Next Article in Journal
Nitrogen Dioxide Optical Sensor Based on Redox-Active Tetrazolium/Pluronic Nanoparticles Embedded in PDMS Membranes
Previous Article in Journal
Preparation of Au@ZnO Nanofilms by Combining Magnetron Sputtering and Post-Annealing for Selective Detection of Isopropanol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Gas-Sensing Properties of Dissolved Gases in Insulating Material Adsorbed on SnO2–GeSe Monolayer

1
State Key Laboratory of Power Transmission Equipment and System Security and New Technology, School of Electrical Engineering, Chongqing University, Chongqing 400044, China
2
State Grid Zhejiang Electric Power Research Institute, Hangzhou 310014, China
*
Author to whom correspondence should be addressed.
Chemosensors 2022, 10(6), 212; https://doi.org/10.3390/chemosensors10060212
Submission received: 3 May 2022 / Revised: 18 May 2022 / Accepted: 2 June 2022 / Published: 5 June 2022
(This article belongs to the Section Nanostructures for Chemical Sensing)

Abstract

:
In a transformer, the insulation materials will produce different dissolved gases due to various faults in the operation of the transformer, in which C2H2, CH4, and H2 are the main dissolved gases. In this study, the adsorption characteristics of the above three gases on the SnO2–GeSe monolayer surface were discussed and analyzed based on the density functional theory. The adsorption energy, transfer charge, geometric structure parameters, electronic density of states, electronic local function, charge difference density, and recovery time were calculated and compared to characterize the gas-sensing adsorption mechanism. The results showed that the SnO2–GeSe monolayer exhibited good adsorption capacity, selectivity, and repeatability for the three characteristic dissolved gases. After adsorbing CH4 gas molecules, the conductivity of the SnO2–GeSe monolayer decreased. After adsorbing C2H2 and H2 gas molecules, the conductivity of the SnO2–GeSe monolayer increased. Therefore, the SnO2–GeSe monolayer has great application potential in the real-time monitoring of dissolved gases in insulating materials, which may become a new type of resistive gas sensor.

1. Introduction

The power system is one of the foundations of modern economic development. As the hub of the power system, the operation condition of the transformer directly affects the stability and reliability of the whole power system [1,2,3]. Therefore, identifying the early latent fault of the transformer is a key problem to be solved for the power industry. Currently, oil-immersed transformers are commonly used in power systems [4,5,6,7]. An oil-immersed transformer in the long-term operation process can incur an electrical fault or thermal fault. When these faults occur, the insulating materials in the transformer will age and decompose, resulting in some characteristic fault gases, such as C2H2, CH4, and H2 [8,9,10,11,12]. Studies have shown that detecting the composition and content of these characteristic fault gases is helpful to discover and judge the type and degree of faults. Among many methods for detecting these characteristic fault gases, chromatographic analysis is most commonly used because of its high detection sensitivity and good selectivity [13,14,15,16,17]. However, this method not only has a long detection time, high experimental environment requirements, expensive equipment, and complex operation but also can not perform real-time monitoring. Therefore, an effective low-cost method with simple operation, good cycle performance, and real-time monitoring of dissolved gases is urgently needed.
With the development of material science, two-dimensional nano-gas-sensing materials, such as graphene, germanium selenide, and boron nitride, have great application potential and prospects in the field of gas-sensing monitoring due to their large specific surface area, excellent electrical properties, and simple preparation process [18,19,20,21,22,23]. The basic principle is that there is chemical or physical adsorption between two-dimensional nanomaterials and the target gas, and the conductivity of the substrate changes significantly before and after the adsorption [24,25,26,27]. Among many two-dimensional nano-gas sensing materials, germanium selenide not only is rich in reserves, low in price, and environmentally friendly but also has direct bandgap and small-carrier effective mass, which is expected to become a new generation of gas-sensing materials with high sensitivity, high selectivity, and good cycle characteristics [28,29,30]. Studies have shown that germanium selenide doped with atoms, clusters, or metal oxides has a stronger gas adsorption capacity than intrinsic materials [31,32,33,34]. As an interesting n-type semiconductor, SnO2 has attracted worldwide researchers due to its wide application prospects. Studies have shown that SnO2-doped semiconductor materials can effectively improve the adsorption effect for specific gases. The gas sensor prepared by it has appropriate working temperature and good cycle characteristics. Therefore, the characteristic fault gas sensor of germanium selenide doped with SnO2 is of great significance for detecting the early latent fault of a transformer. In recent years, the simulation calculation of gas adsorption theory based on density functional theory has been widely used. Many previous studies on gas adsorption that cannot be completed due to complex experimental conditions can carry out theoretical simulation through these technologies to better understand the properties of materials and many complex chemical reaction mechanisms. Based on the first-principles density-functional theory, the electronic properties and microscopic mechanism of gas sensing materials can be obtained, and, on this basis, the macroscopic physical and chemical properties of gas-sensing materials can be deduced [35,36,37,38,39,40,41,42].
In this study, based on the density functional theory (DFT), the adsorption parameters and electrical properties of transformer characteristic fault gases (C2H2, CH4, and H2) on the SnO2-doped GeSe monolayer were analyzed from a microscopic point of view, and the adsorption characteristics of each adsorption system were explored in order to provide a theoretical basis for the synthesis of the SnO2–GeSe monolayer and the preparation of gas sensors applied to the monitoring of the transformer characteristic fault gas.

2. Materials and Methods

Based on the DFT method, all calculations were completed in the Dmol3 and CASTEP modules in Materials Studio [43,44,45,46,47,48,49]. The calculation parameters and methods were set according to the structural model and doping properties, which was very important for accurately and efficiently obtaining the optimal geometric structure and electronic properties of each system. The construction of a reliable exchange–correlation function was the key that directly affected the accuracy of the calculation results. Considering that all electrons in the actual atomic and molecular systems are mainly orientalized, the Perdew–Burke-Ernzehof (PBE) functional of generalized gradient approximation (GGA) was used to deal with the exchange–correlation between electrons, and double numerical plus polarization (DNP) was used to calculate the electron pseudopotential. To accurately obtain the physical and chemical parameters of each system, Monkhorst-Pack was set at 8 × 8 × 1. Considering the van der Waals force and long-distance interaction between the SnO2 unit doping and gas adsorption, the DFT-D method was used to analyze all models. The charge transfer between the substrate and adsorbed gas molecules was calculated by the Mulliken charge analysis method.
In this study, 4 × 4 × 1 superlattice cells were used to construct an intrinsic GeSe monolayer, including 18 Ge atoms and 18 Se atoms. In order to prevent the interaction between adjacent units, a vacuum layer of 20 Å was set. Meanwhile, three gas molecular models of C2H2, CH4, and H2 were constructed and optimized. In order to obtain the optimal SnO2-GeSe monolayer, the SnO2 unit was then doped on the GeSe monolayer surface from different angles and distances. After obtaining the most stable SnO2–GeSe monolayer, the above three gas molecules were placed in different positions near the substrate, and the adsorption calculation was carried out. Finally, the optimal adsorption configurations and parameters of the three gas adsorption systems were obtained, including binding energy (Eb), band energy, the shortest adsorption distance (dsub/gas), absorption energy (Eads), charge transfer (ΔQ), the total density of states (TDOS), electron localization function (ELF), charge density difference (CDD), and recovery time (τ). In order to improve the calculation accuracy and simplify the calculation amount, the energy convergence standard, self-consistent field convergence threshold, maximum force, and maximum displacement were set to 1 × 10−5 Ha, 1 × 10−6 Ha, 2 × 10−3 Ha/Å, and 5 × 10−3 Å, respectively. Considering the influence of a high-humidity environment on the gas sensor, we adopted a higher dielectric constant in Dmol3 solvent. In the aqueous solution environment at 298 K, ε = 78.5 C²/(N-M²).

3. Results and Discussion

Figure 1 shows the optimized geometric structure and adsorption system. In order to compare the stability of the SnO2–GeSe monolayer formed by various doping methods, the geometric structure parameters and binding energy of intrinsic GeSe monolayer and SnO2–GeSe monolayer were calculated. The calculation formula of binding energy (Eb) in this study is as follows:
Eb = ESnO2–GeSeEGeSeESnO2
where ESnO2–GeSe, EGeSe, and ESnO2 represent the energy of the SnO2–GeSe monolayer, pristine GeSe monolayer, and SnO2 unit, respectively.
It is generally believed that the formation energy of a spontaneous doping reaction is negative, and the larger the inverse of the formation energy is, the more stable the doping structure is [50,51,52,53]. It is calculated that the formation energy of the most stable SnO2–GeSe monolayer is –5.600 eV in many doping systems, as shown in Figure 1(b1,b2). The larger adsorption energy indicated that the probability of forming this doped structure was the largest in the actual experimental preparation process. At the same time, the bond length of the doped GeSe monolayer changed from 2.616 to 2.667 Å. The change in bond angles was also mild. Moderate geometric structure change is guaranteed to enhance the gas-sensitive adsorption capacity and stability of the substrate. In addition, since the electronic structure has a great influence on the adsorption performance of the substrate, the band structure of the substrate before and after doping was calculated, as shown in Figure 1(a3–b3). The results show that the top of the valence band was still near the Fermi level after doping, while the position of the conduction band was lower. This reduced the band gap energy from 1.007 to 0.922 eV. According to previous studies, the decrease of band gap energy helps to enhance the gas adsorption capacity of the substrate [54,55,56,57]. This is because the electrons near the forbidden band are more likely to transit from the valence band to the conduction band due to the interaction between the substrate and the gas, thereby changing the conductivity of the substrate, and the specific voltage or current values of different target gases can be obtained macroscopically. Therefore, the adsorption of dissolved gases in the subsequent exploration of insulating materials is based on this SnO2–GeSe monolayer.
The optimized gas adsorption configurations were obtained through optimization calculation, as shown in Figure 1(c1–e2). In order to analyze the adsorption process of the SnO2–GeSe monolayer for each gas, the adsorption energy (Eads), transfer charge (ΔQ), and geometric structure parameters of each adsorption system were calculated. Adsorption energy is a physical and chemical parameter to evaluate the adsorption force of gas molecules on the SnO2–GeSe monolayer. The calculation formula of adsorption energy in this study was:
Eads = EGas/SnO2–GeSeEGasESnO2–GeSe
where EGas/SnO2–GeSe, EGas, and ESnO2–GeSe represent the energy of adsorption systems, gas molecule, and the SnO2–GeSe monolayer, respectively. The calculated adsorption energies for the interaction between C2H2, CH4, and H2 and the SnO2–GeSe monolayer were −0.498 eV, −0.220 eV, and −0.152 eV, respectively.
The calculation formula of charge transfer (ΔQ) in this study was as follows:
ΔQ = Q1Q2
where Q1 and Q2 represent the total charge of gas molecules after and before adsorption, respectively. The calculated transfer charges of CH4, C2H2, and H2 before and after adsorption with the SnO2–GeSe monolayer were −0.262 e, 0.07 e, and 0.027 e, respectively.
The adsorption energy of the three adsorption systems was negative, indicating that the three adsorption reactions were exothermic without external energy input. The SnO2–GeSe monolayer could spontaneously adsorb C2H2, CH4, and H2 gas molecules. At the same time, the adsorption energy of the three adsorption systems was greater than −0.6 eV, indicating that the adsorption between the three gas molecules and SnO2–GeSe monolayer was physical adsorption, and the intensity was C2H2 > CH4 > H2. The moderate adsorption energy was conducive to gas adsorption and desorption, which guaranteed the cycle performance and stability of the gas sensor. In addition, the three gas molecules were adsorbed near the SnO2 unit (all H atoms in the gas molecules and O atoms in the SnO2 unit), and the shortest distance (dsub/gas) from the substrate was 2.560 Å, 2.632 Å, and 2.805 Å, respectively. This showed that with SnO2 as an n-type semiconductor, doping would significantly enhance the gas adsorption capacity of the substrate. The adsorption energy of C2H2 gas molecules on the SnO2–GeSe monolayer surface was the largest, indicating that the adsorption system had the strongest adsorption effect and the most stable adsorption configuration among the three adsorption systems. This also explained why C2H2 gas molecules were the closest to the SnO2–GeSe monolayer in three adsorption systems. Charge transfer refers to the amount of charge transferred from gas molecules to the SnO2–GeSe monolayer during the adsorption of a single molecule. The negative transfer charge of the CH4 molecule indicated that the CH4 gas acted as an electron acceptor in the adsorption process, and the SnO2–GeSe monolayer acted as an electron donor. This led to an increase in the band gap of the adsorption system. On the contrary, in the two adsorption systems, C2H2 gas and H2 gas acted as electron donors. The SnO2–GeSe monolayer acted as an electron acceptor. The absolute value of transfer charge was CH4 > C2H2 > H2, indicating that the charge transfer between the CH4 gas molecules and SnO2–GeSe monolayer was the most intense before and after adsorption. This led to significant changes in conductivity before and after the adsorption system. Therefore, from the perspective of adsorption energy, transfer charge, and geometric structure parameters, the SnO2–GeSe monolayer can distinguish and efficiently monitor the three gas molecules.
In order to further explore the changes in the electrical properties of each system, the total density of state (TDOS) before and after the SnO2–GeSe monolayer doping and adsorption was calculated, as shown in Figure 2. Figure 2a shows that after doping the SnO2 unit, TDOS moved to the left as a whole (that is, moved to the low-level direction), and the formed structure was more stable. Continuous TDOS also indicated that the SnO2–GeSe monolayer had good stability. It was still composed of the upper valence band, lower valence band, and conduction band and was similar to the TDOS of the intrinsic GeSe monolayer, indicating that the crystal structure of the GeSe monolayer itself was not changed after doping. This is similar to the slightly varying geometric structure parameters previously obtained. The obvious change on the right side of the Fermi level indicated that SnO2 unit doping will reduce the band gap of the GeSe monolayer. This is similar to the previous conclusion obtained by band energy. A smaller band gap helps electrons transfer between the valence band and conduction band after gas adsorption. Therefore, the doping of the SnO2 unit was conducive to improving the conductivity and gas-sensing response of the crystal plane. In the three adsorption systems, the TDOS after adsorption has different degrees of left shift (that is, to the low-energy-level direction), as shown in Figure 2b–d. Compared with the other two adsorption systems, the TDOS of the H2 adsorption system had a large gap, and its main change was far below the Fermi level. However, this had little contribution to the change of conductivity, which was also consistent with the small adsorption energy of H2 gas molecules in the adsorption process. Figure 2b shows that after the adsorption of C2H2 gas molecules, a new peak appeared near −10 eV. This may be related to the adsorption of C2H2 gas molecules on the SnO2–GeSe monolayer. Near the Fermi level, the TDOS changes of the C2H2 and CH4 gas adsorption systems were more obvious, and the changes of C2H2 gas molecules after adsorption were more obvious. In addition, the hybrid peaks of TDOS appeared in the two adsorption systems, indicating that the SnO2–GeSe monolayer had an obvious adsorption effect on C2H2 and CH4 gas molecules. This was mainly because the C2H2 and CH4 molecules increased the number of carriers in the adsorption process. Due to the more active movement of electrons here, the conductivity of the system changed dramatically. This also confirmed our previous adsorption strength and selectivity C2H2 > CH4 > H2 based on physical and chemical parameters such as adsorption energy, adsorption shortest distance, and transfer charge.
In order to further explore the interaction mechanism between the three gas molecules and the SnO2–GeSe monolayer surface, the electron localization function (ELF) and charge difference density (CDD) were calculated as shown in Figure 3 and Figure 4. The ELF of the three adsorption systems showed that the yellow region around the gas molecules tended to merge with the SnO2 unit. This showed that there was obvious adsorption between gas molecules and the SnO2 unit. In addition, the blue region between the yellow region also proved that the three adsorption reactions did not generate new chemical bonds, which were physical adsorption. This was consistent with the previous analysis of TDOS conclusions. The CDD diagrams of the three adsorption systems showed that there were a large number of electron dissipation and electron aggregation regions between the gas molecules and SnO2–GeSe monolayer. This indicated that charge transfer mainly occurred between the gas molecules and SnO2–GeSe monolayer. In the CH4 adsorption system, there was an obvious electron dissipation zone between the gas molecule and SnO2 unit, and the aggregation zone was mainly concentrated near the CH4 gas molecule. The results also confirm that CH4 obtained electrons in the adsorption reaction, which was consistent with the charge transfer (−0.262 e) obtained by the Mulliken charge analysis method. In C2H2 and H2 adsorption systems, the dissipation region was mainly concentrated on the C2H2 and H2 gas molecules, and the aggregation region was mainly concentrated on the SnO2 unit, indicating that the SnO2–GeSe monolayer exhibited the obtained electron characteristics in these two adsorption reactions, and C2H2 > H2. Therefore, from the perspective of ELF and CDD, the SnO2–GeSe monolayer had moderate adsorption and strong selectivity for the three gases.
In order to further explore the practicability of the SnO2–GeSe monolayer gas sensor, the recovery time of gas molecules adsorbed on the SnO2–GeSe monolayer was calculated. Recovery time is an important measurement index of gas sensors. A shorter recovery time contributes to the real-time monitoring of dissolved gases in insulation materials. The calculation formula of recovery time in this study was:
τ = v0−1 exp(−Eads/kT)
where v0, k, and T represent the attempt frequency (1 × 1012 s−1), Boltzmann constant (1.38 × 10−23 J/K), and thermodynamic temperature, respectively.
After calculation, the three gases all had a short recovery time in SnO2–GeSe monolayer as shown in Figure 5. This ensured the excellent performance of the SnO2–GeSe monolayer gas sensor, which meant that the sensor could be reused. Combined with the previous gas-sensing parameters such as transfer charge, TDOS, and CDD, this proved that the SnO2–GeSe monolayer has great application potential as a gas-sensing material.

4. Conclusions

In this study based on the density functional theory, the SnO2–GeSe monolayer model was established to optimize the gas adsorption system model, and a series of microscopic parameters characterizing the adsorption process, such as the adsorption energy, transfer charge amount, geometric structure parameters, electronic density of states, electronic local function, charge difference density, and recovery time of three transformer characteristic fault gases (C2H2, CH4, and H2) acting on the SnO2–GeSe monolayer surface, were calculated in detail. The main conclusions are as follows:
(1)
The SnO2 unit’s doping on the GeSe monolayer surface effectively changed the electron cloud distribution, and it became the catalytic activity center for the adsorption of gas molecules on the GeSe monolayer surface, reducing the gas adsorption barrier on the substrate. At the same time, the SnO2 unit provided adsorption sites for gas molecules and promoted the interaction between gas molecules and the SnO2–GeSe monolayer system.
(2)
Three kinds of adsorption systems were optimized based on the density functional theory, and the gas-sensing mechanism of the three gases interacting with the SnO2–GeSe monolayer was obtained. The kind of gas adsorption was physical adsorption, and the intensity and selectivity were C2H2 > CH4 > H2. The recovery time of the order was the opposite. When gas molecules adsorbed on the SnO2–GeSe monolayer surface, C2H2, and H2 gas molecules acted as electron donors, CH4 as an electron acceptor, and the SnO2–GeSe monolayer acted as electron acceptor and electron donor. Macroscopically, there was a significant difference in the system resistance value changes. Therefore, this study provides a theoretical basis for the preparation of the SnO2–GeSe monolayer, explored the potential of the SnO2–GeSe monolayer as a gas sensor for dissolved gases or other uses in insulating materials, and provided a new method of and idea for the development of other gas sensors.

Author Contributions

Conceptualization, L.-Y.G.; methodology, L.-Y.G.; software, Z.H.; validation, S.L.; formal analysis, S.L.; investigation, Z.Y.; data curation, L.J.; writing—original draft preparation, L.-Y.G.; writing—review and editing, Y.T.; visualization, Y.T.; supervision, Y.T.; funding acquisition, Z.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program 2018YFB2100100, the National Natural Science Foundation of China 51907012 and Grant SCITLAB(1009) of Intelligent Terminal Key Laboratory of SiChuan Province.

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. Sun, J.; Xu, M.; Cespedes, M.; Kauffman, M. Data Center Power System Stability—Part I: Power Supply Impedance Modeling. CSEE J. Power Energy Syst. 2022, 8, 403–419. [Google Scholar]
  2. Haes Alhelou, H.; Hamedani-Golshan, M.E.; Njenda, T.C.; Siano, P. A survey on power system blackout and cascading events: Research motivations and challenges. Energies 2019, 12, 682. [Google Scholar] [CrossRef] [Green Version]
  3. Tang, Z.; Yang, Y.; Blaabjerg, F. Power electronics: The enabling technology for renewable energy integration. CSEE J. Power Energy Syst. 2021, 8, 39–52. [Google Scholar]
  4. Yang, J.; Lu, W.; Liu, X. Prediction of Top Oil Temperature for Oil-immersed Transformers Based on PSO-LSTM. In Proceedings of the 2021 4th International Conference on Energy, Electrical and Power Engineering (CEEPE), Chengdu, China, 20–23 September 2017; pp. 278–283. [Google Scholar]
  5. Zheng, H.; Zhang, C.; Zhang, Y.; Liu, J.; Zhang, E.; Shi, Z.; Shao, G.; Shi, K.; Guo, J.; Zhang, C. Optimization of ethanol detection by automatic headspace method for cellulose insulation aging of oil-immersed transformers. Polymers 2020, 12, 1567. [Google Scholar] [CrossRef]
  6. Liu, J.; Fan, X.; Zhang, C.; Lai, C.S.; Zhang, Y.; Zheng, H.; Lai, L.L.; Zhang, E. Moisture diagnosis of transformer oil-immersed insulation with intelligent technique and frequency-domain spectroscopy. IEEE Trans. Ind. Inform. 2020, 17, 4624–4634. [Google Scholar] [CrossRef]
  7. Arguence, O.; Cadoux, F. Sizing power transformers in power systems planning using thermal rating. Int. J. Electr. Power Energy Syst. 2020, 118, 105781. [Google Scholar] [CrossRef]
  8. Ayalew, Z.; Kobayashi, K.; Matsumoto, S.; Kato, M. Dissolved gas analysis (DGA) of arc discharge fault in transformer insulation oils (ester and mineral oils). In Proceedings of the 2018 IEEE Electrical Insulation Conference (EIC), San Antonio, TX, USA, 17–20 June 2018; pp. 150–153. [Google Scholar]
  9. Sami, S.M.; Bhuiyan, M.I.H. An EMD-based Method for the Detection of Power Transformer Faults with a Hierarchical Ensemble Classifier. In Proceedings of the 2020 11th International Conference on Electrical and Computer Engineering (ICECE), Dhaka, Bangladesh, 17–19 December 2020; pp. 206–209. [Google Scholar]
  10. Mharakurwa, E.T.; Nyakoe, G.N.; Akumu, A. Power transformer fault severity estimation based on dissolved gas analysis and energy of fault formation technique. J. Electr. Comput. Eng. 2019, 2019, 9674054. [Google Scholar] [CrossRef] [Green Version]
  11. Rezaie, S.; Bafghi, Z.G.; Manavizadeh, N.; Kordmahale, S.B. Highly Sensitive Detection of Dissolved Gases in Transformer Oil With Carbon-Doped ZnO Nanotube: A DFT Study. IEEE Sens. J. 2021, 22, 82–89. [Google Scholar] [CrossRef]
  12. Naganathan, G.S.; Senthilkumar, M.; Aiswariya, S.; Muthulakshmi, S.; Riyasen, G.S.; Priyadharshini, M.M. Internal fault diagnosis of power transformer using artificial neural network. Mater. Today Proc. 2021; in press. [Google Scholar]
  13. Rodríguez, J.; Contreras, J.; Gaytán, C. Evaluation and Interpretation of Dissolved Gas Analysis of Soybean-Based Natural Ester Insulating Liquid. IEEE Trns. Dielectr. Electr. Insul. 2021, 28, 1343–1348. [Google Scholar] [CrossRef]
  14. Zope, N.; Ali, S.I.; Padmanaban, S.; Bhaskar, M.S.; Mihet-Popa, L. Analysis of 132kV/33kV 15MVA power transformer dissolved gas using transport-X Kelman Kit through Duval’s triangle and Roger’s Ratio prediction. In Proceedings of the 2018 IEEE International Conference on Industrial Technology (ICIT), Lyon, France, 20–22 February 2018; pp. 1160–1164. [Google Scholar]
  15. Pei, L.; Hongbo, L.; Nannan, G.; Yan, Z.; Yanyan, Z.; Ying, P.; Qinghua, Y. Case Analysis of 220 kV Oil-immersed Current Transformer Defect. In Proceedings of the 2020 5th Asia Conference on Power and Electrical Engineering (ACPEE), Chengdu, China, 4–7 June 2020; pp. 2178–2182. [Google Scholar]
  16. Shutenko, O.; Kulyk, O. Analysis of gas content in oil-filled equipment with low energy density discharges. Int. J. Electr. Eng. Inform. 2020, 12, 258–277. [Google Scholar] [CrossRef]
  17. Balaraman, S.; Madavan, R.; Vedhanayaki, S.; Saroja, S.; Srinivasan, M.; Stonier, A.A. Fault Diagnosis and Asset Management of Power Transformer Using Adaptive Boost Machine Learning Algorithm. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1055, 012133. [Google Scholar] [CrossRef]
  18. Hashtroudi, H.; Yu, A.; Juodkazis, S.; Shafiei, M. Two-Dimensional Dy2O3-Pd-PDA/rGO Heterojunction Nanocomposite: Synergistic Effects of Hybridisation, UV Illumination and Relative Humidity on Hydrogen Gas Sensing. Chemosensors 2022, 10, 78. [Google Scholar] [CrossRef]
  19. Li, T.; Yin, W.; Gao, S.; Sun, Y.; Xu, P.; Wu, S.; Kong, H.; Yang, G.; Wei, G. The Combination of Two-Dimensional Nanomaterials with Metal Oxide Nanoparticles for Gas Sensors: A Review. Nanomaterials 2022, 12, 982. [Google Scholar] [CrossRef] [PubMed]
  20. Lee, E.; Yoon, Y.S.; Kim, D.-J. Two-dimensional transition metal dichalcogenides and metal oxide hybrids for gas sensing. ACS Sens. 2018, 3, 2045–2060. [Google Scholar] [CrossRef]
  21. Wang, C.; Li, R.; Feng, L.; Xu, J. The SnO2/MXene Composite Ethanol Sensor Based on MEMS Platform. Chemosensors 2022, 10, 109. [Google Scholar] [CrossRef]
  22. Faglia, G.; Ferroni, M.; Dang, T.T.L.; Donarelli, M.; Rigoni, F.; Baratto, C. Vertically coupling ZnO nanorods onto MoS2 flakes for optical gas sensing. Chemosensors 2020, 8, 19. [Google Scholar] [CrossRef] [Green Version]
  23. Tian, X.; Yao, L.; Cui, X.; Zhao, R.; Chen, T.; Xiao, X.; Wang, Y. A two-dimensional Ti3C2TX MXene@TiO2/MoS2 heterostructure with excellent selectivity for the room temperature detection of ammonia. J. Mater. Chem. A 2022, 10, 5505–5519. [Google Scholar] [CrossRef]
  24. Li, Q.; Li, Y.; Zeng, W. Preparation and Application of 2D MXene-Based Gas Sensors: A Review. Chemosensors 2021, 9, 225. [Google Scholar] [CrossRef]
  25. Maity, A.; Raychaudhuri, A.; Ghosh, B. High sensitivity NH3 gas sensor with electrical readout made on paper with perovskite halide as sensor material. Sci. Rep. 2019, 9, 7777. [Google Scholar] [CrossRef] [Green Version]
  26. Hou, C.; Tai, G.; Liu, Y.; Liu, X. Borophene gas sensor. Nano Res. 2022, 15, 2537–2544. [Google Scholar] [CrossRef]
  27. Qiao, X.; Xu, Y.; Yang, K.; Ma, J.; Li, C.; Wang, H.; Jia, L. Mo doped BiVO4 gas sensor with high sensitivity and selectivity towards H2S. Chem. Eng. J. 2020, 395, 125144. [Google Scholar] [CrossRef]
  28. Shakeel, A.; Rizwan, K.; Farooq, U.; Iqbal, S.; Altaf, A.A. Advanced polymeric/inorganic nanohybrids: An integrated platform for gas sensing applications. Chemosphere 2022, 294, 133772. [Google Scholar] [CrossRef]
  29. Sosa, A.N.; Santana, J.E.; Miranda, Á.; Pérez, L.A.; Rurali, R.; Cruz-Irisson, M. Transition metal-decorated germanene for NO, N2 and O2 sensing: A DFT study. Surf. Interfaces 2022, 30, 101886. [Google Scholar] [CrossRef]
  30. Zhou, K.L.; Wang, Z.; Han, C.B.; Ke, X.; Wang, C.; Jin, Y.; Zhang, Q.; Liu, J.; Wang, H.; Yan, H. Platinum single-atom catalyst coupled with transition metal/metal oxide heterostructure for accelerating alkaline hydrogen evolution reaction. Nat. Commun. 2021, 12, 3783. [Google Scholar] [CrossRef] [PubMed]
  31. Liu, C.; Hu, Y.; Liu, F.; Liu, H.; Xu, X.; Xue, Y.; Zhang, J.; Li, Y.; Tang, C. Electronic structure modulation of CoSe2 nanowire arrays by tin doping toward efficient hydrogen evolution. Int. J. Hydrog. Energy 2021, 46, 17133–17142. [Google Scholar] [CrossRef]
  32. Fan, H.; Mao, P.; Sun, H.; Wang, Y.; Mofarah, S.S.; Koshy, P.; Arandiyan, H.; Wang, Z.; Liu, Y.; Shao, Z. Recent advances of metal telluride anodes for high-performance lithium/sodium-ion batteries. Mater. Horiz. 2022, 9, 524–546. [Google Scholar] [CrossRef]
  33. Hu, Z.; Ding, Y.; Hu, X.; Zhou, W.; Yu, X.; Zhang, S. Recent progress in 2D group IV–IV monochalcogenides: Synthesis, properties and applications. Nanotechnology 2019, 30, 252001. [Google Scholar] [CrossRef]
  34. Liu, X.; Jia, S.; Yang, M.; Tang, Y.; Wen, Y.; Chu, S.; Wang, J.; Shan, B.; Chen, R. Activation of subnanometric Pt on Cu-modified CeO2 via redox-coupled atomic layer deposition for CO oxidation. Nat. Commun. 2020, 11, 4240. [Google Scholar] [CrossRef]
  35. Li, Q.; Liu, Y.; Chen, D.; Miao, J.; Zhi, X.; Deng, S.; Lin, S.; Jin, H.; Cui, D. Nitrogen Dioxide Gas Sensor Based on Ag-Doped Graphene: A First-Principle Study. Chemosensors 2021, 9, 227. [Google Scholar] [CrossRef]
  36. Wang, G.; Yu, J.; Zheng, K.; Huang, Y.; Li, X.; Chen, X.; Tao, L.-Q. A monolayer composite of h-BN doped by a nano graphene domain: As sensitive material for SO2 gas detection. IEEE Electron Device Lett. 2020, 41, 1404–1407. [Google Scholar] [CrossRef]
  37. Peng, Z.; Tao, L.-Q.; Wang, G.; Zhang, F.; Sun, H.; Zhu, C.; Zou, S.; Yu, J.; Chen, X. The promotion of sulfuric vacancy in two-dimensional molybdenum disulfide on the sensing performance of SF6 decomposition components. Appl. Surf. Sci. 2022, 571, 151377. [Google Scholar] [CrossRef]
  38. Wang, G.; Zheng, K.; Huang, Y.; Yu, J.; Wu, H.; Chen, X.; Tao, L.-Q. An investigation of the positive effects of doping an Al atom on the adsorption of CO2 on BN nanosheets: A DFT study. Phys. Chem. Chem. Phys. 2020, 22, 9368–9374. [Google Scholar] [CrossRef] [PubMed]
  39. Sun, H.; Tao, L.-Q.; Li, T.; Gao, X.; Wang, G.; Peng, Z.; Zhu, C.; Zou, S.; Gui, Y.; Xia, S.-Y. Sensing Characteristics of Toxic C₄F₇N Decomposition Products on Metallic-Nanoparticle Co-Doped BN Monolayer: A First Principles Study. IEEE Sens. J. 2021, 21, 13082–13089. [Google Scholar] [CrossRef]
  40. Peng, Z.; Tao, L.-Q.; Zheng, K.; Yu, J.; Wang, G.; Sun, H.; Zhu, C.; Zou, S.; Chen, X. Gas Sensor Based on Semihydrogenated and Semifluorinated h-BN for SF₆ Decomposition Components Detection. IEEE Trans. Electron Devices 2021, 68, 1878–1885. [Google Scholar] [CrossRef]
  41. Xia, S.-Y.; Tao, L.-Q.; Jiang, T.; Sun, H.; Li, J. Rh-doped h-BN monolayer as a high sensitivity SF6 decomposed gases sensor: A DFT study. Appl. Surf. Sci. 2021, 536, 147965. [Google Scholar] [CrossRef]
  42. Zhang, F.; Qiu, J.; Guo, H.; Wu, L.; Zhu, B.; Zheng, K.; Li, H.; Wang, Z.; Chen, X.; Yu, J. Theoretical investigations of novel Janus Pb2SSe monolayer as a potential multifunctional material for piezoelectric, photovoltaic, and thermoelectric applications. Nanoscale 2021, 13, 15611–15623. [Google Scholar] [CrossRef] [PubMed]
  43. Badr, E.A.; Bedair, M.; Shaban, S.M. Adsorption and performance assessment of some imine derivatives as mild steel corrosion inhibitors in 1.0 M HCl solution by chemical, electrochemical and computational methods. Mater. Chem. Phys. 2018, 219, 444–460. [Google Scholar] [CrossRef]
  44. Zhang, X.; Chen, Z.; Chen, D.; Cui, H.; Tang, J. Adsorption behaviour of SO2 and SOF2 gas on Rh-doped BNNT: A DFT study. Mol. Phys. 2020, 118, e1580394. [Google Scholar] [CrossRef]
  45. Zhao, J.; Buldum, A.; Han, J.; Lu, J.P. Gas molecule adsorption in carbon nanotubes and nanotube bundles. Nanotechnology 2002, 13, 195. [Google Scholar] [CrossRef]
  46. Azam, M.A.; Alias, F.M.; Tack, L.W.; Seman, R.N.A.R.; Taib, M.F.M. Electronic properties and gas adsorption behaviour of pristine, silicon-, and boron-doped (8, 0) single-walled carbon nanotube: A first principles study. J. Mol. Graph. 2017, 75, 85–93. [Google Scholar] [CrossRef]
  47. Shukri, M.; Saimin, M.; Yaakob, M.; Yahya, M.; Taib, M. Structural and electronic properties of CO and NO gas molecules on Pd-doped vacancy graphene: A first principles study. Appl. Surf. Sci. 2019, 494, 817–828. [Google Scholar] [CrossRef]
  48. Segall, M.; Lindan, P.J.; Probert, M.A.; Pickard, C.J.; Hasnip, P.J.; Clark, S.; Payne, M. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys.-Condes. Matter 2002, 14, 2717. [Google Scholar] [CrossRef]
  49. Delley, B. Time dependent density functional theory with DMol3. J. Phys. Condes. Matter 2010, 22, 384208. [Google Scholar] [CrossRef] [PubMed]
  50. Zhang, D.; Cao, Y.; Yang, Z.; Wu, J. Nanoheterostructure construction and DFT study of Ni-doped In2O3 nanocubes/WS2 hexagon nanosheets for formaldehyde sensing at room temperature. ACS Appl. Mater. Interfaces 2020, 12, 11979–11989. [Google Scholar] [CrossRef]
  51. Liu, Z.; Gui, Y.; Xu, L.; Chen, X. Adsorption and sensing performances of transition metal (Ag, Pd, Pt, Rh, and Ru) modified WSe2 monolayer upon SF6 decomposition gases (SOF2 and SO2F2). Appl. Surf. Sci. 2022, 581, 152365. [Google Scholar] [CrossRef]
  52. Wang, X.; Gui, Y.; Sun, N.; Ding, Z.; Chen, X. A DFT calculation: Gas sensitivity of defect GeSe to air decomposition products (CO, NO and NO2). IEEE Sens. J. 2022. [Google Scholar] [CrossRef]
  53. Pan, Q.; Li, T.; Zhang, D. Ammonia gas sensing properties and density functional theory investigation of coral-like Au-SnSe2 Schottky junction. Sens. Actuator B-Chem. 2021, 332, 129440. [Google Scholar] [CrossRef]
  54. Liu, Y.; Gui, Y.; Xu, L.; Chen, X. Adsorption property of Co, Rh, and Pd-embedded g-C3N4 monolayer to SO2F2 gas. J. Mater. Res. Technol.-JMRT 2021, 15, 4790–4799. [Google Scholar] [CrossRef]
  55. Zhang, D.; Wu, J.; Li, P.; Cao, Y. Room-temperature SO2 gas-sensing properties based on a metal-doped MoS nanoflower: An experimental and density functional theory investigation. J. Mater. Chem. A 2017, 5, 20666–20677. [Google Scholar] [CrossRef]
  56. Zhang, D.; Li, Q.; Li, P.; Pang, M.; Luo, Y. Fabrication of Pd-decorated MoSe2 nanoflowers and density functional theory simulation toward ammonia sensing. IEEE Electron Device Lett. 2019, 40, 616–619. [Google Scholar] [CrossRef]
  57. Hu, X.; Gui, Y.; Zhu, S.; Chen, X. First-principles study of the adsorption behavior and sensing properties of C2H4 and C2H6 molecules on (CuO/TiO2)n (n = 1–3) cluster modified MoTe2 monolayer. Surf. Interfaces 2022, 31, 102003. [Google Scholar] [CrossRef]
Figure 1. The schematic diagram and geometric structures of (a1,a2) GeSe monolayer, (b1,b2) SnO2–GeSe monolayer, (c1,c2) C2H2 adsorption system, (d1,d2) CH4 adsorption system, (e1,e2) H2 adsorption system, and (a3b3) band energy.
Figure 1. The schematic diagram and geometric structures of (a1,a2) GeSe monolayer, (b1,b2) SnO2–GeSe monolayer, (c1,c2) C2H2 adsorption system, (d1,d2) CH4 adsorption system, (e1,e2) H2 adsorption system, and (a3b3) band energy.
Chemosensors 10 00212 g001
Figure 2. The TDOS of (a) GeSe monolayer and SnO2–GeSe monolayer, (b) C2H2 adsorption system, (c) CH4 adsorption system, (d) H2 adsorption system. The Fermi level is set at zero.
Figure 2. The TDOS of (a) GeSe monolayer and SnO2–GeSe monolayer, (b) C2H2 adsorption system, (c) CH4 adsorption system, (d) H2 adsorption system. The Fermi level is set at zero.
Chemosensors 10 00212 g002
Figure 3. The ELF of (a) C2H2 adsorption system, (b) CH4 adsorption system, and (c) H2 adsorption system.
Figure 3. The ELF of (a) C2H2 adsorption system, (b) CH4 adsorption system, and (c) H2 adsorption system.
Chemosensors 10 00212 g003
Figure 4. The CDD of (a) C2H2 adsorption system, (b) CH4 adsorption system, and (c) H2 adsorption system.
Figure 4. The CDD of (a) C2H2 adsorption system, (b) CH4 adsorption system, and (c) H2 adsorption system.
Chemosensors 10 00212 g004
Figure 5. The recovery time (τ) of the three adsorption systems.
Figure 5. The recovery time (τ) of the three adsorption systems.
Chemosensors 10 00212 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, L.-Y.; Liang, S.; Yang, Z.; Jin, L.; Tan, Y.; Huang, Z. Gas-Sensing Properties of Dissolved Gases in Insulating Material Adsorbed on SnO2–GeSe Monolayer. Chemosensors 2022, 10, 212. https://doi.org/10.3390/chemosensors10060212

AMA Style

Guo L-Y, Liang S, Yang Z, Jin L, Tan Y, Huang Z. Gas-Sensing Properties of Dissolved Gases in Insulating Material Adsorbed on SnO2–GeSe Monolayer. Chemosensors. 2022; 10(6):212. https://doi.org/10.3390/chemosensors10060212

Chicago/Turabian Style

Guo, Liang-Yan, Suning Liang, Zhi Yang, Lingfeng Jin, Yaxiong Tan, and Zhengyong Huang. 2022. "Gas-Sensing Properties of Dissolved Gases in Insulating Material Adsorbed on SnO2–GeSe Monolayer" Chemosensors 10, no. 6: 212. https://doi.org/10.3390/chemosensors10060212

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