Next Article in Journal
Indium Oxide Decorated WS2 Microflakes for Selective Ammonia Sensors at Room Temperature
Next Article in Special Issue
Nanostructuring of SnO2 Thin Films by Associating Glancing Angle Deposition and Sputtering Pressure for Gas Sensing Applications
Previous Article in Journal
A Novel Electrochemical Sensing Platform for the Detection of the Antidepressant Drug, Venlafaxine, in Water and Biological Specimens
Previous Article in Special Issue
Non-Invasive Rapid Detection of Lung Cancer Biomarker Toluene with a Cataluminescence Sensor Based on the Two-Dimensional Nanocomposite Pt/Ti3C2Tx-CNT
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Performance Sulfur Dioxide Gas Sensor Based on Graphite-Phase Carbon-Nitride-Functionalized Tin Diselenide Nanorods Composite

1
College of Mechanical and Electrical Engineering, China University of Petroleum (East China), Qingdao 266580, China
2
Technology Inspection Center, Shengli Oilfield, Dongying 257000, China
3
College of Control Science and Engineering, China University of Petroleum (East China), Qingdao 266580, China
*
Author to whom correspondence should be addressed.
Chemosensors 2022, 10(10), 401; https://doi.org/10.3390/chemosensors10100401
Submission received: 3 September 2022 / Revised: 28 September 2022 / Accepted: 2 October 2022 / Published: 8 October 2022
(This article belongs to the Special Issue Chemical Sensors for Volatile Organic Compound Detection)

Abstract

:
In this paper, a composite of tin diselenide (SnSe2) functionalized by graphite-phase carbon nitride (g-C3N4) was successfully prepared by a hydrothermal method, and was characterized by X-ray diffraction (XRD), scanning electron microscope (SEM) and X-ray photoelectron spectroscopy (XPS). These microstructure characterization results verified the successful synthesis of a multilayer g-C3N4/rod-shaped SnSe2 composite. The gas sensitivity results showed that when the g-C3N4 ratio was 30%, the g-C3N4/SnSe2 composite sensor had the highest response (28.9%) at 200 °C to 20 ppm sulfur dioxide (SO2) gas, which was much higher than those of pristine g-C3N4 and SnSe2 sensors at the optimum temperature. A series of comparative experiments proved that the g-C3N4/SnSe2 composite sensor demonstrated an excellent response, strong reversibility and good selectivity for ppm-level SO2 gas detection. The possible SO2 sensing mechanism was ascribed to the heterostructure between the n-type SnSe2 and n-type g-C3N4 nanomaterials. Furthermore, we also proposed the influence of the special structure of the g-C3N4 functionalized SnSe2 composite on the gas-sensing characteristics.

1. Introduction

As an important indicator of air pollution, sulfur dioxide (SO2) is an irritating, highly toxic and colorless gas, which mainly comes from factory exhaust and automobile exhaust emissions [1,2]. When the emitted concentration of sulfur dioxide in air is too high, it is oxidized to sulfur trioxide and combined with water to form acid rain, which can destroy buildings, pollute the environment and reduce soil fertility [3,4]. In addition, human health is seriously threatened by some harmful gas species including sulfur dioxide gas. Respiratory diseases such as bronchitis and asthma can result from the excessive inhalation of sulfur dioxide gas [5]. There are reports that the human-permissible exposure limit for SO2 gas is 5 ppm and the long-term exposure limit is 2 ppm [6]. Therefore, the development of portable, efficient and reliable gas sensors can be used to monitor the composition and concentration of environmental gas pollutants in the atmosphere, which can help people to deal with dangerous gases in time, and further protect social safety and human health.
Recently, two-dimensional (2D) layered inorganic materials (such as graphene and transition metal materials) have attracted attention due to their unique crystal structures and characteristics. Among them, two-dimensional transition metal dichalcogenides (TMDs) are ideal materials for preparing field effect transistors [7], photodetectors [8] and electronic devices [9] because of their unique electronic, magnetic, optical and mechanical properties. Due to their planar crystal structure, high specific surface area and physical affinity, they also have unique advantages in sensing applications [10]. According to the reports from a few studies, two-dimensional transition metal compounds such as molybdenum disulfide (MoS2) [11,12], tungsten disulfide (WS2) [13,14] and tin diselenide (SnSe2) [15,16] offer good gas sensing characteristics for detecting dangerous gases. SnSe2 is an n-type two-dimensional transition chalcogenide, and its gas-sensitive properties have been widely reported [15,16,17]. The results of a study reported by Moreira et al. on the ammonia (NH3) and nitrogen dioxide (NO2) gas sensitivity of tin-diselenide-based sensors revealed that these sensors had steady repeatability and long-term stability under UV radiation [17]. Zhang et al. fabricated coral-like tin diselenide/metal-organic frameworks (MOFs)-derived nanoflower-like tin dioxide (SnO2) heteronanostructures via a hydrothermal method. The SnSe2/SnO2 nanocomposite sensor exhibited excellent NO2-sensing performance at room temperature, which was significantly improved under UV illumination. The enhanced NO2 sensing performance was attributed to the formation of an n-n heterostructure and light-excited electrons [18]. By using the template-sacrificial approach, Wang et al. created rod-shaped SnSe2 and polyhedral zinc oxide (ZnO) composite nanostructures, and the ZnO/SnSe2 heterostructures exhibited an enhancement of carbon monoxide (CO)-sensing properties at room temperature [19]. Pan et al. reported a coral-like Au-modified SnSe2 Schottky-junction-based ammonia gas sensor and demonstrated good gas sensitivity to ammonia gas detection. In addition, the effect of Au modification on ammonia gas molecules adsorption was also investigated using a first-principles density functional theory (DFT). Because of these studies, it can be concluded that SnSe2 is a feasible material as building block for constructing high-performance gas sensors [20]. However, the use of a single SnSe2 material for gas sensitivity research has certain limitations, and the test results may show poor selectivity and low sensitivity. According to previous studies, the sensing characteristics of gas sensors can be improved by forming a heterojunction [21,22]. Graphite-phase carbon nitride (g-C3N4), as a common two-dimensional material, has a structure like graphene and provides more active sites for gas adsorption. The material has excellent chemical stability, good electron mobility and low cost. In addition, g-C3N4 can also promote the uniform dispersion of active ingredients, so it can be used as a stable catalyst carrier [23,24].
In this paper, a composite of SnSe2 functionalized by g-C3N4 was prepared via a hydrothermal method and served as the sensitive nanomaterial for SO2 gas sensing. The nanostructure of the g-C3N4/SnSe2 composite was characterized by X-ray diffraction (XRD), scanning electron microscope (SEM) and X-ray photoelectron spectroscopy (XPS). The characterization results verified the successful synthesis of a multilayer g-C3N4/rod-shaped SnSe2 composite. The as-prepared g-C3N4/SnSe2 composite sensor showed an excellent gas response and rapid adsorption/desorption ability towards SO2 under the optimal temperature of 200 °C, which was much higher than those of pristine g-C3N4 and SnSe2 sensors. A series of comparative experiments proved that the g-C3N4/SnSe2 composite sensor demonstrated an excellent response, strong reversibility and good selectivity for ppm-level SO2 gas detection. In this paper, two innovative two-dimensional materials, g-C3N4 and SnSe2, were used to discuss in detail the mechanisms that may improve the SO2 sensing performance, such as the synergistic interaction between g-C3N4 and SnSe2, and the effective structural features.

2. Experimental Section

2.1. Materials

Tin chloride dihydrate (SnCl2·2H2O), hydrazine hydrate (N2H4·H2O), selenium dioxide (SeO2), ethanol (CH3CH2OH) and graphitic carbon nitride (g-C3N4) were all supplied by Sinopharm Chemical Reagent Co. Ltd. (Shanghai, China).

2.2. Material Synthesis and Sensor Fabrication

The preparation process of the materials is shown in Figure 1. The SnSe2 was prepared by a hydrothermal reduction method. A total of 0.01 mol SnCl2·2H2O and 0.02 mol SeO2 were added into 70 mL deionized water and stirred for 30 min. A total of 10 mL of hydrazine hydrate was added into the above mixture solution, and then transferred to a 100 mL Teflon-lined stainless-steel autoclave, and hydrothermally treated at 180 °C for 24 h. The resulting precipitate was filtered, washed and dried at 60 °C for 8 h to obtain a black tin diselenide product.
A total of 280 mg of SnSe2 powder and 120 mg of g-C3N4 powder were dissolved in 20 mL of DI water. After vigorously stirring for 1 h, the g-C3N4 was effectively anchored on the surface of SnSe2. The resulting product was dried at 60 °C for overnight. Finally, the g-C3N4/SnSe2 composite was obtained. The composite materials with different ratios (0, 20, 30, and 50%) of g-C3N4 and SnSe2 were prepared by adjusting the quality ratio of g-C3N4 and SnSe2. In addition, the two materials were dispersed in deionized water in a certain proportion, and after a period of ultrasonic treatment, they were effectively combined through strong physical effects and interactions between charges.

2.3. Gas Sensor Fabrication

The structure illustration of the ceramic tube based SO2 gas sensor is shown in Figure 2. It consisted of an Al2O3 ceramic tube and a base. The ceramic tube was 4 mm long and 1.2 mm in diameter, and its surface was equipped with gold electrodes and two pairs of platinum wires for electrical signals. The heating resistor of the Ni–Cr alloy coil passed through a hollow ceramic tube for heating. The sensing layer materials were coated on the surface of the ceramic tube, and the electrodes were led out to complete the preparation of the sensor. After preparing the sensing film, the sensor was dried at 60 °C for 6 h and then aged at 200 °C for 24 h before the test to obtain good resistance stability. The SO2-sensing measurement was performed in a home-made gas sensing detection system [25] as shown in Figure 3. The sensor was placed in a home-made chamber, and the SO2 gas with different concentrations of 1–200 ppm was obtained by diluting 1000 ppm SO2 standard gas with high-purity air. The sensor resistance was measured with an Agilent 34970A digital multimeter and connected with a computer through RS-232 for data acquisition. The operation temperature of the sensor was controlled by an applied voltage to the Ni–Cr heating resistor. A steady power supply of GPD-4303S was employed for applying voltage for heating. The response of the sensor is defined as S = (RA − RG)/RA × 100% (RA: resistance in air; RG: resistance in SO2 gas).

3. Results and Discussion

3.1. Structure Characterization

The crystal structures of the SnSe2, g-C3N4 and g-C3N4/SnSe2 were characterized by X-ray powder diffraction (XRD, Rigaku D/Max-2550, Rigaku, Japan) with Cu Kα radiation (λ = 0.15418 nm). X-ray photoelectron spectroscopy (XPS, Thermo Scientific K-Alpha XPS spectrometer, Thermo Scientific, Waltham, MA, USA) was used to detect the chemical composition of the samples, and the morphology of pristine SnSe2, g-C3N4 and g-C3N4/SnSe2 nanocomposites were observed by a scanning electron microscope (SEM, Hitachi S-4800, Hitachi, Japan). The g-C3N4/SnSe2 sample with a ratio of 30% g-C3N4 was used for characterization.
The crystal phases of the as-prepared materials were identified by XRD analysis. As shown in Figure 4a, the peaks of SnSe2 (JCPDS card number 23-0602) located at 14.38°, 26.98°, 30.88°, 40.17°, 47.77°, 50.19°, 52.58°, 57.92°, 60.34°, 63.96° and 78.22° corresponded to the (001), (100), (011), (012), (110), (111), (103), (201), (004), (202) and (121) diffraction planes of the orthorhombic-phase SnSe2 structures, respectively [26]. The XRD pattern of g-C3N4 was well in accordance with the hexagonal crystal g-C3N4 (JCPDS Card no. 36-1451) [27]. The formation of the two diffraction peaks of (100) and (002) can be attributed to the in-plane structure stacking of the tri-s-triazine part and the in-plane stacking of the conjugated aromatic hydrocarbon system, respectively. There were corresponding characteristic peaks of g-C3N4 and SnSe2 in the XRD pattern of the g-C3N4/SnSe2 nanocomposite, and there were no other characteristic peaks, indicating the successful preparation of the SnSe2/g-C3N4 composite.
XPS is a measurement technique for detecting the elemental composition and chemical valence of materials. Figure 4b is the total spectrum of the g-C3N4/SnSe2 composite, and the existence of the four elements Se, Sn, C and N was confirmed from the peak positions in the figure. The Se 3D spectrum shown in Figure 4c is represented by two peaks with binding energies of 53.4 and 52.2 eV, corresponding to Se 3d5/2 and Se 3d3/2, respectively [28,29,30]. The Sn 3d spectrum shown in Figure 4d shows two distinctive peaks at energies of 486.3 and 495.2 eV, which correspond to Sn 3d5/2 and 3d3/2, respectively. Since there was an energy difference of 8.4 eV between the two peaks, the Sn4+ state can be confirmed [31]. Figure 4e is the characteristic peak of N 1s. Three asymmetric peaks at 398.3, 400.0 and 401.0 eV were attributed to C=N-C, N-(C)3 and C-N-H, respectively. Figure 4f shows the XPS spectrum of C 1s, where the characteristic peak at 284.6 eV was due to the formation of indefinite carbon absorption on the surface of the material, and the diffraction peak at 287.8 eV may be formed by N=C-N coordination [32].
SEM is a detection technology used to observe the micro-morphology of nanomaterials. Figure 5 shows the SEM characterization of the g-C3N4/SnSe2-sensitive material. Figure 5a shows the SEM characterization of the intrinsic SnSe2. The SnSe2 nanorods were composed of numerous nanoparticles, the cross-sectional radius and the length of which were about 300−500 nm and 3−5 μm, respectively. Figure 5b shows the SEM characterization image of the graphite-like-phase carbon nitride. The exfoliated g-C3N4 nanosheets had a 2D layered morphology, which was manifested as multiple thin layers stacking together. Figure 5c,d show the surface morphology of the g-C3N4/SnSe2 composite. It can be clearly observed that the nanorod-shaped SnSe2 and multilayer g-C3N4 grew and aggregated together well, which show s that the g-C3N4/SnSe2 composite was successfully synthesized.

3.2. SO2-Sensing Properties

The working temperature is an important key variable related to the sensing characteristics of the g-C3N4/SnSe2 sensor [33,34,35]. It affects the chemical adsorption and surface reaction of gas molecules. Therefore, to determine the optimal operating temperature as well as the optimum ratio of the two materials, we studied the responses of pristine g-C3N4, pristine SnSe2 and the x% g-C3N4/SnSe2 (x = 0, 20, 30, 50) film sensors to 20 ppm SO2 gas at different operating temperatures. As shown in Figure 6a, the operating temperatures and responses of the sensors presented a “triangular” shape. The optimal temperatures of g-C3N4, SnSe2 and g-C3N4/SnSe2 were 250 °C, 200 °C and 200 °C, respectively. When the g-C3N4 ratio was 30%, the response (28.9%) of g-C3N4/SnSe2 at 200 °C was the highest, which was much higher than those of the pristine g-C3N4 and SnSe2. The gas-sensing properties of the SnSe2, g-C3N4 and 30% g-C3N4/SnSe2 sensors were investigated by recording the changes in resistance when they were exposed to different concentrations (1−200 ppm) of SO2 gas at 200 °C, as shown in Figure 6b. When switched to air, the resistances of the three sensors returned to the base value in air. The response values of 30% g-C3N4/SnSe2 towards 1, 5, 10, 20, 50, 100, 200 ppm SO2 gas were, respectively, 8.82%, 15.59%, 20.22%, 28.52%, 35.09%, 41.54% and 44.34%, higher than those of the pristine SnSe2 and g-C3N4. Figure 6c shows the fitting curves of the three sensors between sensitivity (Y) and concentration of SO2 (X). The fitting functions of g-C3N4, SnSe2 and g-C3N4/SnSe2 sensors were Y = 14.89 − 11.82 × 0.98X, Y = 22.76 − 17.85 × 0.95X and Y = 42.16 − 33.15 × 0.96X, respectively. The R2 of the fitting curve of the g-C3N4/SnSe2 sensor was 0.975. Figure 6d shows the response/recovery curves of the g-C3N4/SnSe2 sensor toward the desired concentrations of SO2 gas. The sensor exhibited stable response and recovery behaviors.
The repeatability of the g-C3N4/SnSe2 composite sensor against a gas with a concentration of 50 ppm SO2 at 200 °C is examined in Figure 7a. For each run, the resistance could fully recover its initial state and changes from 4.5 MΩ to 2.9 MΩ, showing a good reproducibility. Figure 7b shows the response/recovery curves of the g-C3N4, SnSe2, and g-C3N4/SnSe2 composite sensors exposed to 50 ppm SO2 gas. The response/recovery time of the g-C3N4/SnSe2 sensor was 22/24 s, while the values of the pristine g-C3N4 and SnSe2 sensors were 46/55 s and 26/82 s, respectively, suggesting that the g-C3N4/SnSe2 sensor had a faster detection rate towards SO2 gas compared with the pristine g-C3N4 and SnSe2 sensors at 200 °C. Excellent selectivity is also an important factor for nanomaterial-based gas sensors [36,37]. Therefore, we further studied the SO2 selectivity of the g-C3N4/SnSe2 composite sensor at 200 °C. The sensor was exposed to 50 ppm of various interfering gases, such as LPG, CO, CH4, H2S and H2. As shown in Figure 7c, the g-C3N4/SnSe2 composite sensor had the highest response to SO2 gas, indicating that the selectivity to SO2 gas was excellent. Additionally, the long-term stability of the g-C3N4/SnSe2 composite sensor for various SO2 gas concentrations (1, 10 and 50 ppm) at a working temperature of 200 °C was examined. The sensor’s response had no obvious changes in a period under the same experimental conditions and exhibited a good stability, as shown in Figure 7d.
Numerous studies have described the detection of SO2 gases utilizing a variety of sensitive materials up to this point. As far as we know, there have been no studies using g-C3N4/SnSe2 for SO2 gas detection. Table 1 lists the comparison of this work with previously reported works. The different SO2 gas sensors are compared in terms of sensing environment, response value and gas concentration. The comparison results showed that the as-prepared g-C3N4/SnSe2 sensor featured a higher response value and a lower operating temperature.

3.3. SO2 Gas-Sensing Mechanism

The sensing mechanism has been widely explained using the surface charge caused by adsorbed oxygen. For the pristine SnSe2 gas sensor, when exposed to air, oxygen molecules in air adsorb on the surface of the sensing material to form adsorbed oxygen molecules. The adsorbed oxygen molecules extract electrons from the conduction band of SnSe2 to form chemically adsorbed oxygen O at 200 °C. On the surface of the g-C3N4/SnSe2 sensing material, an electron depletion layer consequently forms. The following is a description of the reactions [57,58]:
O2 (gas) → O2 (ads)
O2 (ads) + 2e → 2O (ads)
When the reducing gas SO2 is introduced, the SO2 molecules adsorbed on the surface of the sensing material will further react with O ions and release electrons to the SnSe2 conduction band to form SO3, thereby increasing the number of charge carriers and reducing the resistance of the sensor. The specific reaction process is illustrated as follows [59]:
SO2 (ads) + O (ads) → SO3 + e
In our experiment, it is worth noting that the g-C3N4/SnSe2 sensor exhibited an improved response to SO2 gas than the pristine C3N4 and SnSe2 sensors. This phenomenon could be explained by two positive factors of g-C3N4 decoration. Firstly, the morphology, specific surface area and electrical properties of the material are important factors affecting its gas-sensing performance [60]. The addition of the layered two-dimensional structure of g-C3N4 increases the specific surface area of the composite material and provides more active sites, increasing the production of adsorbed oxygen. Another factor influencing the effectiveness of the gas sensing system is the n-n heterojunction created between g-C3N4 and SnSe2 [61,62,63]. Figure 8a shows the energy band structure of the heterojunction formed by the n-type SnSe2 nanorods and n-type g-C3N4 layered nanosheets in air, where the band gaps of SnSe2 and g-C3N4 were 1.37 and 2.7 eV, respectively. Their Fermi levels were different, and the work function (4.3 eV) of SnSe2 was lower than that (4.67 eV) of g-C3N4 [64]. When the two materials contact with different work functions, electrons will transfer from SnSe2 to g-C3N4 to reach a Fermi energy balance. Therefore, the n-n heterojunction is formed at the interface between SnSe2 and g-C3N4 in air [65]. When exposed to SO2 gas, SO2 molecules absorbed on the surface of the composite material react with O to generate electrons [66]. Therefore, the carrier concentration in the heterojunction increases by receiving electrons, resulting in a narrowing of the depletion layer at the interface of the two materials [61,67,68], which reduces the resistance of the sensor (Figure 8b). The built-in electric field generated by the nano-scale heterojunction can accelerate the separation process of electrons and facilitate electron transfer.

4. Conclusions

In this paper, a g-C3N4 functionalized SnSe2 composite thin film sensor was successfully prepared. XRD, XPS and SEM techniques were used to characterize the elements and structure of the g-C3N4/SnSe2 composite material. The optimal temperature (200 °C) of the g-C3N4/SnSe2 composite sensor was determined through gas-sensing experiments under different working temperatures. The SO2 gas-sensing performance for the g-C3N4/SnSe2 composite sensor was investigated at the optimal temperature. The experimental results showed that under the optimal temperature, the 30% g-C3N4/SnSe2 composite sensor gained the highest sensitivity to SO2 gas. The two-dimensional layered structure of the g-C3N4-modified n-type SnSe2 nanorods not only increased the specific surface area and the gas adsorption site of the g-C3N4/SnSe2 sensor, but also formed an n-n heterojunction between the g-C3N4 nanosheets and SnSe2 nanorods, which improved the sensing performance of the g-C3N4/SnSe2 sensor toward SO2 gas.

Author Contributions

Conceptualization, H.Z. and Q.P.; methodology, D.Z.; validation, H.Z. and Q.P.; formal analysis, H.Z. and Q.P.; investigation, Y.Z. (Yanting Zhang); data curation, H.Z. and Q.P.; writing—original draft preparation, Q.P.; writing—review and editing, Y.Z. (Yating Zhang) and D.Z.; visualization, Q.P. and Y.Z. (Yating Zhang); supervision, D.Z.; funding acquisition, D.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Original Innovation Special Project of Science and Technology Plan of Qingdao West Coast New Area, China (2020-85).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Das, S.; Rana, S.; Mursalin, S.; Rana, P.; Sen, A. Sonochemically prepared nanosized BiFeO3 as novel SO2 sensor. Sens. Actuators B 2015, 218, 122–127. [Google Scholar] [CrossRef]
  2. Jung, G.; Jeong, Y.; Hong, Y.; Wu, M.; Hong, S.; Shin, W.; Park, J.; Jang, D.; Lee, J. SO2 gas sensing characteristics of FET- and resistor-type gas sensors having WO3 as sensing material. Solid-State Electron. 2020, 165, 107747. [Google Scholar] [CrossRef]
  3. Ma, C.; Hao, X.; Yang, X.; Liang, X.; Liu, F.; Liu, T.; Yang, C.; Zhu, H.; Lu, G. Sub-ppb SO2 gas sensor based on NASICON and LaxSm1−xFeO3 sensing electrode. Sens. Actuators B Chem. 2018, 256, 648–655. [Google Scholar] [CrossRef]
  4. Liu, L.; Liu, S. Oxygen vacancies as an efficient strategy for promotion of low concentration SO2 gas sensing: The case of Au-modified SnO2. ACS Sustain. Chem. Eng. 2018, 6, 13427–13434. [Google Scholar] [CrossRef]
  5. Liu, F.; Wang, J.; Jiang, L.; You, R.; Wang, Q.; Wang, C.; Lin, Z.; Yang, Z.; He, J.; Liu, A.; et al. Compact and planar type rapid response ppb-level SO2 sensor based on stabilized zirconia and SrMoO4 sensing electrode. Sens. Actuators B Chem. 2020, 307, 127655. [Google Scholar] [CrossRef]
  6. Tong, P.; Hoa, N.; Nha, H.; Duy, N.; Hung, C.; Hieu, N. SO2 and H2S sensing properties of hydrothermally synthesized CuO nanoplates. J. Electron. Mater. 2018, 47, 7170–7178. [Google Scholar] [CrossRef]
  7. Goel, N.; Kumar, R.; Jain, S.; Rajamani, S.; Roul, B.; Gupta, G.; Kumar, M.; Krupanidhi, S. A high-performance hydrogen sensor based on a reverse-biased MoS2/GaN heterojunction. Nanotechnology 2019, 30, 314001. [Google Scholar] [CrossRef]
  8. Xu, T.; Han, Y.; Lin, L.; Xu, J.; Fu, Q.; He, H.; Song, B.; Gai, Q.; Wang, X. Self-power position-sensitive detector with fast optical relaxation time and large position sensitivity basing on the lateral photovoltaic effect in tin diselenide films. J. Alloys Compd. 2019, 790, 941–946. [Google Scholar] [CrossRef]
  9. Javed, Y.; Mirza, S.; Li, C.; Xu, X.; Rafiq, M. The role of biaxial strain and pressure on the thermoelectric performance of SnSe2: A first principles study. Semicond. Sci. Technol. 2019, 34, 055009. [Google Scholar] [CrossRef]
  10. Gu, D.; Wang, X.; Liu, W.; Li, X.; Lin, S.; Wang, J.; Rumyantseva, M.; Gaskovc, A.; Akbar, S. Visible-light activated room temperature NO2 sensing of SnS2 nanosheets based chemiresistive sensors. Sens. Actuators B Chem. 2020, 305, 127455. [Google Scholar] [CrossRef]
  11. Jaiswal, J.; Sanger, A.; Tiwari, P.; Chandra, R. MoS2 hybrid heterostructure thin film decorated with CdTe quantum dots for room temperature NO2 gas sensor. Sens. Actuators B Chem. 2020, 305, 127437. [Google Scholar] [CrossRef]
  12. Shen, J.; Yanga, Z.; Wang, Y.; Xu, L.; Liu, R.; Liu, X. The gas sensing performance of borophene/MoS2 heterostructure. Appl. Surf. Sci. 2020, 504, 144412. [Google Scholar] [CrossRef]
  13. Liu, D.; Tang, Z.; Zhang, Z. Comparative study on NO2 and H2S sensing mechanisms of gas sensors based on WS2 nanosheets. Sens. Actuators B Chem. 2020, 303, 127114. [Google Scholar] [CrossRef]
  14. Zhou, Q.; Zhu, L.; Zheng, C.; Wang, J. Nanoporous Functionalized WS2/MWCNTs Nanocomposite for Trimethylamine Detection Based on Quartz Crystal Microbalance Gas Sensor. ACS Appl. Mater. Interfaces 2021, 13, 41339–41350. [Google Scholar] [CrossRef] [PubMed]
  15. Pawar, M.; Kadam, S.; Late, D. High-performance sensing behavior using electronic ink of 2D SnSe2 nanosheets. Chem. Sel. 2017, 2, 4068–4075. [Google Scholar] [CrossRef]
  16. Wang, X.; Liu, Y.; Dai, J.; Chen, Q.; Huang, X.; Huang, W. Solution-processed p-SnSe/n-SnSe2 hetero-structure layers for ultrasensitive NO2 detection. Chem. Eur. J. 2020, 26, 3870–3876. [Google Scholar] [CrossRef]
  17. Moreira, O.; Cheng, W.; Fuh, H.; Chien, W.; Yan, W.; Fei, H.; Xu, H.; Zhang, D.; Chen, Y.; Zhao, Y.; et al. High Selectivity Gas Sensing and Charge Transfer of SnSe2. ACS Sens. 2019, 4, 2546–2552. [Google Scholar] [CrossRef] [PubMed]
  18. Li, T.; Zhang, D.; Pan, Q.; Tang, M.; Yu, S. UV enhanced NO2 gas sensing at room temperature based on coral-like tin diselenide/MOFs-derived nanoflower-like tin dioxide heteronanostructures. Sens. Actuators B Chem. 2022, 355, 131049. [Google Scholar] [CrossRef]
  19. Wang, D.; Zhang, D.; Pan, Q.; Wang, T.; Chen, F. Gas sensing performance of carbon monoxide sensor based on rod-shaped tin diselenide/MOFs derived zinc oxide polyhedron at room temperature. Sens. Actuators B Chem. 2022, 371, 132481. [Google Scholar] [CrossRef]
  20. Pan, Q.; Li, T.; Zhang, D. Ammonia gas sensing properties and density functional theory investigation of coral-like Au-SnSe2 Schottky junction. Sens. Actuators B Chem. 2021, 332, 129440. [Google Scholar] [CrossRef]
  21. Sun, K.; Zhan, G.; Chen, H.; Lin, S. Low-Operating-Temperature NO2 Sensor Based on a CeO2/ZnO Heterojunction. Sensors 2022, 21, 8269. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, K.; Liu, H.; Wang, J.; Feng, L. Synthesis and characterization of SnSe2 hexagonal nanoflakes. Mater. Lett. 2009, 63, 512–514. [Google Scholar] [CrossRef]
  23. David, S.P.S.; Veeralakshmi, S.; Priya, M.S.; Nehru, S.; Kalaiselvam, S. Room-temperature chemiresistive g-C3N4/Ag2ZrO3 nanocomposite gas sensor for ethanol detection. J. Mater. Sci.-Mater. El. 2022, 33, 11498–11510. [Google Scholar] [CrossRef]
  24. Pan, W.; Zhang, Y.; Yu, S.; Liu, X.; Zhang, D. Hydrogen sulfide gas sensing properties of metal organic framework-derived α-Fe2O3 hollow nanospheres decorated with MoSe2 nanoflowers. Sens. Actuators B Chem. 2021, 344, 130221. [Google Scholar] [CrossRef]
  25. Yang, Z.; Zhang, D.; Chen, H. MOF-derived indium oxide hollow microtubes/MoS2 nanoparticles for NO2 gas sensing. Sens. Actuators B Chem. 2019, 300, 127037. [Google Scholar] [CrossRef]
  26. Liu, C.; Huang, Z.; Wang, D.; Wang, X.; Miao, L.; Wang, X.; Wu, S.; Toyama, N.; Asaka, T.; Chen, J.; et al. Dynamic Ag+-intercalation with AgSnSe2 nanoprecipitates in Cl-doped polycrystalline SnSe2 toward ultra-high thermoelectric performance. J. Mater. Chem. A 2019, 7, 9761–9772. [Google Scholar] [CrossRef]
  27. Hang, N.; Zhang, S.; Yang, W. Efficient exfoliation of g-C3N4 and NO2 sensing behavior of graphene/g-C3N4 nanocomposite. Sens. Actuators B Chem. 2017, 248, 940–948. [Google Scholar] [CrossRef]
  28. Zhang, D.; Yang, Z.; Li, P.; Pang, M.; Xue, Q. Flexible self-powered high-performance ammonia sensor based on Au-decorated MoSe2 nanoflowers driven by single layer MoS2-flake piezoelectric nanogenerator. Nano Energy 2019, 65, 103974. [Google Scholar] [CrossRef]
  29. Yan, J.; Song, Z.; Xu, H.; Lee, L. Carbon-mediated electron transfer channel between SnO2 QDs and g-C3N4 for enhanced photocatalytic H2 production. Chem. Eng. J. 2021, 425, 131512. [Google Scholar] [CrossRef]
  30. Yan, J.; Wang, T.; Qiu, S.; Song, Z.; Zhu, W.; Liu, X.; Liang, J.; Sun, C.; Li, H. Insights into the efficient charge separation over Nb2O5/2D-C3N4 heterostructure for exceptional visible-light driven H2 evolution. J. Energy Chem. 2022, 65, 548–555. [Google Scholar] [CrossRef]
  31. Xiong, Y.; Lu, W.; Ding, D.; Zhu, L.; Li, X.; Ling, C.; Xue, Q. Enhanced room temperature oxygen sensing properties of LaOCl-SnO2 hollow spheres by UV light illumination. ACS Sens. 2017, 2, 679–686. [Google Scholar] [CrossRef] [PubMed]
  32. Zeng, B.; Zhang, L.; Wan, X. Fabrication of α-Fe2O3/g-C3N4 composites for cataluminescence sensing of H2S. Sens. Actuators B Chem. 2015, 211, 370–376. [Google Scholar] [CrossRef]
  33. Guo, J.; Zhang, J.; Zhu, M.; Ju, D.; Xu, H.; Cao, B. High-performance gas sensor based on ZnO nanowires functionalized by Au nanoparticles. Sens. Actuator B 2014, 199, 339–345. [Google Scholar] [CrossRef]
  34. Zhou, X.; Feng, W.; Wang, C.; Hu, X.; Li, X.; Sun, P.; Shimanoe, K.; Yamazoe, N.; Lu, G. Porous ZnO/ZnCo2O4 hollow spheres: Synthesis, characterization, and applications in gas sensing. J. Mater. Chem. A 2014, 2, 17683–17690. [Google Scholar] [CrossRef]
  35. Bai, S.; Chen, C.; Luo, R.; Chen, A.; Li, D. Synthesis of MoO3/reduced graphene oxide hybrids and mechanism of enhancing H2S sensing performances. Sens. Actuator B 2015, 216, 113–120. [Google Scholar] [CrossRef]
  36. Zhang, D.; Cao, Y.; Wu, J.; Zhang, X. Tungsten trioxide nanoparticles decorated tungsten disulfide nanoheterojunction for highly sensitive ethanol gas sensing application. Appl. Surf. Sci. 2020, 503, 144063. [Google Scholar] [CrossRef]
  37. Zhang, D.; Yang, Z.; Yu, S.; Mi, Q.; Pan, Q. Diversiform metal oxide-based hybrid nanostructure for gas sensing with versatile prospects. Coord. Chem. Rev. 2020, 413, 213272. [Google Scholar] [CrossRef]
  38. Triet, N.M.; Duy, L.T.; Hwang, B.U.; Hanif, A.; Siddiqui, S.; Park, K.H.; Cho, C.Y.; Lee, N.E. High-performance Schottky diode gas sensor based on the heterojunction of three-dimensional nanohybrids of reduced graphene oxide-vertical ZnO vanorods on an AIGaN/GaN layer. ACS Appl. Mater. Interfaces 2017, 9, 30722–30732. [Google Scholar] [CrossRef]
  39. Li, W.; Guo, J.; Cai, L.; Qi, W.; Sun, Y.; Xu, J.; Sun, M.; Zhu, H.; Xiang, L.; Xie, D.; et al. UV light irradiation enhanced gas sensor selectivity of NO2 and SO2 using rGO functionalized with hollow SnO2 nanofibers. Sens. Actuators B Chem. 2019, 290, 443–452. [Google Scholar] [CrossRef]
  40. Chen, D.; Tang, J.; Zhang, X.; Fang, J.; Li, Y.; Zhuo, R. Detecting decompositions of sulfur hexafluoride using reduced graphene oxide decorated with Pt nanoparticles. J. Phys. D Appl. Phys. 2018, 51, 185304. [Google Scholar] [CrossRef]
  41. Chen, A.; Liu, R.; Peng, X.; Chen, Q.; Wu, J. 2D hybrid nanomaterials for selective detection of NO2 and SO2 using “light on and off” strategy. ACS Appl. Mater. Interfaces 2017, 9, 37191–37200. [Google Scholar] [CrossRef] [PubMed]
  42. Tyagi, P.; Sharma, A.; Tomar, M.; Gupta, V. A comparative study of RGO-SnO2 and MWCNT-SnO2 nanocomposites based SO2 gas sensors. Sens. Actuators B Chem. 2017, 248, 980–986. [Google Scholar] [CrossRef]
  43. Chaudhary, V.; Kaur, A. Enhanced room temperature sulfur dioxide sensing behaviour of in situ polymerized polyaniline–tungsten oxide nanocomposite possessing honeycomb morphology. RSC Adv. 2015, 5, 73535–73544. [Google Scholar] [CrossRef]
  44. Stankova, M.; Vilanova, X.; Calderer, J. Detection of SO2 and H2S in CO2 stream by means of WO3-based micro-hotplate sensors. Sens. Actuators B Chem. 2004, 2, 219–225. [Google Scholar] [CrossRef]
  45. Ghimbeu, M.; Lumbreras, M.; Schoonman, J. Electrosprayed metal oxide semiconductor films for sensitive and selective detection of hydrogen sulfide. Sensors 2009, 9, 9122–9132. [Google Scholar] [CrossRef] [PubMed]
  46. Shen, F.; Wang, D.; Liu, R. Edge-tailored graphene oxide nanosheet-based field effect transistors for fast and reversible electronic detection of sulfur dioxide. Nanoscale 2012, 5, 537–540. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, D.; Liu, J.; Jiang, C.; Li, P.; Sun, Y. High-performance sulfur dioxide sensing properties of layer-by-layer self-assembled titania-modified graphene hybrid nanocomposite. Sens. Actuators B Chem. 2017, 245, 560–567. [Google Scholar] [CrossRef]
  48. Tyagi, P.; Sharma, A.; Tomar, M.; Gupta, V. Metal oxide catalyst assisted SnO2 thin film based SO2 gas sensor. Sens. Actuators B Chem. 2016, 224, 282–289. [Google Scholar] [CrossRef]
  49. Zhang, X.; Tie, J.; Zhang, J. A Pt-doped TiO2 nanotube arrays sensor for detecting SF6 decomposition products. Sensors 2013, 13, 14764–14776. [Google Scholar] [CrossRef] [Green Version]
  50. Chaudhary, V.; Kaur, A. Solitary surfactant assisted morphology dependent chemiresistive polyanilne sensors for room temperature monitoring of low ppm sulphur dioxide. Polym. Int. 2015, 64, 1475–1481. [Google Scholar] [CrossRef]
  51. Lupan, O.; Chow, L.; Chai, G. A single ZnO tetrapod-based sensor. Sens. Actuators B Chem. 2009, 141, 511–517. [Google Scholar] [CrossRef]
  52. Zhang, D.; Wu, J.; Peng, L. Room-temperature SO2 gas-sensing properties based on a metal-doped MoS2 nanoflower: An experimental and density functional theory investigation. J. Mater. Chem. A 2017, 5, 20666–20677. [Google Scholar] [CrossRef]
  53. Das, S.; Chakraborty, S.; Parkash, O. Vanadium doped tin dioxide as a novel sulfur dioxide sensor. Talanta 2008, 75, 385–389. [Google Scholar] [CrossRef] [PubMed]
  54. Morris, D.; Egdell, R. Application of V-doped TiO2 as a sensor for detection of SO2. J. Mater. Chem. 2001, 11, 3207–3210. [Google Scholar] [CrossRef]
  55. Kumar, R.; Avasthi, D.; Kaur, A. Fabrication of chemiresistive gas sensors based on multistep reduced graphene oxide for low parts per million monitoring of sulfur dioxide at room temperature. Sens. Actuators B Chem. 2017, 242, 461–468. [Google Scholar] [CrossRef]
  56. Nisar, J.; Topalian, Z.; de Sarkar, A.; Osterlund, L.; Ahuja, R. TiO2 based gas sensor: A possible application to SO2. ACS Appl. Mater. Interfaces 2013, 5, 8516–8522. [Google Scholar] [CrossRef]
  57. Zhang, S.; Song, P.; Liu, L.; Yang, Z.; Wang, Q. In2O3 nanosheets array directly grown on Al2O3 ceramic tube and their gas sensing performance. Ceram. Int. 2017, 43, 7942–7947. [Google Scholar] [CrossRef]
  58. Si, W.; Du, W.; Wang, F.; Wu, L.; Liu, J.; Liu, W.; Cui, P.; Zhang, X. One-pot hydrothermal synthesis of nano-sheet assembled NiO/ZnO microspheres for efficient sulfur dioxide detection. Ceram. Int. 2020, 46, 7279–7287. [Google Scholar] [CrossRef]
  59. Zhuo, Q.; Zeng, W.; Chen, W.; Xu, L.; Kumar, R.; Umar, A. High sensitive and low-concentration sulfur dioxide (SO2) gas sensor application of heterostructure NiO-ZnO nanodisks. Sens. Actuators B Chem. 2019, 298, 126870. [Google Scholar] [CrossRef]
  60. Ye, H.; Liu, L.; Xu, Y.; Wang, L.; Chen, X.; Zhang, K.; Liu, Y.; Koh, S.; Zhang, G. SnSe monolayer: A promising candidate of SO2 sensor with high adsorption quantity. Appl. Surf. Sci. 2019, 484, 33–38. [Google Scholar] [CrossRef]
  61. Seif, A.; nikfarjam, A.; Ghassem, H. UV enhanced ammonia gas sensing properties of PANI/TiO2 core-shell nanofibers. Sens. Actuators B Chem. 2019, 298, 126906. [Google Scholar]
  62. Bai, J.; Zhao, C.; Gong, H.; Wang, Q.; Huang, B.; Sun, G.; Wang, Y.; Zhou, J.; Xie, E.; Wang, F. Debye-length controlled gas sensing performances in NiO@ZnO p-n junctional core-shell nanotubes. J. Phys. D Appl. Phys. 2019, 52, 285103. [Google Scholar] [CrossRef]
  63. Su, P.; Zheng, Y. Room-temperature ppb-level SO2 gas sensors based on RGO/WO3 and MWCNTs/WO3 nanocomposites. Anal. Methods 2021, 13, 782–788. [Google Scholar] [CrossRef] [PubMed]
  64. Veeralingam, S.; Sahatiya, P.; Badhulika, S. Low cost, flexible and disposable SnSe2 based photoresponsive ammonia sensor for detection of ammonia in urine samples. Sens. Actuators B Chem. 2019, 297, 126725. [Google Scholar] [CrossRef]
  65. Yang, Z.; Zhang, D.; Wang, D. Carbon monoxide gas sensing properties of metal-organic frameworks-derived tin dioxide nanoparticles/molybdenum diselenide nanoflowers. Sens. Actuators B Chem. 2020, 304, 127369. [Google Scholar] [CrossRef]
  66. Guo, J.; Li, Y.; Jiang, B.; Gao, H.; Wang, T.; Sun, P.; Liu, F.; Yan, X.; Liang, X.; Gao, Y.; et al. Xylene gas sensing properties of hydrothermal synthesized SnO2-Co3O4 microstructure. Sens. Actuators B Chem. 2020, 310, 127780. [Google Scholar] [CrossRef]
  67. Beniwai, A.; Sunny. Electrospun SnO2/PPy nanocomposite for ultra-low ammonia concentration detection at room temperature. Sens. Actuators B Chem. 2019, 296, 126660. [Google Scholar] [CrossRef]
  68. Xue, Q.; Chen, H.; Li, Q.; Yan, K.; Besenbacher, F.; Dong, M. Room-temperature high-sensitivity detection of ammonia gas using the capacitance of carbon/silicon heterojunctions. Energy Environ. Sci. 2010, 3, 288–291. [Google Scholar] [CrossRef]
Figure 1. Synthesis process of g-C3N4/SnSe2 nanomaterials.
Figure 1. Synthesis process of g-C3N4/SnSe2 nanomaterials.
Chemosensors 10 00401 g001
Figure 2. Structure illustration of the ceramic-tube-based SO2 gas sensor.
Figure 2. Structure illustration of the ceramic-tube-based SO2 gas sensor.
Chemosensors 10 00401 g002
Figure 3. Schematic illustration of sensor performance test platform.
Figure 3. Schematic illustration of sensor performance test platform.
Chemosensors 10 00401 g003
Figure 4. (a) XRD patterns of g-C3N4, SnSe2 and g-C3N4/SnSe2. XPS spectra of g-C3N4/SnSe2 product: (b) full spectrum, (c) Se 3d, (d) Sn 3d, (e) N 1s and (f) C 1s.
Figure 4. (a) XRD patterns of g-C3N4, SnSe2 and g-C3N4/SnSe2. XPS spectra of g-C3N4/SnSe2 product: (b) full spectrum, (c) Se 3d, (d) Sn 3d, (e) N 1s and (f) C 1s.
Chemosensors 10 00401 g004
Figure 5. SEM images of (a) SnSe2, (b) g-C3N4, and (c,d) g-C3N4/SnSe2.
Figure 5. SEM images of (a) SnSe2, (b) g-C3N4, and (c,d) g-C3N4/SnSe2.
Chemosensors 10 00401 g005
Figure 6. (a) Responses of g-C3N4/SnSe2 composites with different quality ratios to 20 ppm SO2 gas at various temperatures. (b) Sensing properties of pristine g-C3N4, SnSe2 and g-C3N4/SnSe2 composite sensors at different concentrations of SO2 gas at 200 °C. (c) Response fitting curves of three sensors versus SO2 concentration at 200 °C. (d) Typical response–recovery curves for various concentrations of SO2 gas at 200 °C.
Figure 6. (a) Responses of g-C3N4/SnSe2 composites with different quality ratios to 20 ppm SO2 gas at various temperatures. (b) Sensing properties of pristine g-C3N4, SnSe2 and g-C3N4/SnSe2 composite sensors at different concentrations of SO2 gas at 200 °C. (c) Response fitting curves of three sensors versus SO2 concentration at 200 °C. (d) Typical response–recovery curves for various concentrations of SO2 gas at 200 °C.
Chemosensors 10 00401 g006
Figure 7. At the optimal temperature of 200 °C (a) repeatability of g-C3N4/SnSe2 composite sensor. (b) Response/recovery curves of g-C3N4, SnSe2 and g-C3N4/SnSe2 composite sensors exposed to 50 ppm SO2 gas. (c) Selectivity and (d) long-term stability of g-C3N4/SnSe2 composite sensor.
Figure 7. At the optimal temperature of 200 °C (a) repeatability of g-C3N4/SnSe2 composite sensor. (b) Response/recovery curves of g-C3N4, SnSe2 and g-C3N4/SnSe2 composite sensors exposed to 50 ppm SO2 gas. (c) Selectivity and (d) long-term stability of g-C3N4/SnSe2 composite sensor.
Chemosensors 10 00401 g007
Figure 8. Schematic diagram of microscopic sensing mechanism of g-C3N4/SnSe2 composite sensor.
Figure 8. Schematic diagram of microscopic sensing mechanism of g-C3N4/SnSe2 composite sensor.
Chemosensors 10 00401 g008
Table 1. Comparison of the SO2-sensing performance between the present work and previous reported studies.
Table 1. Comparison of the SO2-sensing performance between the present work and previous reported studies.
Sensing MaterialSensing EnvironmentResponse ConcentrationRef.
AlGaN/ZnO/rGORT2.5%120 ppb[38]
SnO2/rGO RT/UV1.7%5 ppm[39]
Pt/rGO120 °C5%100 ppm[40]
g-C3N4/rGORT3.2%100 ppm[41]
SnO2/MWCNT60 °C6500 ppm[42]
V2O5/WO3/TiO2400 °C5%20 ppm[43]
WO3350 °C5%1 ppm[44]
Cu–SnO2400 °C1.1%20 ppm[45]
SnO2–PANIRT3.1%4 ppm[46]
TiO2/rGORT11.14%5 ppm[47]
NiO–SnO2180 °C8.3%50 ppm[48]
TiO2200 °C11%10 ppm[49]
ZnORT0.2%100 ppm[50]
WO3–PANIRT4.3%5 ppm[51]
Ni–MoS2RT7.4%5 ppm[52]
V2O5/SnO2350 °C45%5 ppm[53]
V-doped TiO2400 °C10%10 ppm[54]
GORT6%5 ppm[55]
PANIRT4.2%10 ppm[56]
g-C3N4/SnSe2200 °C28.9%20 ppmThis work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, H.; Pan, Q.; Zhang, Y.; Zhang, Y.; Zhang, D. High-Performance Sulfur Dioxide Gas Sensor Based on Graphite-Phase Carbon-Nitride-Functionalized Tin Diselenide Nanorods Composite. Chemosensors 2022, 10, 401. https://doi.org/10.3390/chemosensors10100401

AMA Style

Zhang H, Pan Q, Zhang Y, Zhang Y, Zhang D. High-Performance Sulfur Dioxide Gas Sensor Based on Graphite-Phase Carbon-Nitride-Functionalized Tin Diselenide Nanorods Composite. Chemosensors. 2022; 10(10):401. https://doi.org/10.3390/chemosensors10100401

Chicago/Turabian Style

Zhang, Hao, Qiannan Pan, Yating Zhang, Yanting Zhang, and Dongzhi Zhang. 2022. "High-Performance Sulfur Dioxide Gas Sensor Based on Graphite-Phase Carbon-Nitride-Functionalized Tin Diselenide Nanorods Composite" Chemosensors 10, no. 10: 401. https://doi.org/10.3390/chemosensors10100401

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