cm−1 may be fitted with two sub-peaks with maximum frequencies of 323.6 cm−1 and 351.3 cm−1 leading to 2LA(M) and E12g, respectively. The first-order vibrational mode E12g represents the in-plane vibration between sulfur and tungsten atoms while the A1g vibrational mode at 420 cm−1 corresponds to the out-of-plane vibration of sulfur atoms. It is worth noting that the A1g is sensitive to the number of WS2 layers [31].
Regarding the MoxW1-xS2 heterostructure, the Raman peaks corresponding to 2H-MoS2 and 2H-WS2 are present, as shown in Figure 2c. The positions of the E12g and A1g vibrational modes do not seem to shift compared to the observed peaks in individual samples. This indicates that WS2 and MoS2, obtained through our preparation route, have no effect on each other’s long-range Coulomb interactions between the effective charges as previously reported [24].
The x-ray diffraction (XRD) diagram illustrated in Figure 3a shows clear diffraction peaks at 14.25°, 25.81°, 32.15°, 44.13° and 60.21° corresponding to 2H-MoS2. They are attributed, respectively, to the (002), (004), (103), (006) and (008) planes of the hexagonal 2H-MoS2. For WS2, the XRD diagram shows several significant diffraction peaks at 14.3°, 28.8°, 43.9°, 59.8°, and 77.13° as can be seen in Figure 3b. They are attributed to 2H-WS2 planes (002), (004), (006), (008) and (00,10).
For the MoxW1-xS2 heterostructure, the MoS2 and WS2 peaks are present in the corresponding XRD diagram given in Figure 3c. This confirms the successful fabrication of the heterostructure. The sharp diffraction peaks observed on the spectra are a clear indication of the high crystallinity of the fabricated nanosheets.
The SEM images show different morphologies for the MoS2, WS2 and the heterostructure samples. The MoS2 flakes (Figure 4a) are observed to grow vertically. This is highlighted at a higher magnification in Figure 4b. Similar results were reported previously [32]. On the other hand, WS2 shows accumulated crystals stacked on top of the substrate. A large number of triangular shaped flakes disposed horizontally are visible in Figure 4c,d at low and high magnifications, respectively.
As can be seen in Figure 4e,f, the MoxW1-xS2 heterostructure exhibits a mixed morphology. It consists of both vertically aligned MoS2 nanosheets and stacked layers of WS2. A coherence between the two phases is observed with no visible segregation between the two compounds.
The change in the MoS2 and WS2 morphology could be attributed to the following hypotheses: (1) The high CVD reaction temperature used to process the WS2 could enhance the nucleation kinetics of the first WS2 seeds allowing the coalescence process to occur horizontally. In contrast, for MoS2 the reaction temperature is lower leading to dispersed seeds in the surface of the substrate favoring the coalescence on the top of the first layers; (2) The WS2 weight may impede the vertical shape of WS2; (3) The flow rate used for the processing of WS2 (50 sscm) is lower compared to the one for MoS2 fabrication (70 sccm), which may not allow the evacuation of sulfur excess.
In order to comprehend the mixing mechanism of MoS2 and WS2 structures, we have conducted further microstructure analysis using HRTEM for the three samples as shown in Figure 5.
Figure 5a shows a TEM cross-sectional view of the vertically oriented MoS2 nanosheets (thickness ~100 nm). Higher resolution imaging (Figure 5b) indicates an interplanar spacing of 2H-MoS2 of ~0.62 nm. Moreover, Figure 5c depicts the base region at the interface between the MoS2 and the substrate, showing the nucleation of the 2H-MoS2. The zoomed view in Figure 5c indicates the nature of the flakes’ growth, where certain layers tend to be continuous and few sheets get terminated due to the absence of growth space, which could be at the origin of the vertically aligned 2H-MoS2. Figure 5e–g shows the cross-sectional views of the WS2 sample. The smooth planar growth of the WS2 is clearly visible compared to MoS2, in agreement with the observation on the SEM images. An interplanar distance of 2H-WS2 is determined at ~0.65 nm. Figure 5h shows typical TEM bright field images obtained from the MoxW1-xS2 heterostructure showing an overlapping region (red and purple boxes). Zoomed images of both boxes indicate the presence of 2H-MoS2 (red box, Figure 5i) and 2H-WS2 (purple box, Figure 5j), which is a signature of the MoS2-WS2 heterostructure form with a crystallographic relationship (200)MoS2//(101)WS2. Moreover, it is worth noting that 2D materials in their single-phase form are often considered to be thermally stable. Nevertheless, in their heterostructure form, they may suffer from thermal defects and induced stresses that could affect their electronic and optical properties. In the current work, these inherent stresses have not been evaluated as well as their impact on the heterostructure physical properties. It is believed that the effects of these stresses on the heterostructure structural and electronic properties cannot be neglected.
To precisely investigate the chemical composition of the fabricated samples, we have conducted XPS analyses, as shown in Figure 6.
The XPS survey scan of the MoS2 sample, shown in Figure 6a, indicates the presence of MoS2 constituting elements. This figure shows strong peaks for Mo 3d and S 2p orbitals. The peak at 227.02 eV corresponds to the S 2s peak. The two strong peaks at 229.85 eV and 232.99 eV are attributed to Mo4+ 3d5/2 and 3d3/2 (Figure 6b). A small peak appearing at 236.2 eV indicates a minor oxidation of the Mo material [33]. For Sulfur, S 2p peaks are recorded and shown in Figure 6c. The two strong peaks at 162.54 eV and 163.73 eV are attributed to S2− 2p3/2 and p1/2 states. Regarding the WS2, the XPS survey scan (Figure 6d) indicates the presence of W and S in the material, translated by the strong peaks for W 4f7/2 and W 4f5/2 appearing at 33.28 eV and 35.43 eV, respectively (Figure 6e). An additional peak appears at 38.72 eV that could be attributed to W p3/2. The S peaks appearing at 162.88 eV and 164.07 eV are due to the S2− 2p3/2 and p1/2 states, respectively (Figure 6f). Moreover, the XPS survey scan obtained from the MoxW1-xS2 heterostructure sample is provided in Figure 6g, indicating the presence of W, Mo, and S in the heterostructure. The Mo 3d peaks are found at 229.79 eV and 232.95 eV (Figure 6h), and the S 2s peak is found at 227.2 eV. Additionally, strong peaks of W appear at 33.3 eV and 35.42 eV, corresponding to W 4f7/2 and W 4f5/2. The peak at 35.91 eV could be attributed to W-O bond, while the 38.80 eV peak is attributed to W 5p3/2 (Figure 6i). The S peaks appearing at 162.84 eV and 164.0 eV correspond to the S2− 2p3/2 and p1/2 states, respectively (Figure 6j).

3.2. Optical Properties

Density functional theory (DFT) calculations, using generalized gradient approximation (GGA) and Perdew Burke Ernzerhof (PBE) methods (implemented in Quantum Espresso) were first used to estimate the structural optimizations and electronic attributes (see Supplementary Materials). We have summarized in Table 1 the crystal structures of MoS2 and WS2, as well as the cell parameters, the cutoff wave function and charge densities for both materials implemented in DFT calculations. For all configurations (monolayer and bilayer), the cell parameters and atomic locations were fully relaxed using the Broyden-Fletcher-Goldfarb-Shanno (BFGS) approach until the remaining force on each atom was less than 10−3 Ryd/Bohr (see Supplementary Materials).
To simulate monolayers in our calculations, a vacuum space of 15 Å was created along both sides of the z-axis to isolate the crystal and prevent interactions between the adjacent layers. For the sampling of the Brillouin zone, a Monkhorst-Pack technique is used, with k-point meshes of 9 × 9 × 2 for the bulk and 9 × 9 × 1 for the monolayer and bilayer structures. The generated optimal structure was utilized to compute the band structures for various setups. The results revealed a transition from an indirect bandgap to a direct bandgap as the structure changed from bulk to monolayer for both WS2 and MoS2. Similar results have been reported [29,34,35,36]. The band structure of the MoS2 and WS2 layers were well conserved, as seen in Figure 7.
The conduction band minimum (CBmin) and the valence band maximum (VBmin) of the MoS2 and WS2 monolayers are positioned at the K point, respectively. MoxW1-xS2 is an indirect semiconductor with a 1.45 eV indirect bandgap. Unlike their homogeneous bilayers’ counterparts, the CBmin of MoxW1-xS2 heterostructures is positioned at the K point, whilst the VBmin is located at the point Γ. Through van der Waals interactions, the MoS2 and WS2 monolayers produce an atomically sharp type-II heterointerface, which may be favorable for electron–hole pair separation. Free electrons and holes will spontaneously separate in a type II heterostructure, which is useful for optoelectronics and solar energy conversion applications [37,38,39].
Furthermore, the optical reflectance of all samples was measured at room temperature in the wavelength range 400–800 nm as shown in Figure 8a.
MoS2 shows the lowest reflectance compared to the other samples, with the presence of both excitons, A and B, clearly visible at 636 nm and 688 nm positions, respectively [40,41]. This is due to the high optical absorption of the vertical morphology of the MoS2 nanosheets with high specific area and light trapping via the multiple scattering effects [42]. On the other hand, WS2 exciton appears clearly at 620 nm, showing the highest reflectance caused by its planar morphology. Finally, the reflectance of the MoxW1-xS2 sample shows a mixed behavior between MoS2 and WS2 with an enhancement of the MoS2 excitons. This validates the successful fabrication of the heterostructure as confirmed by the Raman spectroscopy and HRTEM analyses discussed earlier.
To obtain the optical bandgap, the reflectance measurements recorded for all investigated samples are implemented in Kubelka-Munk model as per the following equation [43]:
F ( R ) = K S = ( 1 R ) 2 2 R
where K represents the molar absorption coefficient, S is the scattering factor, and R is the reflectance. Our results show that both MoS2 and WS2 exhibit a bandgap of 1.77 eV and 1.85 eV, respectively, approximately equivalent to bandgap obtained using DFT calculations. In contrast, the MoxW1-xS2 sample shows a low bandgap of 1.63 eV compared to DFT calculations, which is probably due to the heterostructure construction and implementation in DFT that does not seem to reproduce the effective form of the heterostructure.

3.3. Photoresponse Measurements

To evaluate the optoelectronic properties of the MoS2, WS2, and MoxW1-xS2 nanocomposite films, we deposited a pair of Au electrodes onto the device surface, as illustrated in Figure 9. The electrical measurements were conducted at room temperature under dark and illumination conditions using a halogen lamp (70 mW/cm2), and under different excitation wavelengths ranging from 400 nm to 700 nm. The effective detection area of the samples was 0.075 cm2.
The J-V curves were collected using a voltage sweep program −/+ 5 V at 0.1 V step. The JPh at 5 V bias was subsequently computed using the following formula:
J P h   ( m A / c m 2 ) = I l i g h t I d a r k A
where Ilight and Idark represent the current obtained in the light and the dark conditions, respectively. A is the active detection area.
The photoresponse (P) of our samples was computed using the following equation:
P ( % ) = 100 I l i g h t I d a r k I d a r k
The responsivity (Rλ) and the relative detectivity (D*) [17] of the photodetectors were obtained using the following equations:
R λ = I p h P l i g h t
D * = R λ ( 2 q I d a r k ) 1 2
where q is the absolute value of an electron charge (1.6 × 10−19 Coulombs), Rλ is the responsivity given in units of mA W−1, and D* is the relative detectivity given in units of Jones.
In Figure 10, compared to the other samples, MoS2 exhibits the highest JPh achieving 4.8 mA/cm2, compared to the other samples, while WS2 and the heterostructure MoxW1-xS2 have shown lower values of 0.8 mA/cm2 and 3.7 mA/cm2, respectively. Moreover, the highest photoresponse is also achieved by MoS2 ~ 6.8 × 10 4 %, while WS2 exhibits the lowest one of 1.5 × 103%. The photoresponse of MoxW1-xS2 heterostructure of ~5.8 × 103% is similar to previously reported values [44]. This strong photoresponse is due to the high optical absorbance of the vertical MoS2 nanosheets, known for possessing a high ability to capture light and a quick charge transfer [45] (e.g., Figure 10a–c).
For MoS2, we obtain maximum values of Rλ and D*, respectively at 68 mA W−1 and 6.3 × 103 Jones, with a 5 V bias voltage. It is worth noting that higher Rλ values were also reported [18,19,46,47], however in those works the considered active area was extremely smaller (~10−7 cm2) and a high applied bias was considerably higher (~50 V) compared to our present study. Instead, our findings concur that WS2 and MoxW1-xS2 heterostructure exhibit Rλ and D* of 8.9 mA W−1, 2.1 × 1010 Jones and 47.4 mA W−1, 1.4 × 1011 Jones, respectively compared to MoS2.
For further examinations of the photoresponse of our samples, we conducted a series of photoresponse measurements under monochromatic light excitations in the range between 400 nm and 700 nm wavelengths. Figure 11d shows the relative detectivity D* computed at various wavelengths using the above-mentioned formula (Equations (4) and (5)), which is in agreement with the measured values in Figure 10d. From the latter, one can notice that Rλ and D* are decreasing from 77.2 to 10.9 mA W−1 and from 7.2 × 1011 to 1.8 × 1010 Jones, respectively, with an increasing excitation wavelength from 400 nm to 700 nm.
The maximum responsivity Rλ of 77.2 mA W−1 and the relative detectivity D* of 7.2 × 1011 Jones, were obtained at 400 nm excitation for MoS2 sample as reported elsewhere [18,19,46,47].
To further correlate our investigation with existing works, we conducted a survey of available data for sole MoS2, WS2 and the heterostructure made out of these compounds. The survey is summarized in Table 2.
The photodetection obtained for MoS2 shows similar performances in terms of detectivity with slightly better performance of our CVD-fabricated MoS2 compared to the sample processed by PLD. However, there is a difference in responsivity, mainly attributed to applied high bias voltages (~50 V) and very low active area used 10−7 cm2, compared to our active area of 10−2 cm2. This shows that our one-step fabricated MoS2 has exhibited high photodetection performances despite the larger effective area used. To the best of our knowledge, no previous results were reported on WS2 better than our results on CVD-grown samples achieving a responsivity of 11.8 mA W−1 and a detectivity of 1010 Jones, obtained at a very low bias voltage of 5 V. The control of CVD parameters to grow a high quality MoxW1-xS2 monolayer was already reported [28]. Similarly, we have fabricated the MoxW1-xS2 using a one-step CVD fabrication method achieving a responsivity and a detectivity of 47 mA W−1 and 1011 Jones, respectively. Similar results were reported before using a mixture of MoS2/WS2 monolayers and graphene or using MoS2 core shell containing WS2. We believe that our results indicate that our processing route consisting of a one-step and low-cost CVD fabrication technique of MoS2, WS2 and MoxW1-xS2 hold a strong promising potential for the future development of scalable photodetector devices.

4. Conclusions

A MoxW1-xS2 heterostructure was successfully synthesized by a one-step CVD route. The high purity and high quality of the samples have been confirmed by multiple characterization techniques. The incorporation of the MoS2, WS2 and the MoxW1-xS2 heterostructure into photoconductive devices demonstrated a promising potential of these compounds to be used as broadband photodetectors. In particular, the MoxW1-xS2 heterostructure has achieved a responsivity of 47.4 mA W−1 and a relative detectivity of 1.4 × 1011 Jones under visible light excitation ranging from 400 to 700 nm.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13010024/s1, Figure S1: Crystals configuration used in DFT for (a) MoS2, (b) WS2 and (c) MoxW1-xS2; Figure S2: Bandgap computed by DFT simulations (a) direct bandgap for monolayer MoS2, (b) indirect bandgap for bilayer MoS2; Figure S3: Bandgap obtained using Kubelka-Munck model (a) direct bandgap for monolayer MoS2, (b) indirect bandgap for bilayer MoS2; Figure S4: Bandgap computed using DFT simulations(a) direct bandgap for monolayer WS2, (b) indirect bandgap for bilayer WS2; Figure S5: Bandgap obtained using Kubelka-Munck model (a) direct bandgap for monolayer WS2, (b) indirect bandgap for bilayer WS2.

Author Contributions

M.A.Q. and M.J. conceived the study; M.A.Q., N.S.R. and A.K. carried out the experimental and theoretical investigations; M.A.Q., A.K., M.E.M., G.M., C.K., A.B. and M.J. analyzed the data. All authors contributed equally on writing, editing and reviewing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Authors acknowledge the financial support from Technology Innovation Institute under grant number TII/DERC/2091/2020 (Ref. UPJV-2021-DR-83).

Data Availability Statement

Data would be made available upon request to the corresponding author.

Acknowledgments

This work is a research collaboration between the University of Picardie Jules Verne (UPJV), France and Technology Innovation Institute (TII), UAE.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hui, R. Photodetectors. In Introduction to Fiber-Optic Communications; Elsevier: Amsterdam, The Netherlands, 2020; pp. 125–154. [Google Scholar]
  2. Malik, M.; Iqbal, M.A.; Choi, J.R.; Pham, P.V. 2D Materials for Efficient Photodetection: Overview, Mechanisms, Performance and UV-IR Range Applications. Front. Chem. 2022, 10, 905404. [Google Scholar] [CrossRef] [PubMed]
  3. Sciuto, A.; Roccaforte, F.; Raineri, V. Electro-Optical Response of Ion-Irradiated 4H-SiC Schottky Ultraviolet Photodetectors. Appl. Phys. Lett. 2008, 92, 093505. [Google Scholar] [CrossRef]
  4. Li, L.; Chen, H.; Fang, Z.; Meng, X.; Zuo, C.; Lv, M.; Tian, Y.; Fang, Y.; Xiao, Z.; Shan, C.; et al. An Electrically Modulated Single-Color/Dual-Color Imaging Photodetector. Adv. Mater. 2020, 32, 1907257. [Google Scholar] [CrossRef] [PubMed]
  5. Tyagi, D.; Wang, H.; Huang, W.; Hu, L.; Tang, Y.; Guo, Z.; Ouyang, Z.; Zhang, H. Recent Advances in Two-Dimensional-Material-Based Sensing Technology toward Health and Environmental Monitoring Applications. Nanoscale 2020, 12, 3535–3559. [Google Scholar] [CrossRef] [PubMed]
  6. Arams, F.R. Photodetectors for Optical Communication Systems. Proc. IEEE 1970, 58, 1466–1486. [Google Scholar] [CrossRef]
  7. Rogalski, A.; Antoszewski, J.; Faraone, L. Third-Generation Infrared Photodetector Arrays. J. Appl. Phys. 2009, 105, 4. [Google Scholar] [CrossRef] [Green Version]
  8. Xi, F.; Wang, H.; Xiao, D.; Dubey, M.; Ramasubramaniam, A. Two-Dimensional Material Nanophotonics. Nat. Photonics 2010, 4, 882. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, L.; Meric, I.; Huang, P.Y.; Gao, Q.; Gao, Y.; Tran, H.; Taniguchi, T.; Watanabe, K.; Campos, L.M.; Muller, D.A.; et al. One-Dimensional Electrical Contact to a Two-Dimensional Material. Science 2013, 342, 614–617. [Google Scholar] [CrossRef] [Green Version]
  10. Koski, K.J.; Cui, Y. The New Skinny in Two-Dimensional Nanomaterials. ACS Nano 2013, 7, 3739–3743. [Google Scholar] [CrossRef]
  11. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and Optoelectronics of Two-Dimensional Transition Metal Dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. [Google Scholar] [CrossRef]
  12. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-Layer MoS2 Transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef] [PubMed]
  13. Lee, E.W.; Lee, C.H.; Paul, P.K.; Ma, L.; McCulloch, W.D.; Krishnamoorthy, S.; Wu, Y.; Arehart, A.R.; Rajan, S. Layer-Transferred MoS2/GaN PN Diodes. Appl. Phys. Lett. 2015, 107, 103505. [Google Scholar] [CrossRef] [Green Version]
  14. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.Y.; Galli, G.; Wang, F. Emerging Photoluminescence in Monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef] [PubMed]
  15. Ponomarev, E.; Gutiérrez-Lezama, I.; Ubrig, N.; Morpurgo, A.F. Ambipolar Light-Emitting Transistors on Chemical Vapor Deposited Monolayer MoS2. Nano Lett. 2015, 15, 8289–8294. [Google Scholar] [CrossRef] [Green Version]
  16. George, A.; Fistul, M.V.; Gruenewald, M.; Kaiser, D.; Lehnert, T.; Mupparapu, R.; Neumann, C.; Hübner, U.; Schaal, M.; Masurkar, N.; et al. Giant Persistent Photoconductivity in Monolayer MoS2 Field-Effect Transistors. npj 2D Mater. Appl. 2021, 5, 15. [Google Scholar] [CrossRef]
  17. 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]
  18. Xie, Y.; Zhang, B.; Wang, S.; Wang, D.; Wang, A.; Wang, Z.; Yu, H.; Zhang, H.; Chen, Y.; Zhao, M.; et al. Ultrabroadband MoS2 Photodetector with Spectral Response from 445 to 2717 Nm. Adv. Mater. 2017, 29, 1605972. [Google Scholar] [CrossRef]
  19. Li, S.; Chen, X.; Liu, F.; Chen, Y.; Liu, B.; Deng, W.; An, B.; Chu, F.; Zhang, G.; Li, S.; et al. Enhanced Performance of a CVD MoS2 Photodetector by Chemical in Situ N-Type Doping. ACS Appl. Mater. Interfaces 2019, 11, 11636–11644. [Google Scholar] [CrossRef]
  20. Magnozzi, M.; Pflug, T.; Ferrera, M.; Pace, S.; Ramó, L.; Olbrich, M.; Canepa, P.; Ağircan, H.; Horn, A.; Forti, S.; et al. Local Optical Properties in CVD-Grown Monolayer WS2 Flakes. J. Phys. Chem. C 2021, 125, 16059–16065. [Google Scholar] [CrossRef]
  21. Zhu, Z.Y.; Cheng, Y.C.; Schwingenschlögl, U. Giant Spin-Orbit-Induced Spin Splitting in Two-Dimensional Transition-Metal Dichalcogenide Semiconductors. Phys. Rev. B Cover. Condens. Matter Mater. Phys. 2011, 84, 153402. [Google Scholar] [CrossRef]
  22. Li, J.; Han, J.; Li, H.; Fan, X.; Huang, K. Large-Area, Flexible Broadband Photodetector Based on WS2 Nanosheets Films. Mater. Sci. Semicond. Process 2020, 107, 104804. [Google Scholar] [CrossRef]
  23. Tan, H.; Fan, Y.; Zhou, Y.; Chen, Q.; Xu, W.; Warner, J.H. Ultrathin 2D Photodetectors Utilizing Chemical Vapor Deposition Grown WS2 with Graphene Electrodes. ACS Nano 2016, 10, 7866–7873. [Google Scholar] [CrossRef] [PubMed]
  24. Xue, Y.; Zhang, Y.; Liu, Y.; Liu, H.; Song, J.; Sophia, J.; Liu, J.; Xu, Z.; Xu, Q.; Wang, Z.; et al. Scalable Production of a Few-Layer MoS2/WS2 Vertical Heterojunction Array and Its Application for Photodetectors. ACS Nano 2016, 10, 573–580. [Google Scholar] [CrossRef]
  25. Ye, K.; Liu, L.; Liu, Y.; Nie, A.; Zhai, K.; Xiang, J.; Wang, B.; Wen, F.; Mu, C.; Zhao, Z.; et al. Lateral Bilayer MoS2–WS2 Heterostructure Photodetectors with High Responsivity and Detectivity. Adv. Opt. Mater. 2019, 7, 1900815. [Google Scholar] [CrossRef]
  26. Susarla, S.; Kutana, A.; Hachtel, J.A.; Kochat, V.; Apte, A.; Vajtai, R.; Idrobo, J.C.; Yakobson, B.I.; Tiwary, C.S.; Ajayan, P.M. Quaternary 2D Transition Metal Dichalcogenides (TMDs) with Tunable Bandgap. Adv. Mater. 2017, 29, 1702457. [Google Scholar] [CrossRef]
  27. Mann, J.; Ma, Q.; Odenthal, P.M.; Isarraraz, M.; Le, D.; Preciado, E.; Barroso, D.; Yamaguchi, K.; von Son Palacio, G.; Nguyen, A.; et al. 2-Dimensional Transition Metal Dichalcogenides with Tunable Direct Band Gaps: MoS2(1-x)Se2x Monolayers. Adv. Mater. 2014, 26, 1399–1404. [Google Scholar] [CrossRef]
  28. Wang, Z.; Liu, P.; Ito, Y.; Ning, S.; Tan, Y.; Fujita, T.; Hirata, A.; Chen, M. Chemical Vapor Deposition of Monolayer Mo1-XWxS2 Crystals with Tunable Band Gaps. Sci. Rep. 2016, 6, 21536. [Google Scholar] [CrossRef] [Green Version]
  29. Mouloua, D.; Kotbi, A.; Deokar, G.; Kaja, K.; el Marssi, M.; el Khakani, M.A.; Jouiad, M. Recent Progress in the Synthesis of MoS2 Thin Films for Sensing, Photovoltaic and Plasmonic Applications: A Review. Materials 2021, 14, 3283. [Google Scholar] [CrossRef]
  30. 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]
  31. Zhao, W.; Ghorannevis, Z.; Amara, K.K.; Pang, J.R.; Toh, M.; Zhang, X.; Kloc, C.; Tan, P.H.; Eda, G. Lattice Dynamics in Mono- and Few-Layer Sheets of WS2 and WSe2. Nanoscale 2013, 5, 9677–9683. [Google Scholar] [CrossRef]
  32. Lan, F.; Lai, Z.; Xu, Y.; Cheng, H.; Wang, Z.; Qi, C.; Chen, J.; Zhang, S. Synthesis of Vertically Standing MoS2 Triangles on SiC. Sci. Rep. 2016, 6, 31980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Rajput, N.S.; Kotbi, A.; Kaja, K.; Jouiad, M. Long-Term Aging of CVD Grown 2D-MoS2 Nanosheets in Ambient Environment. npj Mater. Degrad. 2022, 6, 75. [Google Scholar] [CrossRef]
  34. Zhao, W.; Pan, J.; Fang, Y.; Che, X.; Wang, D.; Bu, K.; Huang, F. Metastable MoS2: Crystal Structure, Electronic Band Structure, Synthetic Approach and Intriguing Physical Properties. Chemistry—Eur. J. 2018, 24, 15942–15954. [Google Scholar] [CrossRef]
  35. Wang, L.; Kutana, A.; Yakobson, B.I. Many-Body and Spin-Orbit Effects on Direct-Indirect Band Gap Transition of Strained Monolayer MoS2 and WS2. Ann. Phys. 2014, 526, L7–L12. [Google Scholar] [CrossRef] [Green Version]
  36. Brumme, T.; Calandra, M.; Mauri, F. First-Principles Theory of Field-Effect Doping in Transition-Metal Dichalcogenides: Structural Properties, Electronic Structure, Hall Coefficient, and Electrical Conductivity. Phys. Rev. B Cover. Condens. Matter Mater. Phys. 2015, 91, 155436. [Google Scholar] [CrossRef] [Green Version]
  37. Li, W.; Wang, T.; Dai, X.; Wang, X.; Zhai, C.; Ma, Y.; Chang, S.; Tang, Y. Electric Field Modulation of the Band Structure in MoS2/WS2 van Der Waals Heterostructure. Solid State Commun. 2017, 250, 9–13. [Google Scholar] [CrossRef]
  38. dos Santos, R.B.; Rivelino, R.; Mota, F.D.B.; Kakanakova-Georgieva, A.; Gueorguiev, G.K. Feasibility of Novel (H3C)NX(SiH3)3-n Compounds (X = B, Al, Ga, In): Structure, Stability, Reactivity, and Raman Characterization from Ab Initio Calculations. Dalton Trans. 2015, 44, 3356–3366. [Google Scholar] [CrossRef] [Green Version]
  39. Kakanakova-Georgieva, A.; Giannazzo, F.; Nicotra, G.; Cora, I.; Gueorguiev, G.K.; Persson, P.O.Å.; Pécz, B. Material Proposal for 2D Indium Oxide. Appl. Surf. Sci. 2021, 548, 149275. [Google Scholar] [CrossRef]
  40. Mukherjee, B.; Tseng, F.; Gunlycke, D.; Amara, K.K.; Eda, G.; Simsek, E. Complex Electrical Permittivity of the Monolayer Molybdenum Disulfide (MoS_2) in near UV and Visible. Opt. Mater. Express 2015, 5, 447. [Google Scholar] [CrossRef] [Green Version]
  41. Vaquero, D.; Clericò, V.; Salvador-Sánchez, J.; Martín-Ramos, A.; Díaz, E.; Domínguez-Adame, F.; Meziani, Y.M.; Diez, E.; Quereda, J. Excitons, Trions and Rydberg States in Monolayer MoS2 Revealed by Low-Temperature Photocurrent Spectroscopy. Commun. Phys. 2020, 3, 194. [Google Scholar] [CrossRef]
  42. Mouloua, D.; Rajput, N.S.; Blach, J.-F.; Lejeune, M.; El Marssi, M.; El Khakani, M.A.; Jouiad, M. Fabrication Control of MoS2/MoO2 Nanocomposite via Chemical Vapor Deposition for Optoelectronic Applications. Mater. Sci. Eng. B 2022, 286, 116035. [Google Scholar] [CrossRef]
  43. Landi, S.; Segundo, I.R.; Freitas, E.; Vasilevskiy, M.; Carneiro, J.; Tavares, C.J. Use and Misuse of the Kubelka-Munk Function to Obtain the Band Gap Energy from Diffuse Reflectance Measurements. Solid State Commun. 2022, 341, 114573. [Google Scholar] [CrossRef]
  44. Tan, H.; Xu, W.; Sheng, Y.; Lau, C.S.; Fan, Y.; Chen, Q.; Tweedie, M.; Wang, X.; Zhou, Y.; Warner, J.H. Lateral Graphene-Contacted Vertically Stacked WS2/MoS2 Hybrid Photodetectors with Large Gain. Adv. Mater. 2017, 29, 1702917. [Google Scholar] [CrossRef] [PubMed]
  45. Kufer, D.; Konstantatos, G. Highly Sensitive, Encapsulated MoS2 Photodetector with Gate Controllable Gain and Speed. Nano Lett. 2015, 15, 7307–7313. [Google Scholar] [CrossRef]
  46. Taffelli, A.; Dirè, S.; Quaranta, A.; Pancheri, L. MoS2 Based Photodetectors: A Review. Sensors 2021, 21, 2758. [Google Scholar] [CrossRef] [PubMed]
  47. Liu, X.; Hu, S.; Lin, Z.; Li, X.; Song, L.; Yu, W.; Wang, Q.; He, W. High-Performance MoS2 Photodetectors Prepared Using a Patterned Gallium Nitride Substrate. ACS Appl. Mater. Interfaces 2021, 13, 15820–15826. [Google Scholar] [CrossRef] [PubMed]
  48. Han, P.; Marie, L.S.; Wang, Q.X.; Quirk, N.; El Fatimy, A.; Ishigami, M.; Barbara, P. Highly Sensitive MoS2 Photodetectors with Graphene Contacts. Nanotechnology 2018, 29, 20LT01. [Google Scholar] [CrossRef] [Green Version]
  49. Kim, B.H.; Gu, H.H.; Yoon, Y.J. Large-Area and Low-Temperature Synthesis of Few-Layered WS2 Films for Photodetectors. 2D Mater. 2018, 5, 045030. [Google Scholar] [CrossRef]
  50. Zeng, L.; Tao, L.; Tang, C.; Zhou, B.; Long, H.; Chai, Y.; Lau, S.P.; Tsang, Y.H. High-Responsivity UV-Vis Photodetector Based on Transferable WS2 Film Deposited by Magnetron Sputtering. Sci. Rep. 2016, 6, 20343. [Google Scholar] [CrossRef] [Green Version]
  51. Lan, C.; Zhou, Z.; Zhou, Z.; Li, C.; Shu, L.; Shen, L.; Li, D.; Dong, R.; Yip, S.P.; Ho, J.C. Wafer-Scale Synthesis of Monolayer WS2 for High-Performance Flexible Photodetectors by Enhanced Chemical Vapor Deposition. Nano Res. 2018, 11, 3371–3384. [Google Scholar] [CrossRef]
Figure 1. Fabrication protocol used for the MoS2, WS2 and MoxW1-xS2 heterostructure samples.
Figure 1. Fabrication protocol used for the MoS2, WS2 and MoxW1-xS2 heterostructure samples.
Nanomaterials 13 00024 g001
Figure 2. Vibrational modes obtained under 532 nm laser excitation for the (a) MoS2, (b) WS2 and (c) MoxW1-xS2 samples.
Figure 2. Vibrational modes obtained under 532 nm laser excitation for the (a) MoS2, (b) WS2 and (c) MoxW1-xS2 samples.
Nanomaterials 13 00024 g002
Figure 3. XRD diagrams recorded for the as-grown (a) MoS2, (b) WS2 and (c) MoxW1-xS2.
Figure 3. XRD diagrams recorded for the as-grown (a) MoS2, (b) WS2 and (c) MoxW1-xS2.
Nanomaterials 13 00024 g003
Figure 4. Low and high magnification SEM images of (a,b) MoS2 vertically aligned nanosheets (c,d) WS2 stacked layers and (e,f) MoxW1-xS2 co-existent vertical and stacked layers.
Figure 4. Low and high magnification SEM images of (a,b) MoS2 vertically aligned nanosheets (c,d) WS2 stacked layers and (e,f) MoxW1-xS2 co-existent vertical and stacked layers.
Nanomaterials 13 00024 g004
Figure 5. Typical TEM microstructure at low and high magnifications. (a,b) Cross-sectional bright field TEM images of vertically oriented MoS2 flakes; (c,d) TEM image of in-plane MoS2; (eg) cross-sectional views of deposited multi-stacked WS2 layers; (hj) bright-field image of the MoxW1-xS2 heterostructure.
Figure 5. Typical TEM microstructure at low and high magnifications. (a,b) Cross-sectional bright field TEM images of vertically oriented MoS2 flakes; (c,d) TEM image of in-plane MoS2; (eg) cross-sectional views of deposited multi-stacked WS2 layers; (hj) bright-field image of the MoxW1-xS2 heterostructure.
Nanomaterials 13 00024 g005
Figure 6. XPS analyses for the (ac) MoS2, (df) WS2 and (gj) MoxW1-xS2 samples.
Figure 6. XPS analyses for the (ac) MoS2, (df) WS2 and (gj) MoxW1-xS2 samples.
Nanomaterials 13 00024 g006
Figure 7. DFT computed electronic bands structure for monolayer (a) MoS2, (b) WS2 and (c) MoxW1-xS2.
Figure 7. DFT computed electronic bands structure for monolayer (a) MoS2, (b) WS2 and (c) MoxW1-xS2.
Nanomaterials 13 00024 g007
Figure 8. Reflectance for all samples (a) (red: MoS2; blue: WS2 and green: MoxW1-xS2). The insets (bd) show the variation of the reflectance function versus energy determined by the Kubelka-Munk model. The respective bandgap energies are indicated by dashed lines.
Figure 8. Reflectance for all samples (a) (red: MoS2; blue: WS2 and green: MoxW1-xS2). The insets (bd) show the variation of the reflectance function versus energy determined by the Kubelka-Munk model. The respective bandgap energies are indicated by dashed lines.
Nanomaterials 13 00024 g008
Figure 9. Schematic of the photoresponse measurements set up used for all samples.
Figure 9. Schematic of the photoresponse measurements set up used for all samples.
Nanomaterials 13 00024 g009
Figure 10. (a) Photocurrent density measurements for (a) MoS2, (b) WS2, (c) MoxW1-xS2, and their respective (d) responsivity and relative detectivity obtained under halogen lamp illumination (70 mW/cm2).
Figure 10. (a) Photocurrent density measurements for (a) MoS2, (b) WS2, (c) MoxW1-xS2, and their respective (d) responsivity and relative detectivity obtained under halogen lamp illumination (70 mW/cm2).
Nanomaterials 13 00024 g010
Figure 11. Photocurrent density Jph measured for (a) MoS2, (b) WS2, (c) MoxW1-xS2, and (d) their respective relative detectivity under 400–700 nm wavelength excitations.
Figure 11. Photocurrent density Jph measured for (a) MoS2, (b) WS2, (c) MoxW1-xS2, and (d) their respective relative detectivity under 400–700 nm wavelength excitations.
Nanomaterials 13 00024 g011
Table 1. Crystal systems, cut off wave function and cell parameters implemented in DFT calculations.
Table 1. Crystal systems, cut off wave function and cell parameters implemented in DFT calculations.
MaterialCrystal SystemCut Off/ Wave Function (Ryd)Lattice Parameters (Å)
MoS2P63/mmc70/700a = b = 3.18; c = 15
WS2P63/mmc50/500a = b = 3.19; c = 15
MoxW1-xS2P63/mmc60/600a = b = 3.18; c = 15
Table 2. Photodetection performances of our samples with respect to the literature.
Table 2. Photodetection performances of our samples with respect to the literature.
MaterialFabricationBias (V)Power
Density (mW cm−2)
Active Area (cm2)ExcitationResponsivityDetectivity (Jones)Ref.
(nm)(mA W−1)
MoS2PLD1081.2 × 10−3445–271750.71.55 × 109[18]
MoS2-HfO2Exfoliation5-1.5 × 10−7550–8001047.7 × 1011[45]
MoS2/GaN substrateCVD202.94.7 × 10−446025 × 1035.6 × 108[47]
MoS2/Graphene CVD1016 × 10−6532–6331.4 × 1038.7 × 1014[48]
MoS2CVD5076.8 × 10−7450–7501059.4 × 1012[19]
MoS2CVD5707.5 × 10−2400–70077.27.2 × 1011[This study]
WS2Sputtering1014.99 × 10−7450–6350.44.4 × 106[49]
WS2Sputtering5--36553.3 × 1031.22 × 1011[50]
WS2-Graphene CVD52.5 × 1074 × 10−125323.5 × 1031.6 × 1010[23]
WS2Exfoliation511.7-532–10644.12.6 × 109[22]
WS2CVD100.071.7 × 10−65320.54.9 × 109[51]
WS2CVD5707.5 × 10−2400–70011.82.9 × 1010[This study]
MoS2/WS2 Graphene CVD101.7 × 1023.1 × 10−85322340 × 1034.1 × 1011[44]
MoS2/WS22-steps CVD4-1.2 × 10−54502.3 × 103-[24]
WS2/MoS22-steps CVD51.3 × 1036.2 × 10−7457–6716.7 × 1033.1 × 1013[25]
MoxW1-xS2CVD5707.5 × 10−2400–70047.41.4 × 1011[This study]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Al Qaydi, M.; Kotbi, A.; Rajput, N.S.; Bouchalkha, A.; El Marssi, M.; Matras, G.; Kasmi, C.; Jouiad, M. Photodetection Properties of MoS2, WS2 and MoxW1-xS2 Heterostructure: A Comparative Study. Nanomaterials 2023, 13, 24. https://doi.org/10.3390/nano13010024

AMA Style

Al Qaydi M, Kotbi A, Rajput NS, Bouchalkha A, El Marssi M, Matras G, Kasmi C, Jouiad M. Photodetection Properties of MoS2, WS2 and MoxW1-xS2 Heterostructure: A Comparative Study. Nanomaterials. 2023; 13(1):24. https://doi.org/10.3390/nano13010024

Chicago/Turabian Style

Al Qaydi, Maryam, Ahmed Kotbi, Nitul S. Rajput, Abdellatif Bouchalkha, Mimoun El Marssi, Guillaume Matras, Chaouki Kasmi, and Mustapha Jouiad. 2023. "Photodetection Properties of MoS2, WS2 and MoxW1-xS2 Heterostructure: A Comparative Study" Nanomaterials 13, no. 1: 24. https://doi.org/10.3390/nano13010024

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