Shraddha Hambirae,
Shashikant Shindeb,
H. M. Pathana,
Som Datta Kaushikc,
Chandra Sekhar Routd and
Shweta Jagtap
*e
aDepartment of Physics, Savitribai Phule Pune University, India
bMES's Department of Physics, Nowrosjee Wadia College, Pune 411001, India
cUGC-DAE Consortium for Scientific Research Mumbai Centre, BARC, Mumbai, India
dCentre for Nano and Material Sciences, Jain (Deemed-to-be University), Jain Global Campus, Ramanagaram, Bangalore, India
eDepartment of Electronic and Instrumentation Science, Savitribai Phule Pune University, India. E-mail: shweta.jagtap@gmail.com
First published on 1st August 2024
In this study, we explored the gas-sensing capabilities of MoS2 petaloid nanosheets in the metallic 1T phase with the commonly investigated semiconducting 2H phase. By synthesizing SnS2 nanoparticles and MoS2 petaloid nanosheets through a hydrothermal method, we achieve notable sensing performance for NO2 gas at room temperature (27 °C). This investigation represents a novel study, and to the best of our knowledge no, prior similar investigations have been reported in the literature for 1T@2HMoS2/SnS2 heterostructures for room temperature NO2 gas sensing. The formed heterostructure between SnS2 nanoparticles and petaloid MoS2 nanosheets exhibits synergistic effects, offering highly active sites for NO2 gas adsorption, consequently enhancing sensor response. Our sensor demonstrated a remarkable sensing response (Ra/Rg = 7.49) towards 1 ppm of NO2, rapid response time of 54 s, baseline recovery in 345 s, good selectivity and long-term stability, underscoring its potential for practical gas-sensing applications.
Several methods are available for detection of toxic gases, including chemiresistive, calorimetric, optical, and electrochemical techniques.5 However, few methods encounter challenges like limited accessibility, higher costs, and sensitivity constraints etc. Resistive gas sensors have garnered considerable interest due to their compact size, ease of fabrication, straightforward operation, and cost-effectiveness in manufacturing.6,7 Metal oxide semiconductors (MOS) have served as effective gas sensors in the past, but their reliance on high operating temperatures led to increased power consumption. Lately, two-dimensional (2D) layer-structured materials have gained significant interest across several domains.8 Notably, two-dimensional semiconductor materials distinguish themselves through their extensive band gap coverage, interface free of dangling bonds, high mobility, and rapid carrier transport. A high specific surface area and rich active sites make it possible to adsorbed a large number of gas molecules, which is one of the best properties that make 2D structures advantageous in gas sensing applications.9 In addition, the electrical characteristics of 2D materials can be modulated by altering the number of layers.10 These materials undoubtedly provide a reasonably good foundation for developing high-performance gas sensors.9 In particular, tin disulfide (SnS2), a characteristic 2D layered material with weak van der Waals interactions between its layers has received significant interest.11 The non-solubility and non-toxicity of SnS2 in aqueous solutions make it a very promising material. SnS2 is characterised by its unique structure at the atomic layer level, which results in an abundance of available functional active sites.12 Being a more electronegative nanomaterial, SnS2 accumulates chemically adsorbed molecules on its surface, generating an electrostatic potential. The depletion region of this surface potential corresponds to the size of the entire nanostructure. This unique attribute of n-type SnS2 nanosheets enhances gas adsorption sites, making it a strong option for designing and making high-performance NO2 gas monitors.12,13 Moreover, the two-dimensional surface of SnS2 is favourable for incorporation other semiconductors, which results in the formation of well-contacted heterojunctions that improve carrier conduction.14,15 It has been found that the formation of heterojunctions is a highly effective method for manipulating the electronic state of the SnS2 surface, thereby substantially enhancing its gas sensing characteristics.16 Comparing heterojunction materials to individual materials, they often perform better due to their diverse morphologies and band alignments. This improvement may be linked to the excellent heterointerface, which promotes fast charge transfer. Furthermore, the curved shape of the conductive band and the valence band in a typical heterojunction structure often causes the Fermi level to attempt equilibrium, which ultimately leads in the formation of a depletion layer. This effect reduces response and recovery times by directly contributing to high conductivity.16,17 The construction of two-dimensional heterojunction nanomaterials has been a popular and advanced method for designing gas sensors in recent years. Similar to SnS2, MoS2 possesses a traditional layer arrangement with weak van der Waals interactions.15 Therefore, the development of improved heterointerfaces by employing 2D SnS2 and MoS2 nanosheets stands as a promising approach for further improving electrical capabilities. Furthermore, the metastable 1T phase exhibits superior electrical conductivity for charge transfer and higher adsorption energy to NOx compared to the semiconducting 2H phase, suggesting its suitability for gas sensing applications. For example, MoS2 nanosheets coated with SnS2 nanoparticles (MoS2/SnS2) were developed by Jia-Bei Liu and colleagues by using the simple hydrothermal process and mechanical exfoliation method. The results of this study indicated that the SnS2 nanoparticles, which serve as an efficient antioxidative decoration, may increase the stability of MoS2 nanosheets. This provides a potential way to create high-stability NO2 gas sensors at ambient temperature.18
Considering the synergistic effects of heterostructures to boost gas sensing performance. This novel investigation explored the 1T@2H MoS2/SnS2 heterostructure by a simple hydrothermal approach to construct a two-dimensional layered 2H-SnS2 decorated on petaloid 1T-MoS2 nanosheets for room temperature NO2 gas sensing, marking the first instance of such research in the literature without prior similar studies. Utilizing a range of characterization techniques, comprehensive structural and functional investigations were conducted on the synthesized material. The hierarchical structure of the SnS2/MoS2 sensor proved superior to the pristine SnS2 sensor, exhibiting good response (Rg/Ra = 7.49) to 1 ppm NO2 at ambient temperature. The sensor is particularly noteworthy for its excellent selectivity and consistent repeatability.
C1V1 = C2V2 |
Therefore, to achieve a 1 ppm concentration in a 250 ml chamber, we need 0.25 ml of gas from the cylinder. We then used syringe to collect 0.25 ml of the gas from the 1000 ppm cylinder.
X-ray Photoelectron Spectroscopy (XPS) measurements were performed to determine the presence of the 1T@2H-MoS2 mixed phase and to examine the bonding arrangement, chemical composition, and electronic structure. The entire XPS spectrum of a SnS2-decorated MoS2 heterostructures in Fig. 2a shows Sn, Mo, S, and tiny quantities of carbon and oxygen. The presence of oxygen is a result of ambient oxygen adsorption on the composite surface. In Fig. 2b two prominent peaks, corresponding to Mo 3d5/2 and Mo 3d3/2 at around 228.9 and 231.84 eV, respectively, in the XPS spectra of the Mo 3d region for the SnS2/1T@2H-MoS2 composite show the existence of the 1T phase of MoS2. Two smaller 2H phase peaks with binding energies at 229.9 and 233.2 eV shift by about 1 eV to higher binding energies than 1T-MoS2.23–26 This change indicates the presence of a trace quantity of semiconducting MoS2 in the 2H phase. The +6-oxidation state of Mo (Mo6+ 3d3/2) is responsible for another weak peak at 236.9 eV which shows that pure MoS2 is partially oxidized. High-resolution S 2p spectra (Fig. 2c) show two peaks at 160.82 and 162.51 eV for S 2p1/2 and S 2p3/2 of S2− from MoS2 one unidentified peak were observed at 160.03. The SnS2/MoS2 heterostructures is further confirmed by two distinctive peaks in the S 2p spectra of SnS2/MoS2 that are located at 160.40 and 163.47 eV, respectively and correspond to S2− 2p1/2 and S2− 2p3/2 from SnS2.27 In the Sn 3d region from Fig. 2d, the peaks at 487.5 eV and 496.1 eV are typical Sn(IV) peaks. Interestingly, two more peaks at 485.2 eV and 493.6 eV suggest that Sn has changed in chemical states, which may be due to SnS2 and MoS2 interactions.28
![]() | ||
Fig. 2 (a) XPS survey spectra of SnS2/1T@2H-MoS2 heterostructures, high resolution XPS spectra of (b) Mo 3d. (c) S 2p. (d) Sn 3d. |
The microstructures of 1T@2H-MoS2 and its heterojunction with SnS2 were investigated using Field Emission Scanning Electron Microscopy (FESEM). Fig. 3a shows pristine 1T@2H-MoS2, which features a combination of small nanoparticles and petaloid nanosheets ranging from 1–3 μm in length. High-resolution Fig. 3b reveals small pores on the petaloid nanosheets and width of the nanosheets is around 30–80 nm. Pure SnS2, illustrated in Fig. S4(a),† consists of closely packed small nanoparticles. The pristine 1T@2H-MoS2 also displays some nanoparticle morphology, whereas in the composite, as shown in Fig. 3c, spherical SnS2 particles developed on the surface of the porous 1T@2H-MoS2 petaloid nanosheets. Fig. 3d shows that the significant aggregation of the SnS2 nanoparticles, resulting in the formation of a small sheet-like structure. Fig. 3e showed TEM image of 1T@2H-MoS2/SnS2 which consist of SnS2 and MoS2 nanoparticles which are around 10 to 20 nm in size, along with MoS2 nanosheets. Further to confirm the presence of nanoparticles and petaloid nanosheets in 1T@2H-MoS2/SnS2, EDS analysis was also carried out as shown in Fig. 3f which clearly shows presence of sulfur (S), molybdenum (Mo), and tin (Sn) elements. Individual elemental distribution is also shown in Fig. 3g–i for Mo, Sn and S respectively.
![]() | ||
Fig. 3 FESEM images of (a and b) 1T@2H-MoS2, (c and d) 1T@2H-MoS2/SnS2, (e) TEM image 1T@2H-MoS2/SnS2, (f–i) EDS mapping of 1T@2H-MoS2/SnS2 heterostructures. |
The 1T@2H-MoS2/SnS2 heterostructures demonstrates remarkable potential for NO2 gas sensing with its significantly faster response and recovery times compared to individual SnS2 and MoS2. In particular, the SnS2/1T@2H-MoS2 gas sensor with response times of about 54 seconds and recovery times of about 345 seconds. This performance is higher than MoS2 (response time 80 s and recovery time 102 s) and the SnS2 sensor (response and recovery time 75 s and 91 s respectively). The faster response and recovery of the heterostructures may be attributed to a synergistic effect between SnS2 and MoS2, leading to enhanced gas adsorption and improved charge transfer dynamics.
Table 1 presents comparison between the 1T@2H-MoS2/SnS2 heterostructures and other relevant sensors, including SnS2-based sensors and MoS2-based sensors reported in the literature for NO2 sensing. This comparative analysis aims to provide a comprehensive evaluation of sensing performance, taking into account crucial parameters such as response, response speed, recovery time, and operating temperature. Despite some literature reports indicating sensors with higher responses, our sensor stands out due to its room temperature sensing capability, exceptional selectivity, and stability. This recognition of performance at room temperature, along with its outstanding selectivity and stability, positions our sensor as a promising candidate for widespread use, even in scenarios where other sensors may have demonstrated greater responses according to existing literature.
Sensing material | NO2 concentration (ppm) | Working temperature (°C) | Response | Response/recovery time (s) | References |
---|---|---|---|---|---|
Au/SnS2/SnO2 heterojunctions | 8 | 80 | 22.3 | 174/359.6 | 31 |
MoS2/rGO composite | 3 | 160 | 23% | — | 32 |
3D-MoS2/PbS | 100 | RT | 25% | 30/235 | 33 |
SnS2/MoS2 | 10 | RT | 6.2 | 3.3/25.3 | 34 |
SnO2@SnS2 | 0.2 | RT | 5.5 | 950/1160 | 35 |
SnS2/SnS | 0.5 | RT | 2.5 | 375/1590 | 36 |
SnS2/vertical flakes | 50 | 120 | 1.64 | 41/379 | 37 |
1T@2H-MoS2/SnS2 | 1 | RT | 7.49 | 54/345 | This work |
To get more knowledge of the NO2 detecting capabilities of the gas sensors, the dynamic response curves were determined at NO2 concentrations of 1, 5, 10, 25, and 50 ppm, respectively. As shown in Fig. 5a, the response values of the SnS2/MoS2 sensors steadily rise with increasing NO2 concentrations, with the 1T@2H-MoS2/SnS2 gas sensor exhibiting the greatest response value over the entire test range. Furthermore, by plotting the logarithm of sensor response on the Y-axis and the logarithm of gas concentration on the X-axis in Fig. 5b linear correlation is observed across the entire concentration range. The SnS2/1T@2H-MoS2 sensor demonstrates linear relationship, supported by an R-squared value of 0.987.
The selectivity of the 1T@2H-MoS2/SnS2 sensor were carried out using 1000 ppm concentration of other interfering gases, such as, ammonia (NH3), carbon monoxide (CO), ethanol (C2H5OH), acetone (CH3COCH3), and carbon dioxide (CO2) and 100 ppm of NO2 gas. According to the findings, which are shown in Fig. 5c, resistance increases and decreases with gas exposure are represented by positive and negative response values, respectively. The trend of selectivity for individual gas molecules is determined by their oxidation potential and electrophilicity, which are impacted by partial charge transfer and adsorption energy. Compared to other gases (CO, C2H5OH, H2, C3H5OH), the sensor notably shows a greater detecting response to nitrogen-based compounds (e.g., NO2, NH3). This conclusion is consistent with work by Ray et al.,38 who used first-principal calculations to demonstrate that the gas adsorption energy of MoS2 for NO2 (268.6 MeV) is larger than that for NH3 (110.1 MeV), supporting the claim that adsorption and selectivity are related. Additionally, because of the physiosorbed paramagnetic NO2 molecules on the surface of SnS2, which provide a magnetic dipole and greater physical affinity, the device exhibits a higher sensing response to N-based substance.39,40 To explore the impact of relative humidity on the sensing properties of the sensor, 10 ppm NO2 was exposed to 1T@2H-MoS2/SnS2 under varying humidity levels (30% RH, 50% RH, 75% RH, 90% RH). The results from Fig. 5d indicate a slight decrease in the response value (13.25 ± 0.06) as the RH increases. This phenomenon can be attributed to the adsorption of water molecules, which reduces the active sites available for the target gas, consequently leading to a decline in sensor response. Therefore, it is reasonable to conclude that the influence of humidity is minimal, ensuring the reliability of the sensor in practical applications at room temperature. Furthermore, the sensor stability was examined over a twelve-week period. Results indicate that throughout this duration, the 1T@2H-MoS2/SnS2 sensor consistently shows sensing response ranging from 16.8 to 17.76 (in Fig. 6a). Therefore, development of a heterojunction between MoS2 and SnS2 considerably improves the sensors stability in air.
The findings demonstrate that 1T@2H-MoS2/SnS2 heterostructures exhibit superior gas response compared to their individual counterparts. Enhanced sensor response due to the combined action of the geometric (as shown in Fig. 7) and electronic aspects. The geometric aspect results in SnS2 nanoparticles on the MoS2 surface with more exposed active sites, while the electronic aspect creates a heterostructures at the interface. According to the UV-vis spectra, pure SnS2 shows a high absorption, especially in the ultraviolet region, with a notable decrease in absorption beyond 400 nm. SnS2 coated 1T@-2HMoS2, on the other hand, exhibits absorption from UV to near-infrared light regions. Which indicates that it is capable of absorbing superior amounts of ultraviolet and visible light. Because of the presence of the 1T-MoS2 metallic phase in the composite, there is a possibility that the increased absorption in the visible light range is associated with plasmon resonance absorption (illustrate in Fig. 6b). In case of semiconductor band gap is affected by a number of parameters, such as grain size, doping, and composition. A very small change was observed by the introduction of SnS2 in 1T@2H-MoS2/SnS2 than the pristine 1T@2H-MoS2. The relationships between (αhv)2 and photon energy are shown in Fig. 6c–e. The bandgap of SnS2 and 1T@2H-MoS2/SnS2 are 2.34 eV and 1.36 eV respectively. The lower band gap observed in the composite material than the pristine materials. It is often simpler for electrons to go from the valence band to the conduction band when the band gap is narrower. This promotes the growth of chemisorbed oxygen on the surface of SnS2/1T@2H-MoS2, ultimately resulting in an elevated reaction rate.42
To further understand the interface charge transfers on the sensor surface, Electrochemical impedance spectroscopy (EIS) with equivalent circuit (illustrate in Fig. S5 and S6†) was employed in presence of 10 ppm of NO2 gas. Typically, impedance spectra exhibit semi-circles at low-frequency regions (illustrate in Fig. 6e), reflecting the surface charge characteristics of the sensor material. Notably, 1T@2H-MoS2/SnS2 displays a smaller semicircle indicating lower electron transfer resistance compared to other devices. These results underscore the superior gas sensing properties of 1T@2H-MoS2/SnS2 as a sensing material. Consequently, the findings suggest that the resistance and charge transfer resistance of pure 1T@2H-MoS2 (5.48 MΩ) is higher than those of 1T@2H-MoS2/SnS2 (3.72 MΩ). In addition to this I–V characteristics of the synthesized materials obtained at room temperature under air are illustrated in Fig. 6g. Both forward and reverse-biased regimes of the corresponding I–V curves exhibit linearity, suggesting the ohmic nature of the contact. The average electrical resistances in air are 15 MΩ, and 6 MΩ of 1T@2H-MoS2 and 1T@2H-MoS2/SnS2 respectively. However, a recent study by L. Liu et al.34 observed lower conductivity (resulting in higher resistance) in the case of SnS2/MoS2-II (SMS-II) compared to the pure SnS2 and MoS2 in air at room temperature. Contrary to these findings, we observed higher conductivity in the case of the 1T@2H-MoS2/SnS2 heterostructure compared to the pristine materials. This enhanced conductivity in the heterostructure may be attributed to two main factors: firstly, the introduction of SnS2 into the MoS2 lattice can act as a dopant, introducing additional charge carriers. Secondly, the presence of SnS2 can modify the band structure of MoS2, leading to alterations in the electronic properties of the heterostructure. Furthermore, the addition of SnS2 to the MoS2 structure improves porosity and surface area, contributing to the enhanced sensing response. Moreover, SnS2 and MoS2 possessing work functions of 4.2–4.5 eV and 5.2–5.4 eV respectively, heterojunctions form between them. Electron flow occurs from SnS2 to MoS2 until their Fermi levels are balanced, creating an electron depletion layer on MoS2 and bending the energy band of SnS2.34 This leads to a change in electrical resistance in 1T@2H-MoS2/SnS2 heterostructures. Upon exposed to NO2 gas at the optimal operating temperature, trapped electrons are released back to the conduction band of 1T@2H-MoS2/SnS2 heterostructures due to the reaction between adsorbed O2 species and NO2 molecules. Consequently, the electrical resistance of 1T@2H-MoS2/SnS2 heterostructures significantly decreases, resulting in an enhanced gas sensing response.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra03194f |
This journal is © The Royal Society of Chemistry 2024 |