Next Article in Journal
A Scaled Proteomic Discovery Study for Prostate Cancer Diagnostic Markers Using ProteographTM and Trapped Ion Mobility Mass Spectrometry
Previous Article in Journal
Optimization of an Ischemic Retinopathy Mouse Model and the Consequences of Hypoxia in a Time-Dependent Manner
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gas-Sensing Performance of Metal Oxide Heterojunction Materials for SF6 Decomposition Gases: A DFT Study

by
Tingting Zeng
1,
Donglin Ma
1 and
Yingang Gui
2,*
1
College of Physics and Engineering, Chengdu Normal University, Chengdu 611130, China
2
College of Engineering and Technology, Southwest University, Chongqing 400715, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(15), 8009; https://doi.org/10.3390/ijms25158009
Submission received: 28 May 2024 / Revised: 18 July 2024 / Accepted: 19 July 2024 / Published: 23 July 2024

Abstract

:
The online monitoring of GIS equipment can be realized through detecting SF6 decomposition gasses. Metal oxide heterojunctions are widely used as gas-sensing materials. In this study, the structural and electrical properties of In2O3-ZnO and TiO2-ZnO heterojunctions were analyzed based on density functional theory calculations. After heterojunction structural optimization, the electrical conductivity of these two heterojunctions was enhanced compared to each intrinsic model, and the electrical conductivity is ranked as follows: In2O3-ZnO heterojunction > TiO2-ZnO heterojunction. The gas-sensing response of these two heterojunctions to four SF6 decomposition gasses, H2S, SO2, SOF2, and SO2F2, was investigated. For gas adsorption systems, the adsorption energy, charge transfer, density of states, charge difference density, and frontier molecular orbitals were calculated to analyze the adsorption and gas-sensing performance. For gas adsorption on the In2O3-ZnO heterojunction surface, the induced conductivity changes are in the following order: H2S > SO2F2 > SOF2 > SO2. For gas adsorption on the TiO2-ZnO heterojunction surface, H2S and SOF2 increase conductivity, and SO2 and SO2F2 decrease conductivity.

1. Introduction

In modern power systems, the gas insulated switchgear (GIS) is an essential critical piece of equipment to ensure the safe and stable operation of the power grid [1]. Sulfur hexafluoride gas (SF6) is widely used in GIS equipment due to its excellent insulation and arc extinguishing performance [2]. However, during the long-term operation of GIS equipment, external environmental factors (such as overvoltage and high temperature) and inherent insulation defects of the equipment itself may lead to internal discharge and overheating faults [3]; furthermore, this can lead to SF6 decomposition, which has a serious impact on GIS reliability. Therefore, in order to ensure the stable operation of GIS equipment, effective online monitoring measures must be taken to quickly identify and eliminate the potential insulation faults [4]. Research has confirmed that SF6 decomposition gasses mainly include H2S, SO2, SOF2, and SO2F2, and the operating state of the GIS device can be reflected by detecting the presence and concentration of these gasses [5,6]. Therefore, it is necessary to explore gas-sensitive materials with high sensitivity and selectivity for SF6 decomposition gasses detection.
In recent years, the adsorption research on SF6 decomposition gas mainly focuses on graphene, two-dimensional metal sulfur compounds (TMDCs), metal oxides, and surface-modified metal oxides [7,8,9]. Among them, metal oxides are widely used materials with the characteristics of low cost, low power consumption, and easy synthesis for gas detection [10], including ZnO, TiO2, In2O3, SnO2, and WO3. ZnO is an n-type semiconductor, which is widely used for detecting toxic and hazardous gasses due to its wide bandgap and high electron mobility [11]. At the same time, researchers have also proposed various forms of nanostructures, such as nanosheets [12], nanorings [13], nanowires [14], and nanorods [15], to improve the sensitivity, selectivity, and stability of the ZnO-based gas sensor, and to reduce the working temperature and reaction recovery time. Razavi et al. prepared ZnO@TiO2 nanocomposites and verified their potential in the determination of hydrazine (N2H4) [16]. The research showed that TiO2 exhibits high selectivity and sensitivity towards small gas molecules [17,18,19]. Shooshtariet al. prepared an electronic nose based on carbon nanotube-titanium dioxide hybrid nanostructures for the detection and discrimination of volatile organic compounds (VOCs) [20]. In addition, In2O3 is also widely used in gas detection and monitoring due to its suitable bandgap and good chemical stability, and excellent gas-sensing performance [21,22]. The gas-sensing property of In2O3 can also be improved through surface metal modification or the formation of heterojunctions [23]. However, intrinsic metal oxides have limited gas sensitivity to specific gasses, and metal oxide heterojunctions are often used to improve the gas sensitivity of metal oxides-based gas sensors due to Fermi level effects and synergistic effects between different metal oxides [24,25]. Cheng et al. reported an ammonia sensor based on the ZnO/CuO heterojunction using a hydrothermal method, which effectively improved the gas-sensing property of the ZnO sensor for NH3 at room temperature [26].
In order to improve the sensitivity and selectivity of gas-sensitive materials towards the SF6 decomposition gasses of H2S, SO2, SOF2, and SO2F2, this study adopts In2O3-ZnO and TiO2-ZnO heterojunctions as gas-sensitive materials, and studies their gas-sensing properties to SF6 decomposition gasses based on density functional theory (DFT). Firstly, the structural models of In2O3-ZnO and TiO2-ZnO heterojunctions were constructed and optimized, and then the structural and energy band properties were analyzed. Secondly, the adsorption structures of different SF6 decomposition gasses on the heterojunction surface were constructed and optimized. Finally, by analyzing the density of states (DOS) and charge difference density (CDD) of gas adsorption systems, the adsorption performance of the metal oxide heterojunction materials for SF6 decomposition gasses was analyzed. The research results play an important role in promoting the application of gas-sensing technology in the online monitoring of power insulation equipment.

2. Results and Discussion

2.1. Adsorption Properties of In2O3-ZnO Heterojunction on SF6 Decomposition Gases

2.1.1. Construction and Energy Band Analysis of In2O3-ZnO Heterojunction Model

In the section, the most stable structure of the In2O3-ZnO heterojunction model was constructed and optimized, as shown in Figure 1a,b. The In2O3-ZnO heterojunction contains a total of 120 atoms, including 40 atoms in the In2O3 layer and 80 atoms in the ZnO heterojunction layer. The ZnO heterojunction crystal plane is placed in the lower layer, and the In2O3 crystal plane is placed in the upper layer. The lattice parameters of the ZnO heterojunction are as follows: a = 11.2558 Å, b = 16.2464 Å, and γ = 90°. The lattice parameters of In2O3 are as follows: a = 11.0324 Å, b = 16.0283 Å, and γ = 90°. The lattice parameters of the constructed In2O3-ZnO heterojunction are as follows: a = 11.1441 Å, b = 16.1373 Å, and γ = 90°. In heterostructures, there is a formation of chemical bonds between the In2O3 layer and the ZnO layer, which implies a strong interaction force between the two metal oxides. In addition, its enormous formation energy (−18.017 eV) and significant geometric deformation also prove that the formation between the two materials is highly stable. The energy band structure of the In2O3-ZnO heterojunction shown in Figure 1c intuitively displays the electron property of the heterojunction. According to the calculation results, the energy required for electrons to transition from the valence band to the conduction band is 0.652 eV. Compared to the intrinsic In2O3 (1.538 eV) and intrinsic ZnO heterojunction (1.769 eV), the significant decrease in energy indicates that the conductivity of the heterojunction is superior.

2.1.2. Adsorption Properties of In2O3-ZnO Heterojunction

  • Gas adsorption structure analysis
Figure 2 shows the most stable adsorption structures for H2S, SO2, SOF2, and SO2F2 on the In2O3-ZnO heterojunction surface, and Table 1 lists the corresponding adsorption parameters. For the H2S adsorption system shown in Figure 2(a1,a2), the strong binding force between the O and H atoms leads to the breaking of the H-S bond, resulting in the formation of a new H-O bond (1.008 Å); here, a new chemical bond is formed between the S atom and the nearest In atom with a length of 2.486 Å. As listed in Table 1, the adsorption energy of the H2S adsorption system reaches −1.924 eV, indicating chemical adsorption. H2S molecules act as electron contributors and transfer the 0.255 |e| charge to the In2O3-ZnO heterojunction, which is consistent with the reducibility properties of H2S. As shown in Figure 2(b1,b2), the O atoms of the SO2 molecule form chemical bonds with In atoms, while the S atom of the SO2 molecule forms chemical bonds with O atoms, resulting in varying degrees of elongation of the S-O bond in SO2. Meanwhile, the system has an adsorption energy of −1.992 eV and a charge transfer capacity of −0.361 |e|, indicating a strong binding force between the SO2 molecules and the In2O3-ZnO heterojunction. For the adsorption systems of SOF2 and SO2F2, shown in Figure 2(c1,c2,d1,d2), there is a significant distance between the gas molecules and heterojunctions; gas molecules can easily desorb from the surface of the In2O3-ZnO heterojunction, indicating that the adsorption process is physical adsorption. In addition, the small adsorption energy (SOF2: Eads = −0.465 eV, SO2F2: Eads = −0.501 eV) and the small amount of charge transfer (SOF2: Qt = 0.016 |e|, SO2F2: Qt = 0.011 |e|) also confirm this physical adsorption effect. In summary, the adsorption process of H2S and SO2 is chemical adsorption, and the molecular structure of the H2S gas molecules is destroyed, making it difficult to carry out the desorption process. The adsorption process of SOF2 and SO2F2 is physical adsorption and the adsorption energy is moderate, which together can have a certain adsorption capacity for gas molecules and make desorption possible.
  • DOS, CDD, and molecular orbital analysis
Figure 3 shows the density of states distribution of H2S, SO2, SOF2, and SO2F2 adsorption on the In2O3-ZnO heterojunction surface. In the H2S adsorption system shown in Figure 3(a1,a2), there is no significant change in TDOS. It can also be observed from PDOS that the hybridization phenomenon between atomic orbitals is very weak. This phenomenon indicates that the conductivity of the system remains almost unchanged before and after H2S adsorption. Figure 3(b1,b2) shows the TDOS and PDOS of the SO2 gas adsorption system. It can be observed that TDOS almost completely overlaps above the Fermi level. However, the overall density of states shows an increasing trend below the Fermi level, with the peak value increasing most significantly around −5 eV. Due to the increase in electron filling probability at this location, the conductivity of the system increases. In the PDOS distribution, there is a very obvious hybridization phenomenon between the 5p orbitals of the In atoms and the 2p orbitals of the S atoms between −10 eV and 5 eV. This phenomenon confirms the strong chemical interaction between the SO2 and In2O3-ZnO heterojunction, as well as the increase in TDOS near −5 eV. During the adsorption process of SOF2 shown in Figure 3(c1,c2), TDOS remains almost unchanged, which is consistent with its small adsorption energy and small charge transfer performance. In Figure 3(d1,d2), TDOS also keeps overlapping before and after the adsorption of SO2F2 gas, except in two new small peaks, which appear near −10 eV. In the PDOS analysis of the SO2F2 adsorption process, only In-5s orbitals and F-2p orbitals have a certain degree of hybridization, but the hybridization is not significant, which has a certain impact on improving gas adsorption performance. This result explains that compared to SOF2 gas molecules, the SO2F2 gas adsorption system has a greater adsorption energy.
To further understand the adsorption mechanism of the In2O3-ZnO heterojunction on the four SF6 decomposition gasses, the charge difference density before and after adsorption were studied as shown in Figure 4; the increase in charge density is shown in red, while the decrease is shown in blue. For the SOF2 and SO2F2 gas molecules adsorption, there is almost no charge transfer between the In2O3-ZnO heterojunction and the gas molecules. When adsorbing H2S gas molecules, the S atom acts as an electron acceptor, indicating a decrease in the number of charges around the S atom. However, the red color around the S atom is very weak, and the entire H2S gas molecule transfers 0.255 |e| to the In2O3-ZnO heterojunction substrate during the adsorption process. When adsorbing SO2 gas molecules, the charge density around the O atom of SO2 molecules significantly increases, indicating that the charge transfers from the In2O3-ZnO heterojunction to SO2. This is attributed to the formation of chemical bonds between gas molecules and the substrate, as well as the large adsorption energy. The In2O3-ZnO heterojunction has a good adsorption effect on the SO2 gas molecule, but this adsorption force is not sufficient to break the original chemical bonds of the SO2 gas molecules. Therefore, the SO2 gas molecules have the potential for desorption.
Figure 5 shows the changes in the frontier molecular orbitals of theIn2O3-ZnO heterojunction before and after H2S, SO2, SOF2, and SO2F2 molecules adsorption, where LUMO and HOMO represent the lowest unoccupied molecular orbitals and the highest occupied molecular orbitals, respectively. It can be observed that after gas molecule adsorption, the molecular orbitals and corresponding energies change. Usually, changes in the energy gap can directly affect conductivity. During the adsorption process of H2S and SO2, there is a significant change in the energy gap value, which is related to chemical adsorption. Specifically, the conductivity of the H2S adsorption system increases, while the conductivity of the SO2 adsorption system decreases. After the adsorption of SOF2 and SO2F2 gas molecules, the energy gap increases, indicating a decrease in conductivity. Although this change is not very obvious, it is consistent with the previous analysis results. According to the extent of the energy gap change, the conductivity changes caused by gas adsorption are ranked in descending order: SO2 > H2S > SO2F2 > SOF2.
This section calculated and analyzed the adsorption performance of the In2O3-ZnO heterojunction on four gas molecules: H2S, SO2, SOF2, and SO2F2. Firstly, the most stable heterostructure and gas adsorption model were structured; each structure has a negative binding or adsorption energy, indicating a spontaneously performed reaction. There is a formation of chemical bonds in the adsorption process of H2S and SO2, indicating a chemical adsorption, while the adsorption of SOF2 and SO2F2 belongs to physical adsorption with small adsorption energy and longer adsorption distance. According to the density of states and molecular orbital theory analysis, the conductivity of the H2S adsorption system increases, while the conductivity of the SO2 adsorption system decreases. However, in the H2S adsorption system, the structure of gas molecules is disrupted, and cannot be desorbed from the In2O3-ZnO heterojunction surface. In the SO2 adsorption system, the original chemical bonds of SO2 gas molecules keep intact, showing a certain desorption potential. Therefore, the In2O3-ZnO heterojunction can serve as an adsorbent material for H2S gas molecules, while it can also be used as a gas-sensitive material for SO2 gas molecules.

2.2. Adsorption Properties of TiO2-ZnO Heterojunction on SF6 Decomposition Gases

2.2.1. Construction and Energy Band Analysis of TiO2-ZnO Heterojunction Model

This section constructs and optimizes the most stable model by combining the TiO2 and ZnO heterojunction crystal planes. Firstly, the TiO2 crystal plane and ZnO heterojunction crystal plane were constructed and optimized separately. The lattice parameters of the TiO2 crystal were as follows: a = b = 3.776 Å, c = 9.486 Å, γ = 90°. The ZnO heterojunction crystal’s lattice parameters were as follows: a = b = 3.249 Å, c = 5.205 Å, γ = 90°. Then, the TiO2 and ZnO heterojunction crystal planes were cut and expanded, and the obtained parameters were as follows: a = 18.880 Å, b = 10.210 Å, γ = 90° (TiO2 heterojunction); a = 19.496 Å, b = 10.411 Å, γ = 90° (ZnO heterojunction). In this model, the lower layer is the TiO2 layer, and the upper layer is the ZnO heterojunction layer. There are a total of 156 atoms, including 96 atoms in the ZnO layer and 60 atoms in the TiO2 layer. Figure 6a,b shows the structure of TiO2-ZnO heterojunction after structural optimization, with a binding energy of 12.819 eV. In the model, the TiO2 layer and the ZnO layer are connected by chemical bonds. In addition, the formation of chemical bonds makes it easier for electrons to migrate from the ZnO heterojunction layer to the TiO2 layer, as evidenced by the small energy band (1.292 eV), shown in Figure 6c. Therefore, the conductivity of the TiO2-ZnO heterojunction is enhanced.

2.2.2. Adsorption Properties of TiO2-ZnO Heterojunction

  • Gas adsorption structure analysis
As shown in Figure 7, the most stable adsorption models are calculated by adjusting the various adsorption sites of gas molecules. As listed in Table 2, the H2S gas adsorption system has the maximum adsorption (−0.788 eV) and the maximum charge transfer amount (0.249 |e|). From Figure 7(a1,a2), it can be observed that the shortest distance between the S atoms and Ti atoms is 2.780 Å, and the long adsorption distance makes it impossible to form a chemical bond between the two. In the adsorption system of the TiO2-ZnO heterojunction for the four gasses, only chemical bonds form between the SO2 gas molecule and the substrate material, while the other three gas molecules still maintained a relatively long adsorption distance from the TiO2-ZnO heterojunction substrate. In the adsorption system of SOF2 and SO2F2, they have a long adsorption distance, small adsorption energy, and a small amount of charge transfer. Therefore, the adsorption of SOF2 and SO2F2 gas molecules on the TiO2-ZnO heterojunction is physical adsorption. In addition, only the charge transfer amount of the SO2 adsorption system is negative, signifying that gas molecules act as electron acceptors to obtain electrons from the substrate. It is the formation of chemical bonds (S-O: 1.766 Å; Ti-O: 1.958 Å) that facilitate electron migration, allowing SO2 gas molecules to form a whole with the TiO2 layer, and obtain more electrons from the ZnO heterojunction layer. In the other three adsorption systems, the TiO2 layer acts as a stronger electron acceptor, causing gas molecules to lose electrons. In the above gas adsorption, the molecular structure of the gas molecules is not disrupted. Therefore, under certain conditions, the four gas molecules can desorb from the surface of the TiO2-ZnO heterojunction.
  • DOS, CDD, and molecular orbital analysis
To investigate the gas-sensing performance of the TiO2-ZnO heterojunction on H2S, SO2, SOF2, and SO2F2 gas molecules, the density of states distribution of the four gas adsorption systems was analyzed, as shown in Figure 8. There is no significant change in TDOS before and after the four gas molecules’ adsorption on the surface of the TiO2-ZnO heterojunction. From PDOS analysis, it can be seen that the degree of atomic orbital hybridization is very weak. Therefore, it is not possible to accurately determine the changes in conductivity from the density of states analysis. Figure 9 shows the distribution of charges upon gas adsorption. From Figure 9a, it can be observed that the electron density around H2S gas is relatively low, which is consistent with the conclusion that the H2S molecule acts as an electron donor. In Figure 9b, the blue color is around the S atom of the SO2 gas molecule and the red color is around the O atoms of the TiO2 layer, indicating that SO2 receives electrons from the TiO2 layer. This conclusion stays consistent with the charge transfer of SO2 of−0.142 |e| during the adsorption process. In the SOF2 and SO2F2 adsorption systems, there is no significant change in the distribution of electrons, indicating a small amount of charge transfer.
Figure 10 shows the frontier molecular orbitals’ distribution and the related energy of the gas molecules before and after adsorption on the surface of the TiO2-ZnO heterojunction. Before gas adsorption, LUMO is mainly distributed on the TiO2 layer, while HOMO is mainly distributed on the ZnO heterojunction layer, and both are evenly distributed. The energy corresponding to LUMO is −4.921 eV, and the energy corresponding to HOMO is −6.213 eV. In the three gas adsorption systems of H2S, SOF2, and SO2F2, the energy gap only undergoes slight changes, within a range of less than 0.03 eV. The conductivity of the H2S and SOF2 gas adsorption systems increases, while the conductivity of the SO2F2 gas adsorption systems decreases. In the SO2 gas adsorption system, the distribution of LUMO on the TiO2 layer becomes thinner, followed by a significant increase in the energy gap value, which is 0.086 eV. This means there is a significant increase in conductivity after gas adsorption, providing a theoretical basis for the development of resistive chemical sensors.
This section investigated the gas-sensing response of TiO2-ZnO heterojunctions to four gas molecules. Based on DFT calculations, the most stable heterostructure and gas adsorption model were calculated, and the gas-sensing performance of the TiO2-ZnO heterojunction heterostructure on four gas molecules was explored by analyzing density of states, CDD, and the HOMO-LUMO gap. The calculation results show that the TiO2 and ZnO crystal planes can form stable heterojunctions with high formation energy. The adsorption type of H2S, SOF2, and SO2F2 by the TiO2-ZnO heterojunction is physical adsorption, while the adsorption type of SO2 is chemical adsorption. The conductivity of the TiO2-ZnO heterojunction increases after adsorbing H2S and SOF2 gas molecules; after adsorbing SO2 and SO2F2, there is a decrease in conductivity. Specifically, the change in conductivity is most significant after the adsorption of SO2, indicating that the TiO2-ZnO heterojunction can effectively detect SO2 gas.

3. Methods

All calculations were performed based on DFT [27]. Specifically, the generalized gradient approximation method with a GGA-PBE function was used for geometric optimization and energy calculation [28], while the Tkatchenko–Scheffler method was used to correct the van der Waals force to obtain accurate results [29]. In addition, a DFT semi-core pseudopotential (DSSP) was adopted and its value was set to 6 [30]. The K-points in the Brillouin area were set to 5 × 5 × 1. The energy convergence accuracy, maximum stress, and displacement were set to 2.721 × 10−4 eV, 0.109 eV/Å, and 0.005 Å, respectively [31,32]. To eliminate the influence of interlayer interactions, the thickness of the vacuum layer was set to 20 Å.

4. Conclusions

In order to achieve the online monitoring of GIS faults, this study analyzed the adsorption and gas-sensing properties of In2O3-ZnO and TiO2-ZnO heterojunctions on four SF6 decomposition gasses (H2S, SO2, SOF2, and SO2F2). Compared to intrinsic In2O3 and ZnO, the band value of the In2O3-ZnO heterojunction significantly decreases. The energy band value of the TiO2-ZnO heterojunction is larger than that of the In2O3-ZnO heterojunction. The In2O3-ZnO heterojunction shows large adsorption energy for H2S and SO2 via chemical adsorption, and small adsorption energy for SOF2 and SO2F2 via physical adsorption. After gas adsorption, the conductivity of the H2S adsorption system increases, while the conductivity of the SO2 adsorption system decreases. The In2O3-ZnO heterojunction can serve as an adsorbent material for H2S gas molecules, while it can also be used as a gas-sensitive material for SO2 gas molecules. While TiO2-ZnO only shows chemical adsorption for SO2, there is a distinct change in the conductivity of the adsorption system. The research results play an important role in exploring novel gas-sensitive materials and preparing a specific gas sensor using the online monitoring of power insulation equipment.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ijms25158009/s1.

Author Contributions

Conceptualization, T.Z.; methodology, Y.G.; software, D.M.; validation, T.Z.; formal analysis, T.Z.; investigation, T.Z.; resources, D.M.; data curation, T.Z.; writing—original draft preparation, T.Z.; writing—review and editing, Y.G.; visualization, Y.G.; supervision, Y.G.; project administration, Y.G.; funding acquisition, T.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Talent Introduction Program of Chengdu Normal University (No. YJRC 2020-14), and Innovative Experimental Project of Chengdu Normal University (No. 2023JG09, No. CXSY2024009).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

SF6sulfur hexafluoride
GISgas insulated switchgear
TMDCstwo-dimensional metal sulfur compounds
VOCsvolatile organic compounds
DFTdensity functional theory
DOSdensity of states
TDOStotal density of states
PDOSpartial density of states
CDDcharge difference density
LUMOlowest unoccupied molecular orbital
HOMOhighest occupied molecular orbital
GGAgeneralized gradient approximation
PBEPerdew–Burke–Ernzerhof
DSSPDFT semi-core pseudopotential

References

  1. Wang, J.; Chen, W.; Wang, P.; Wan, F.; Zhang, Z.; Gao, S.; Wang, Y. Analysis of SF6 decomposed products by fibre-enhanced Raman spectroscopy for gas-insulated switchgear diagnosis. High Volt. 2024, 9, 206–216. [Google Scholar] [CrossRef]
  2. Khan, Q.; Refaat, S.S.; Abu-Rub, H.; Toliyat, H.A. Partial discharge detection and diagnosis in gas insulated switchgear: State of the art. IEEE Electr. Insul. Mag. 2019, 35, 16–33. [Google Scholar] [CrossRef]
  3. Zhuang, Y.; Hu, X.; Tang, B.; Wang, S.; Cui, A.; Hou, K.; He, Y.; Zhu, L.; Li, W.; Chu, J. Effects of SF6 decomposition components and concentrations on the discharge faults and insulation defects in GIS equipment. Sci. Rep. 2020, 10, 15039. [Google Scholar] [CrossRef] [PubMed]
  4. Faizol, Z.; Zubir, F.; Saman, N.M.; Ahmad, M.H.; Rahim, M.K.A.; Ayop, O.; Jusoh, M.; Majid, H.A.; Yusoff, Z. Detection Method of Partial Discharge on Transformer and Gas-Insulated Switchgear: A Review. Appl. Sci. 2023, 13, 9605. [Google Scholar] [CrossRef]
  5. Fan, X.; Li, L.; Zhou, Y.; Tang, N.; Zou, Z.; Li, X.; Huang, G.; Liu, M. Online detection technology for SF6 decomposition products in electrical equipment: A review. IET Sci. Meas. Technol. 2018, 12, 707–711. [Google Scholar] [CrossRef]
  6. Fang, H.; Chen, H.; Zhang, F.; Tan, R.; Han, C.; Ai, X.; Ma, X.; Cui, Y.; Wang, D. Trace detection of SF6 gas decomposition component H2S based on Pr6O11-In2O3. Sens. Actuators B Chem. 2024, 412, 135822. [Google Scholar] [CrossRef]
  7. Li, Z.; Jia, L.; Chen, J.; Cui, X.; Zeng, W.; Zhou, Q. Ag-modified hexagonal GaN monolayer as an innovative gas detector toward SF6 decomposed species: Insights from the first-principles computations. Appl. Surf. Sci. 2022, 589, 153000. [Google Scholar] [CrossRef]
  8. Wang, L.; Wang, X.; Yan, Y.; Qiu, H.; Guo, X.; Tang, J.; Zeng, F. Theoretical analysis of Ni atom-doped MXene for improving the catalytic degradation performance of SF6. Comput. Theor. Chem. 2023, 1229, 114325. [Google Scholar] [CrossRef]
  9. Meng, H.; Liu, Z.; Wang, X.; Feng, L. A fluorinated greenhouse gas sensor based on N-doped tin oxide materials. Environ. Sci.-Nano 2024, 11, 459–469. [Google Scholar] [CrossRef]
  10. Krishna, K.G.; Parne, S.; Pothukanuri, N.; Kathirvelu, V.; Gandi, S.; Joshi, D. Nanostructured metal oxide semiconductor-based gas sensors: A comprehensive review. Sens. Actuators A Phys. 2022, 341, 113578. [Google Scholar] [CrossRef]
  11. Franco, M.A.; Conti, P.P.; Andre, R.S.; Correa, D.S. A review on chemiresistive ZnO gas sensors. Sens. Actuators Rep. 2022, 4, 100100. [Google Scholar] [CrossRef]
  12. Yuan, H.; Aljneibi, S.A.A.A.; Yuan, J.; Wang, Y.; Liu, H.; Fang, J.; Tang, C.; Yan, X.; Cai, H.; Gu, Y.; et al. ZnO Nanosheets Abundant in Oxygen Vacancies Derived from Metal-Organic Frameworks for ppb-Level Gas Sensing. Adv. Mater. 2019, 31, 1807161. [Google Scholar] [CrossRef] [PubMed]
  13. Alenezi, M.R. Hierarchical zinc oxide nanorings with superior sensing properties. Mater. Sci. Eng. B 2018, 236–237, 132–138. [Google Scholar] [CrossRef]
  14. Zhao, S.; Shen, Y.; Li, A.; Chen, Y.; Gao, S.; Liu, W.; Wei, D. Effects of rare earth elements doping on gas sensing properties of ZnO nanowires. Ceram. Int. 2021, 47, 24218–24226. [Google Scholar] [CrossRef]
  15. Zhang, B.; Sun, J.-Y.; Gao, P.-X. Low-Concentration NOx Gas Analysis Using Single Bimodular ZnO Nanorod Sensor. ACS Sens. 2021, 6, 2979–2987. [Google Scholar] [CrossRef] [PubMed]
  16. Razavi, R.; Nejad, F.G.; Ahmadi, S.A.; Beitollahi, H. Synthesis of ZnO@TiO2 nanoparticles and its application to construct an electrochemical sensor for determination of hydrazine. Electrochem. Commun. 2024, 159, 107639. [Google Scholar] [CrossRef]
  17. Tian, X.; Cui, X.; Lai, T.; Ren, J.; Yang, Z.; Xiao, M.; Wang, B.; Xiao, X.; Wang, Y. Gas sensors based on TiO2 nanostructured materials for the detection of hazardous gases: A review. Nano Mater. Sci. 2021, 3, 390–403. [Google Scholar] [CrossRef]
  18. Haidry, A.A.; Xie, L.; Wang, Z.; Zavabeti, A.; Li, Z.; Plecenik, T.; Gregor, M.; Roch, T.; Plecenik, A. Remarkable Improvement in Hydrogen Sensing Characteristics with Pt/TiO2 Interface Control. ACS Sens. 2019, 4, 2997–3006. [Google Scholar] [CrossRef] [PubMed]
  19. Tshabalala, Z.P.; Swart, H.C.; Motaung, D.E. Fabrication of TiO2 nanofibers based sensors for enhanced CH4 performance induced by notable surface area and acid treatment. Vacuum 2021, 187, 110102. [Google Scholar] [CrossRef]
  20. Shooshtari, M.; Salehi, A. An electronic nose based on carbon nanotube-titanium dioxide hybrid nanostructures for detection and discrimination of volatile organic compounds. Sens. Actuators B-Chem. 2022, 357. [Google Scholar] [CrossRef]
  21. Shah, S.; Hussain, S.; Din, S.T.U.; Shahid, A.; Amu-Darko, J.N.O.; Wang, M.; Tianyan, Y.; Liu, G.; Qiao, G. A review on In2O3 nanostructures for gas sensing applications. J. Environ. Chem. Eng. 2024, 12, 112538. [Google Scholar] [CrossRef]
  22. Sui, N.; Zhang, P.; Zhou, T.; Zhang, T. Selective ppb-level ozone gas sensor based on hierarchical branch-like In2O3 nanostructure. Sens. Actuators B Chem. 2021, 336, 129612. [Google Scholar] [CrossRef]
  23. Liang, X.; Zhang, J.; Du, L.; Zhang, M. Effect of resonant tunneling modulation on ZnO/In2O3 heterojunction nanocomposite in efficient detection of NO2 gas at room temperature. Sens. Actuators B Chem. 2021, 329, 129230. [Google Scholar] [CrossRef]
  24. Yang, S.; Lei, G.; Xu, H.; Lan, Z.; Wang, Z.; Gu, H. Metal Oxide Based Heterojunctions for Gas Sensors: A Review. Nanomaterials 2021, 11, 1026. [Google Scholar] [CrossRef] [PubMed]
  25. Gai, L.-Y.; Lai, R.-P.; Dong, X.-H.; Wu, X.; Luan, Q.-T.; Wang, J.; Lin, H.-F.; Ding, W.-H.; Wu, G.-L.; Xie, W.-F. Recent advances in ethanol gas sensors based on metal oxide semiconductor heterojunctions. Rare Met. 2022, 41, 1818–1842. [Google Scholar] [CrossRef]
  26. Cheng, C.; Chen, C.; Zhang, H.; Zhang, Y. Preparation and study of ammonia gas sensor based on ZnO/CuO heterojunction with high performance at room temperature. Mater. Sci. Semicond. Process. 2022, 146, 106700. [Google Scholar] [CrossRef]
  27. Zoromba, M.S.; Alharbi, F.; Al-Hossainy, A.F.; Abdel-Aziz, M.H. Preparation of hybrid conducting polymers blend nanocomposite for energy conversion using experimental data and TD-DFT/DMOl3 computations. J. Mater. Res. Technol. 2023, 23, 2852–2867. [Google Scholar] [CrossRef]
  28. Haas, P.; Tran, F.; Blaha, P. Calculation of the lattice constant of solids with semilocal functionals. Phys. Rev. B 2009, 79, 085104. [Google Scholar] [CrossRef]
  29. Bučko, T.; Lebègue, S.; Hafner, J.; Ángyán, J.G. Tkatchenko-Scheffler van der Waals correction method with and without self-consistent screening applied to solids. Phys. Rev. B 2013, 87, 064110. [Google Scholar] [CrossRef]
  30. Muvva, C.; Murugan, N.A.; Subramanian, V. Assessment of Amyloid Forming Tendency of Peptide Sequences from Amyloid Beta and Tau Proteins Using Force-Field, Semi-Empirical, and Density Functional Theory Calculations. Int. J. Mol. Sci. 2021, 22, 3244. [Google Scholar] [CrossRef]
  31. Gui, Y.; Lin, Y.; Ji, C.; Luo, P.; Chen, X. Density Functional Theory Calculations for the Adsorption Property of Hazardous Industrial Gasses on Transition-Metal-Modified MoS2 Nanosheets. ACS Appl. Nano Mater. 2022, 5, 11111–11118. [Google Scholar] [CrossRef]
  32. Wang, Y.; Gui, Y.; Ji, C.; Tang, C.; Zhou, Q.; Li, J.; Zhang, X. Adsorption of SF6 decomposition components on Pt3-TiO2(1 0 1) surface: A DFT study. Appl. Surf. Sci. 2018, 459, 242–248. [Google Scholar] [CrossRef]
Figure 1. (a,b) The structure of In2O3-ZnO heterojunction, (c) the band structure of In2O3-ZnO heterojunction.
Figure 1. (a,b) The structure of In2O3-ZnO heterojunction, (c) the band structure of In2O3-ZnO heterojunction.
Ijms 25 08009 g001
Figure 2. The gasses adsorption structures of (a1,a2) H2S, (b1,b2) SO2, (c1,c2) SOF2, and (d1,d2) SO2F2 on In2O3-ZnO heterojunction.
Figure 2. The gasses adsorption structures of (a1,a2) H2S, (b1,b2) SO2, (c1,c2) SOF2, and (d1,d2) SO2F2 on In2O3-ZnO heterojunction.
Ijms 25 08009 g002
Figure 3. The TDOS and PDOS of (a1,a2) H2S, (b1,b2) SO2, (c1,c2) SOF2, and (d1,d2) SO2F2 on In2O3-ZnO heterojunction.
Figure 3. The TDOS and PDOS of (a1,a2) H2S, (b1,b2) SO2, (c1,c2) SOF2, and (d1,d2) SO2F2 on In2O3-ZnO heterojunction.
Ijms 25 08009 g003
Figure 4. The CDD of (a) H2S, (b) SO2, (c) SOF2, and (d) SO2F2 on In2O3-ZnO heterojunction.
Figure 4. The CDD of (a) H2S, (b) SO2, (c) SOF2, and (d) SO2F2 on In2O3-ZnO heterojunction.
Ijms 25 08009 g004
Figure 5. Calculated frontier molecular orbitals, HOMO and LUMO, before and after gas adsorption on In2O3-ZnO heterojunction.
Figure 5. Calculated frontier molecular orbitals, HOMO and LUMO, before and after gas adsorption on In2O3-ZnO heterojunction.
Ijms 25 08009 g005
Figure 6. (a,b) The model of TiO2-ZnO heterojunction, (c) the band structure of TiO2-ZnO heterojunction.
Figure 6. (a,b) The model of TiO2-ZnO heterojunction, (c) the band structure of TiO2-ZnO heterojunction.
Ijms 25 08009 g006
Figure 7. The gasses adsorption structures of (a1,a2) H2S, (b1,b2) SO2, (c1,c2) SOF2, and (d1,d2) SO2F2 on TiO2-ZnO heterojunction.
Figure 7. The gasses adsorption structures of (a1,a2) H2S, (b1,b2) SO2, (c1,c2) SOF2, and (d1,d2) SO2F2 on TiO2-ZnO heterojunction.
Ijms 25 08009 g007
Figure 8. The TDOS and PDOS of (a1,a2) H2S, (b1,b2) SO2, (c1,c2) SOF2, and (d1,d2) SO2F2 on TiO2-ZnO heterojunction.
Figure 8. The TDOS and PDOS of (a1,a2) H2S, (b1,b2) SO2, (c1,c2) SOF2, and (d1,d2) SO2F2 on TiO2-ZnO heterojunction.
Ijms 25 08009 g008
Figure 9. The CDD of (a) H2S, (b) SO2, (c) SOF2, and (d) SO2F2 on TiO2-ZnO heterojunction.
Figure 9. The CDD of (a) H2S, (b) SO2, (c) SOF2, and (d) SO2F2 on TiO2-ZnO heterojunction.
Ijms 25 08009 g009
Figure 10. Calculated frontier molecular orbitals, HOMO and LUMO, before and after gas adsorption on TiO2-ZnO heterojunction.
Figure 10. Calculated frontier molecular orbitals, HOMO and LUMO, before and after gas adsorption on TiO2-ZnO heterojunction.
Ijms 25 08009 g010
Table 1. The adsorption parameters of H2S, SO2, SOF2, and SO2F2 on In2O3-ZnO heterojunction surface.
Table 1. The adsorption parameters of H2S, SO2, SOF2, and SO2F2 on In2O3-ZnO heterojunction surface.
SystemDistance (Å)Eads (eV)Qt (e)
H2S/In2O3-ZnOIn-S:2.486
H-O:1.008
−1.9240.255
SO2/In2O3-ZnOIn-O:2.190
S-O:1.561
−1.992−0.361
SOF2/In2O3-ZnOIn-S:3.687−0.4650.016
SO2F2/In2O3-ZnOIn-S:3.902−0.5010.011
Table 2. The adsorption parameters of H2S, SO2, SOF2, and SO2F2 on TiO2-ZnO heterojunction surface.
Table 2. The adsorption parameters of H2S, SO2, SOF2, and SO2F2 on TiO2-ZnO heterojunction surface.
SystemDistance (Å)Eads (eV)Qt (e)
H2S/TiO2-ZnOTi-S:2.780−0.7880.249
SO2/TiO2-ZnOS-O:1.766
Ti-O:1.958
−0.608−0.142
SOF2/TiO2-ZnOTi-S:3.084−0.4020.083
SO2F2/TiO2-ZnOS-O:3.214−0.5490.061
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

Zeng, T.; Ma, D.; Gui, Y. Gas-Sensing Performance of Metal Oxide Heterojunction Materials for SF6 Decomposition Gases: A DFT Study. Int. J. Mol. Sci. 2024, 25, 8009. https://doi.org/10.3390/ijms25158009

AMA Style

Zeng T, Ma D, Gui Y. Gas-Sensing Performance of Metal Oxide Heterojunction Materials for SF6 Decomposition Gases: A DFT Study. International Journal of Molecular Sciences. 2024; 25(15):8009. https://doi.org/10.3390/ijms25158009

Chicago/Turabian Style

Zeng, Tingting, Donglin Ma, and Yingang Gui. 2024. "Gas-Sensing Performance of Metal Oxide Heterojunction Materials for SF6 Decomposition Gases: A DFT Study" International Journal of Molecular Sciences 25, no. 15: 8009. https://doi.org/10.3390/ijms25158009

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