Defects in monolayer MoS2 (M-MoS2) can cause complex electronic states that significantly affect its optical and electrical properties. Understanding and describing the impact of these defects, especially the role of sulfur vacancy (Vs) in M-MoS2 when integrating them into practical technologies, is crucial. However, a significant challenge exists in precisely controlling Vs generation in M-MoS2. This article presents an in situ defect engineering procedure for M-MoS2, considering the influence of external stimuli. We investigated how Vs changes and its impact on the optoelectronic characteristics of M-MoS2 after it is directly exposed to various gas environments. A photodetector device was fabricated, which exhibited an outstanding responsivity of 1.02 × 104 A/W, a detectivity of 1.2 × 1012 Jones, and an ultralow noise equivalent power of 1.56 × 10−18 W Hz−1/2. When the device is exposed to a reducing gas (H2S) environment, the performance increases by 136%, and in an oxidizing gas environment (NO2), it decreases by 68% in terms of responsivity due to a change in the concentration of Vs. We studied the photoresponse characteristics of the device by using Vs as the key parameter. This research contributes to the field of defect engineering in M-MoS2, expanding our knowledge of gas–surface interactions and assisting in producing highly sensitive optoelectronic devices.
The integration of next-generation multifunctional semiconducting technology is based on the ability to tune the characteristics of 2D semiconductors effortlessly.1 Defect engineering is an established approach for tuning the electrical and optical characteristics of semiconductors.2 Although several studies have recently been conducted on 2D materials and their fascinating qualities, controlling the creation of defects and, consequently, the inherent characteristics of these materials is a critical step toward making them realistically feasible.2 Among the various 2D materials, transition metal dichalcogenides (TMDs), such as monolayer MoS2 (M-MoS2), are considered promising 2D materials for next-generation optoelectronic devices.3 It has an atomically thin structure with a high surface-to-volume ratio, high absorption, stability, flexibility, and an appropriate bandgap of ∼1.8 eV.4,5 These unique characteristics of M-MoS2 make it a substitute for semiconducting materials in fabricating photovoltaic cells, LEDs, and photodetectors (PDs) because they have different properties that suit them well with existing semiconductor technologies.6,7 The built-in defects in M-MoS2, specifically sulfur vacancies (Vs), significantly influence the performance of devices by altering carrier concentration, mobility, and interaction with external stimuli.8,9 Therefore, an investigation is essential to fully understand the impact of inherent defects, such as Vs, on the optical behavior of M-MoS2 under different gas circumstances.
There are several obstacles to achieving scalable and controlled defect engineering in M-MoS2. Various approaches have been reported to investigate the defects, including UV photon irradiation, chemical functionalization, electron/ion beam irradiation, laser processing, thermal annealing, and plasma treatment.2,10–12 Despite their effectiveness in inducing defects, they often lead to structural degradation and unpredictable properties in MoS2-based devices, increase production costs, and limit practical application. This study presents a novel method for defect engineering in M-MoS2 caused by exposure to gaseous environments. In particular, Vs can be controlled by the exposure of M-MoS2 to gaseous environments of H2S and NO2 without compromising its structural integrity. We also investigated the effect of external perturbation on dynamic optoelectronic properties and emphasized the significant role played by Vs in determining their photoresponse behavior. These gases introduce charge states while modifying the photoconductivity of the materials, thus influencing carrier generation and recombination rates. The findings could significantly advance the area of sensing based on 2D materials and enable the development of highly selective and sensitive sensors for next-generation sensing applications. It also offers a reliable and reproducible defect engineering solution for 2D materials.
Using a two-zone CVD technique, the M-MoS2 growth was optimized on a SiO2/Si substrate. The precursors such as molybdenum oxide (MoO3) and sulfur (S) powders were loaded in ceramic boats in different zones after flushing the tube at 1050 °C. Zone-1 was maintained at 160 °C (S), zone-2 was maintained at 780 °C (MoO3), and argon gas continuously flowed at 40 SCCM in the precisely controlled environment. For the optimization of M-MoS2, Raman spectroscopy (Jobin Yvon instrument) and atomic force microscopy (AFM; Multimode V Veeco system) were used to analyze the vibrational modes and high-resolution topographic images to confirm its formation. Furthermore, the surface composition of the thin film was examined using x-ray photoelectron spectroscopy (Scienta Omicron with an Al Kα source) by calibrating the C 1s position at 284.6 eV. The optical properties of the film are analyzed using the AvaSoft-8 UV–visible spectroscopy software (Avantes spectrometer) and the Edinburgh photoluminescence (PL) (FLS-980) D2D2 spectrometer, respectively. The device was fabricated on a single MoS2 flake using the lithography technique. The size of the triangular flake is 26 μm, and the spacing between the electrodes is 1 μm. We obtained the photoresponse of the PD device using a Keithley 2450 source meter as the data collection unit and a probe station equipped with an S-10 Triax cable. The current detection limit of the PD device is in the femtoampere range and has a precision of 100 ns in response time evaluation.
The schematic presentation and the temperature profile of the experimental setup used to synthesize M-MoS2 are shown in Figs. 1(a) and 1(b), respectively. The optical microscopic image of grown M-MoS2 is shown in Fig. 1(c), indicating well-defined, isolated triangular MoS2-flakes with sharp edges of high crystallinity. The triangular shape of the crystallites mirrors the threefold symmetry of MoS2, suggesting that they are single-crystalline.13 The top view of the Field Emission Scanning Electron Microscopy (FESEM) image of one flake and the optical image of the monolayer grown sample is shown in Figs. S1(a) and S1(b) of the supplementary material. The high-resolution topographic images obtained by AFM [Fig. 1(d)] reveal a uniform flake and thickness. The determined thickness is ∼0.7 nm. Moreover, the Raman spectra show peaks at 385.24 and 403.24 cm−1, revealing the presence of two distinct peaks with a Raman shift of 18.0 cm−1, corresponding to the E12g vibrational mode for the I-plane atom and the A1g mode for out-of-plane vibration [illustrated in Fig. 1(e)]. A satellite peak around 380 cm−1 can be observed, attributed to a localized vibrational mode of a longitudinal optical branch [LO(M)] caused by Vs. In addition, to support the claim of uniform growth, Raman mapping was observed, revealing the uniform distribution of MoS2 in the flake shown in Fig. 1(f).14
Furthermore, to analyze the chemical composition of M-MoS2, Mo 3d core level spectra are shown in Fig. 1(g). [The survey scan illustrated in Fig. S1(c) confirms the purity of M-MoS2.] The spectra exhibited doublets for Mo 3d5/2 (229.31 eV) and Mo 3d3/2 (232.42 eV), with a spin–orbit parting of 3.11 eV and distinctly separated Mo–S (Mo+4) bonding in MoS2. The evidence peak of the Mo–S bond can be observed at 226.39 eV. The S 2p core level [Fig. 1(d) of the supplementary material] confirms the oxidation state. In particular, the doublet 2p1/2 and 2p3/2 appear at 163.54 and 162.40 eV, respectively, with a peak splitting energy of 1.14 eV.14 The stoichiometry was calculated using the Mo 3d core level of spectra of XPS.15 [Equation (S1)] The estimated ratio of Mo and S, around 1:1.8, suggests a sulfur deficiency in the grown MoS2 flakes. Further analysis was conducted on the optical characteristic UV–Vis absorption spectra; distinct excitonic peaks at ∼695 and 610 nm were observed. These peaks represent the excitons A and B, confirming that M-MoS2 exhibits a semiconducting nature and a direct bandgap, which is essential for its application in PDs.16
Figure 1 confirms the growth of M-MoS2 with a sulfur deficiency that consists of a Mo atom sandwiched between two S atoms, as shown in Fig. 1S(e) of the supplementary material. To play with the defect, the device was fabricated, and the engineering of the device was observed in different stimuli. We have used the concentration of gases (H2S: 10 ppm, NO2: 10 ppm) and the power of the light source (10 mW/cm2 for 636 nm wavelength). Figure 2(a) depicts the schematic image [optical image of the device displayed in Fig. 1(f) of the supplementary material] of the device used to study the impact of various gases on the performance of the device. The I–V characteristics of gases without light illumination are shown in Fig. 2(b). The dark current (Id) at 0 V in the air was found to be 3.02 pA, while Id for the gases H2S and NO2 was 2.27 and 5.07 pA, respectively.
The changes in Id are caused by the interactions of H2S and NO2 gases with the surface of MoS2. Interactions with H2S reduce Id by the passivation of defect states, while NO2 increases the defects, which act as trap states, leading to higher Id.17 To understand the optoelectronic properties, the I–V in 626 nm light illumination under different gases is explained in Fig. S2 of the supplementary material. Furthermore, the transit response of the device using a 626 nm light source under different gas environments (H2S, air, and NO2) has been recorded. The power of the light was adjustable, and the applied bias was maintained at 0.5 V. Figure 2(c) shows Iph with time under various gas environments, and the source was periodically switched on/off at 10-s intervals. We have observed significant changes in Id at 0.5 V under air, H2S, and NO2; the values were 0.79, 0.58 and 1.4 nA, respectively. The current was increased by light illumination with air, H2S, and NO2 exposure. The values were 7, 9, and 4.8 nA, respectively, indicating responsiveness to different gases. Using Iph, the performance of the device was evaluated in terms of R, detectivity (D), and noise equivalent power (NEP).18,19 In the air environment, the device exhibited R, D, and NEP of 10 296 A/W, 1.2 × 1012 Jones, and 1.56 × 10−18 W Hz−1/2, respectively, as illustrated in Fig. 2(d) and Fig. S3 of the supplementary material (Table S1 of the supplementary material). When we exposed the device to H2S gas, the R was increased by 136%, and another parameter was also enhanced. In contrast, the device performance in the NO2 gas environment was decreased, as discussed in the supplementary material. In addition, we also investigated the influence of different gases, as shown in Fig. 2(e). Under H2S, air, and NO2 gases, the rise time was calculated as 615, 625, and 687 ms and the decay times were 846 ms, 1.5 s, and 1.58 s, respectively. The response time (τ) under air was taken as a baseline. When exposed to H2S, the device displayed significantly faster rise and decay times, indicating enhanced photocarrier generation and recombination due to decreased Schottky barrier height and increased electron concentration. However, in the case of the NO2 gas environment, slower rise and decay times were exhibited, suggesting slower photocarrier dynamics due to increased Schottky barrier height and reduced electron concentration. The fitted curve closely matched the experimental data, confirming the positive impact of H2S on enhancing the photodetection performance of the device.
Inducing defects in the device due to exposure to environmental gas happens in three stages. (1) H2S gas molecules come into contact with the surface of MoS2, where they can become physically or chemically adsorbed. At defect sites or edges with a higher binding energy, H2S molecules typically split into HS− and H+ ions. (2) The adsorbed H2S molecules act as electron donors, transferring electrons to the MoS2, which changes the electron concentration and impacts the sensitivity and response time of the PD. (3) Passivation or modification of trap states occurs through an interaction with H2S.17,20 In addition, H2S modifies the height of the Schottky barrier, influencing the carrier injection efficiency and affecting the performance of PD. On the contrary, exposure to NO2 induces p-type doping in MoS2 by accepting electrons from the surface, as shown in Fig. S4 of the supplementary material. This reduces the number of electrons and changes the performance of PD. The interaction of NO2 and H2S with M-MoS2 affects the carrier and changes the Schottky barrier height inversely, as shown in Figs. 3(e) and 3(f). Understanding the interactions of M-MoS2 with gas-specific changes and PD performance is important for future research.
In summary, we have explored the growth of M-MoS2 using the CVD technique and investigated the effect of defect engineering on the performance of PD. The fabricated device shows an excellent responsivity of 1.02 × 104 A/W and an ultralow NEP on the order of 10−18 W Hz−1/2 in the air environment. The impact of Vs and its interactions with H2S, NO2, and ambient conditions have been studied. It was observed that the device performance improved (136% in responsivity) in the H2S and decreased (68%) in the NO2 environment compared to the air. Our findings highlight the critical role of Vs in determining the sensitivity and selectivity of MoS2-based optical sensors. Exposure to H2S led to the passivation of Vs, resulting in enhanced PL intensity, faster carrier recombination times, and decreased Id due to reduced non-radiative recombination pathways. Meanwhile, exposure to NO2 caused PL quenching, slower recombination times, and increased dark current, which indicates enhanced non-radiative recombination and charge trapping. These changes underscore the sensitivity of M-MoS2-based devices to gas interactions and highlight their potential for sensing applications in various environments. The photocurrent and response time measurements also demonstrated the sensitivity of M-MoS2 to gas exposure, with faster response times and increased photocurrent upon H2S exposure, while NO2 exposure resulted in prolonged response times and decreased photocurrent. This research significantly advances the understanding of defect engineering in 2D materials for gas-sensing applications, providing detailed mechanistic insights into how Vs influences gas interactions. This work establishes a robust foundation for future research in optimizing MoS2-based devices for environmental monitoring and industrial applications.
SUPPLEMENTARY MATERIAL
The supplementary material provides additional data analysis and characterization of the gas-modulated optoelectronic characteristics of M-MoS2 for PD applications. The morphology and chemical analysis of M-MoS2 using a FESEM image, an optical image of the grown flakes, XPS spectra, the schematic of M-MoS2 structure, and an optical image of a fabricated device on a single flake are shown in Fig. S1, and Fig. S2 shows the I–V characteristics under different gas conditions. In addition, Figs. S3 and S4 display performance assessments and Schottky barrier changes. Tables S1 and S2 compare the device performance parameters and TRPL decay durations under various circumstances. This additional information provides a detailed explanation of the MoS2-based device’s development, composition, and performance, emphasizing its potential for complex sensing applications.
The authors thank the Director, CSIR-NPL, for his continuous support and encouragement. P.P. acknowledges CSIR, India, for the financial support under the CSIR-SRF fellowship. The authors also thank Dr. Preetam Singh, Dr. Jai Tawale, and Rajat Mukherjee for UV–visible, FESEM, and Raman spectroscopy measurements, respectively.
AUTHOR DECLARATIONS
Conflict of Interest
The authors affirm that they have no known financial or interpersonal conflicts that may have appeared to have influenced the research presented in this study.
Author Contributions
Pukhraj Prajapat: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Validation (equal); Writing – original draft (equal). Pargam Vashishtha: Conceptualization (supporting); Data curation (supporting); Formal analysis (equal); Investigation (equal); Methodology (supporting); Resources (supporting); Validation (equal); Writing – original draft (supporting). Govind Gupta: Conceptualization (equal); Funding acquisition (lead); Project administration (lead); Resources (lead); Supervision (lead); Validation (equal); Visualization (equal); Writing – review & editing (lead).
DATA AVAILABILITY
The data that support the findings of this study are available within the article and its supplementary material.