Next Article in Journal
Thermal, Physical, and Optical Properties of the Solution and Melt Synthesized Single Crystal CsPbBr3 Halide Perovskite
Next Article in Special Issue
Adsorption and Sensing of CO2, CH4 and N2O Molecules by Ti-Doped HfSe2 Monolayer Based on the First-Principle
Previous Article in Journal
Nanocomposites of Carbon Quantum Dots and Graphene Quantum Dots: Environmental Applications as Sensors
Previous Article in Special Issue
Ni-Decorated ZnO Monolayer for Sensing CO and HCHO in Dry-Type Transformers: A First-Principles Theory
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mo2C-Based Microfluidic Gas Sensor Detects SF6 Decomposition Components: A First-Principles Study

1
School of Electrical and Electronic Engineering, Hubei University of Technology, Wuhan 430068, China
2
Hubei Engineering Research Center for Safety Monitoring of New Energy and Power Grid Equipment, Hubei University of Technology, Wuhan 430068, China
3
School of Electrical Engineering and Automation, Wuhan University, Wuhan 430072, China
*
Author to whom correspondence should be addressed.
Chemosensors 2022, 10(9), 368; https://doi.org/10.3390/chemosensors10090368
Submission received: 11 August 2022 / Revised: 7 September 2022 / Accepted: 14 September 2022 / Published: 16 September 2022
(This article belongs to the Special Issue Application and Advance of Gas Sensors)

Abstract

:
Mo2C is a two-dimensional material with high electrical conductivity, low power consumption, and catalytic effect, which has promising applications in the field of microfluidic gas detection. First principles were used to study the adsorption characteristics of Mo2C monolayer on four typical decomposition gases of SF6 (H2S, SO2, SOF2, and SO2F2), and to explore the feasibility of its application in the detection of SF6 decomposition components. The results showed that Mo2C chemisorbed all four gases, and the adsorption capacity was H2S < SO2 < SOF2 < SO2F2. The adsorption mechanism of Mo2C as a microfluidic sensor was analyzed in combination with its charge-density difference and density of states. On the other hand, the different work-function change trends after adsorbing gases provide the possibility for Mo2C to selectively detect gases as a low-power field-effect transistor sensor. All content can be used as theoretical guidance in the realization of Mo2C as a gas-sensitive material for the detection of SF6 decomposition components.

1. Introduction

SF6, as a gas medium with excellent insulation and arc extinguishing properties, plays an important role in reducing the footprint of high-voltage equipment and optimizing the installation design, etc. [1,2,3]. Because of the excellent chemical properties of SF6 insulating gas, it is more and more widely used in the field of electric power. During its long-term operation, SF6 will be decomposed due to partial discharge and partial overheating, and the generated series of products will react with the moisture and oxygen present in the equipment to produce a stable mixture of gas components, represented by H2S, SO2, SOF2, and SO2F2 [4,5]. The presence of these gases will reduce the insulation performance of SF6, and eventually lead to equipment failure. Studies have shown that the type and content of these gases can be used to determine the type, size, and cause of partial discharges and to provide reliable information for assessing the insulation status of the equipment [6,7]. Therefore, it is necessary to test these gases to ensure the insulation status of electrical equipment.
A microfluidic gas-sensing-and-detection method based on nano gas-sensitive materials has received wide attention for its high sensitivity, high selectivity, high reliability and long-term stability, and it is the future development direction of built-in sensing and online monitoring of gas insulation equipment due to its high integration and small size. With nanogas-sensitive materials as its core working part, many scholars have done a lot of research work from theory to experiment on the gas-sensitive properties of these materials: transition metal sulfides represented by MoS2 [8,9], transition metal carbides represented by GaN, InN [10,11,12], metal oxides represented by SnO2, ZnO [13,14], etc. More and more gas-sensitive materials are being developed and utilized. Mo2C, as one of the transition metal carbide nanomaterials, has been used in various application studies because of its superior physical, chemical, and mechanical properties. Soo-Yeon Cho et al. [15] used molybdenum carbide to fabricate solid-state gas sensors with high sensitivity and stability. Wang Tao et al. [16] theoretically investigated the relationship between the stability of different Mo2C surfaces and CO adsorption. In particular, Mo2C has a high conductivity similar to that of pure metals [17], which offers the possibility to be used as low-power device development.
Two-dimensional Mo2C planes were selected as gas adsorption surfaces for the study, and firstly, various adsorption structures of four typical SF6 decomposition components, SO2, SOF2, SO2F2, and H2S, were constructed on the Mo2C surface, respectively. The optimal adsorption configuration of each adsorption system is derived from the adsorption energy, and then the chemical interaction between the gas molecule and the surface and the electronic structure of the system are investigated for each adsorption structure, and the Charge Density Difference (CDD) and Density of States (DOS) are calculated to analyze the bonding between the atoms and the surface within the adsorbed molecule in detail. In order to further explore and broaden the possibility of the application of Mo2C-based microfluidic devices in practical engineering, the theoretical desorption time of each adsorption system and the possibility of application in field-effect transistor sensors were calculated in this paper. The interaction behavior of the Mo2C surface and SF6 decomposition components is discussed in depth in theoretical calculations, which provides a theoretical basis for developing Mo2C-based microfluidic gas sensors with better adsorption performance for SF6 decomposition component detection [18,19,20,21].

2. Computational Details and Models

All models were constructed under the Materials Studio software and all calculations were performed under the Dmol3 module [22]. The Perdew–Burke–Ernzerhof (PBE) generalized function of the Generalized Gradient Approximation (GGA) method was adopted to handle the exchange correlations of the electron parts [23]. The p-polarized double numerical polarization (DNP) is adopted as the basis set function of the atomic orbital. Since Mo is a transition metal element, the dispersion-corrected density functional theory (DFT-D) method developed by Tkachenko and Scheffler and the TS method are used to understand the weak interaction and van der Waals force, instead of using Grimme’s empirical diffusion-corrected DFT-D [24]. The orbital truncation radius was set to 5.0 Å (1 Å = 1 × 10−10 m) [25]. For the geometry-optimized relaxation of all structures, the convergence criteria were set as follows: the difference in energy between the two geometry optimizations was less than 1.0 × 10−5 Ha (1 Ha = 27.21 eV), the force per atom was less than 0.002 Ha/Å, and the maximum displacement distance per atom was less than 0.005 Å [26]. For the electron step iteration in each geometry optimization step, the energy difference between the two calculations is less than 1.0 × 10−6 Ha.
A 4 × 4 × 1 monolayer of Mo2C contains 16 C atoms and 32 Mo atoms, and a 15 Å vacuum layer is set up to eliminate the effect of interlayer interactions.
The four gas molecules approach the Mo2C surface in different orientations to finally obtain multiple adsorption structures with minimum total local energy. Among them, the adsorption energy (Eads) of a single gas molecule on the Mo2C surface is calculated as [27]:
E ads = E monolayer + g a s E monolayer E gas
where E monolayer + g a s is the total energy of the system after adsorption of gas molecules on the Mo2C surface, E monolayer is the energy of Mo2C, and E gas is the energy of gas molecules; if Eads is negative, it indicates that this is a spontaneous and stable exothermic reaction [28].
Mulliken’s method was used to analyze the charge transfer between the gas molecules and the adsorbed substrate, and the corresponding calculation equation is shown below:
Δ Q = Q 1 Q 2
where Q1 represents the charge of the gas molecule after adsorption, and Q2 represents the charge of the gas molecule before adsorption, and define ΔQ to quantify the charge change; if ΔQ is positive, it indicates that the gas molecule loses electrons and the crystal surface gains electrons during the adsorption process, and the opposite if ΔQ is negative [29].

3. Results and Discussion

A model of the most stable two-dimensional Mo2C is shown in Figure 1a. The structure is derived from the 111 surface of the cubic Mo2C with the original lattice parameter of a = 4.155 Å. Mo2C is a typical three-layer interlayer material with a top view demonstrating a graphene-like six-membered ring structure, a configuration similar to other reported two-dimensional transition metal sulfides typified by MoS2 [30]. The band structure of two-dimensional Mo2C is shown in Figure 1b, and the calculated results show a band gap of 0, confirming its high conductivity with a metallic nature [17]. Figure 1c shows the structures of the four gas molecules, from which it can be seen that the structures of the four gas molecules are basically symmetrical. Combining the setting scheme of the adsorption sites and the structural characteristics of the gas molecules in the relevant adsorption studies, three different adsorption configurations were constructed for H2S, SO2 gas molecules, while four different adsorption configurations were constructed for SOF2 and SO2F2 gases. The data such as adsorption energy for the best adsorption configuration of the four gas molecules are shown in Table 1. The adsorption energy of MoS2 (Pt-doped) [31] for SO2F2, SOF2 is −1.39 eV and −2.65 eV, respectively, while the adsorption energy of GaN (Cr-doped) [4] for H2S and SO2 is −1.02 eV and −2.84 eV, respectively, which is significantly larger than that of Mo2C. In order to better analyze the adsorption behavior of Mo2C on these four gases, the HOMO (Highest Occupied Molecular Orbital)-LUMO (Lowest Unoccupied Molecular Orbital) energies of these four gases were calculated, and the specific values are shown in Table 2. SO2F2 has the largest energy gap, followed by SOF2, SO2 and H2S, which indicates that the SO2F2 molecule is the most stable. In terms of HOMO energies, SO2F2 is the largest, which means that it is the least bound, so it is the most active and the most variable, which is verified by the adsorption energy. In the next section, the adsorption behavior of the four gas molecules will be discussed in depth in terms of the band structure, DOS, and charge transfer.

3.1. Adsorption Characteristic of Mo2C

Based on the obtained most stable Mo2C configuration, for the H2S adsorption system, three different adsorption models: H2S molecules parallel to the Mo2C surface, two H atoms facing downward perpendicular to the Mo2C surface, and S atoms facing downward perpendicular to the Mo2C surface were constructed respectively to perform adsorption calculations, and the most stable configuration of H2S gas molecules adsorbed on the Mo2C surface was obtained according to the adsorption energy calculation formula (Equation (1)), as shown in Figure 2a. It is obvious from the figure that H2S interacts with Mo2C in a form slightly parallel to the Mo2C surface as a whole, with an adsorption energy of −1.560 eV, which is less than −0.800 eV, and it is easy to conclude that the adsorption of H2S is chemisorption [32]. Further analysis shows that the adsorption distance of H2S is 2.543 Å and there is no significant deformation of the H2S molecular structure after adsorption, and from the Mulliken aspect, Mo2C gets 0.125 e after the adsorption is completed. As seen in the corresponding CDD of Figure 3c, there is a more obvious overlap of electron dissipation and accumulation between the downward-facing F atom and the two O atoms and the Mo atom below it, explaining the interaction between them [33]. Similar to the H2S adsorption system, three different adsorption models: SO2 molecules parallel to the Mo2C surface, two O atoms facing downward perpendicular to the Mo2C surface, and S atoms facing downward perpendicular to the Mo2C surface were constructed to perform adsorption calculations. The most stable configuration of SO2 gas molecules adsorbed on the Mo2C surface was obtained, as shown in Figure 2b. It is obvious from the figure that SO2 is adsorbed on the Mo2C surface in the form of two O atoms facing downward and S atoms facing upward, with an adsorption energy of −2.538 eV, it is easier to conclude that the adsorption of SO2 is chemisorption. Upon further analysis, the adsorption distance of SO2 is 2.152 Å, and there is no significant deformation of the SO2 molecular structure after adsorption, and from the Mulliken aspect, Mo2C loses 0.224 e after the adsorption is completed. The electron transfer process of SO2 adsorption on the Mo2C surface can be directly seen in the corresponding CDD of Figure 3b. A large amount of blue charge accumulation is wrapped around the whole SO2 molecule, while the Mo atom directly below is surrounded by a yellow charge dissipation region, verifying the electron gaining behavior of SO2 during the adsorption. The interaction between the O atom and the Mo atom below it is explained by the presence of a certain overlap of electron dissipation and accumulation.
For the SOF2 adsorption system, four different adsorption models were constructed with S atoms facing the Mo2C surface and other atoms facing up, S atoms facing up and other atoms facing the Mo2C surface, O atoms facing the Mo2C surface and other atoms facing up and F atoms facing the Mo2C surface and other atoms facing up, respectively. The most stable configuration of SOF2 gas molecules adsorbed on the Mo2C surface was obtained, as shown in Figure 2c. It is obvious from the figure that SOF2 is adsorbed on the Mo2C surface in the form of two F atoms and one O atom facing downward and sulfur atoms facing upward, with an adsorption energy of −3.458 eV, which is a more obvious chemisorption. Further analysis shows that the adsorption distance of SOF2 is 1.888 Å. After adsorption, the two F atoms of the SOF2 molecule are obviously far away from the molecular body and finally located directly above the Mo atom, respectively. Mo2C loses 0.634 e after the adsorption is completed. As seen in the corresponding CDD of Figure 3c, there is a more pronounced overlap of electron dissipation and accumulation between the downward-facing O atom and the two F atoms and the Mo atom below them, explaining the stronger interaction between them. For the SO2F2 adsorption system, four different adsorption models were constructed with S atoms facing the Mo2C surface and other atoms facing upward, F atoms facing the Mo2C surface and other atoms facing upward, two O atoms facing the Mo2C surface and other atoms facing upward and two F atoms facing the Mo2C surface and other atoms facing upward. The most stable configuration of SO2F2 gas molecules adsorbed on the Mo2C surface was obtained as shown in Figure 2d. It is obvious from the figure that SO2F2 is adsorbed on the Mo2C surface in the form of one F-atom facing upward, two O atoms facing slightly downward, and another F atom facing downward, with an adsorption energy of −4.944 eV, which is a more obvious chemisorption as well. Further analysis shows that the adsorption distance of SO2F2 is 1.511 Å. After adsorption, one F atom of the SO2F2 molecule is far away from the molecular body and finally located directly above the C atom. From the Mulliken aspect, Mo2C loses 0.488 e after the adsorption is completed. From the corresponding CDD in Figure 3d, it can be intuitively seen that the electron dissipation and accumulation with large overlap between SO2F2 and its underlying Mo atoms explain the strong interaction between them.

3.2. Electronic Behavior

The band structure of the most stable adsorption configuration of H2S is shown in Figure 4a. Compared with the band structure of the Mo2C substrate before adsorption, the band gap value remains zero, but the band near the Fermi level increases. To further analyze the adsorption mechanism, the DOS is plotted as shown in Figure 5(a1,a2). There is no change in the total DOS at the Fermi level after the adsorption of H2S by Mo2C, while the partial overlap of the S 3p orbitals and their nearest Mo 4d orbitals can be seen from the partial density of states map, but the overlap region is small due to the small percentage of the Mo 4d orbital region. This means that there is a certain orbital hybridization between H2S and Mo2C, which means that there is some interaction between them [34]. The band structure of the most stable adsorption configuration of SO2 is shown in Figure 4c, compared with the band structure of the Mo2C substrate before adsorption, the band gap value remains 0 without change, but the energy band near the Fermi level increases further compared to the H2S adsorption system. The DOS is plotted for further analysis of the adsorption mechanism, as shown in Figure 5(b1,b2), the trend of the total density of states changes is unchanged after the adsorption of SO2 by Mo2C, but the peak at −0.4 eV increases compared to that before the adsorption, and the change of this peak is known from the corresponding partial density of states plots contributed by the Mo 4d orbital. Similarly, the peak around −4.4 eV is contributed by the O 2p orbital. In terms of the degree of the partial density of states crossover, the O 2p orbitals and Mo 4p orbitals are more crossed at −3 eV to −5 eV and less crossed at −2 eV−1 eV. Overall, the orbital hybridization existing between SO2 and Mo2C is stronger than that of the H2S adsorption system, which is also verified in the adsorption energy.
The band structure of the most stable adsorption configuration of SOF2 is shown in Figure 4c, and its band gap value is the same as that of the Mo2C substrate before adsorption, but the energy band near the Fermi level is further increased compared with that of the SO2 adsorption system. The peak at −0.4 eV is lower than that before adsorption, but the DOS in the region to the left of −3.5 eV is higher than that before adsorption. Overall, the orbital hybridization existing between SOF2 and Mo2C is stronger than that of the SO2 adsorption system. Figure 4d shows the band structure of the most stable adsorption configuration of SO2F2, where the band gap value is the same as that of the Mo2C substrate before adsorption, but the energy band near the Fermi level is further increased compared to the other three adsorption systems. The corresponding partial density of states plots shows that the F 2p orbitals, Mo 4d orbitals, and C 2p orbitals in the corresponding range have different degrees of crossover, and the degree of crossover at the Fermi level is relatively small. In general, the orbital hybridization between SO2F2 and Mo2C is the strongest among the four adsorption systems, which is also verified in the adsorption energy.
Through the analysis of the above four adsorption systems, it was found that Mo2C is in contact with the four SF6 decomposition component gas, the adsorption energy is larger, more easily captured by the Mo2C surface, and its surface electron transfer, resulting in an increasing in the energy band, but the band structure shows that it still shows metallic, which increases the possibility of Mo2C as low-power gas-sensitive sensing materials for industrial applications. The next section will explore other applications in terms of low-power material properties. The feasibility of industrial applications of Mo2C will be further explored in the next section in terms of the theoretical sensing response and recovery time of gas-sensitive sensors.

3.3. Resistance-Type Sensor Exploration

In theoretical analysis, calculation formulas are usually used to directly reflect the sensitivity of gas-sensing materials to gases, but for Mo2C, which has no band gap, it cannot be calculated and directly reflected by this method. The good adsorption characteristics of Mo2C for all four gas molecules are indirectly known through the detailed analysis in the previous section, and this section intuitively reflects it from the theoretical recovery time.
The recovery time is the process of gas detachment from the sensor surface and its duration can be used to determine whether the material can be used as a basis for multiple gas detection over a long period of time. The recovery time (τ) can be calculated from the Van’t Hoff Arrhenius formula [17].
τ = A 1 e ( E a k B T )
where A is the attempt frequency factor, whose magnitude is defined as 1012 (s−1) [35], Ea (eV) is the potential barrier to be overcome for desorption, which here is equivalent to the opposite of the adsorption energy. kB is the Boltzmann constant, and T is the operating temperature. Equation (3) shows that the larger the potential barrier to be overcome for desorption at the same temperature, the longer the desorption time; in other words, the smaller the adsorption energy and the more difficult it is to desorb. In addition, for the same material, the time for the gas to come off the surface of the gas-sensitive material can be shortened by increasing the working temperature. The desorption times of these four gases at 298 K, 398 K, 498 K and 598 K were calculated by Equation (3). For H2S gas, when the temperature rises to 598 K, the desorption time is an ideal 13.9 s. For the other three gas molecules with smaller adsorption energy, the sensor recovery characteristics are not ideal when there is no external interference; it is necessary to increase the external conditions to enhance the recovery performance, and increase the life of the sensor. For example, when the sensor ends a service cycle, by adding high-power ultraviolet irradiation or high heat in an oxygen-rich environment, the desorption of the adsorbed gas is accelerated, increasing the repeatability of the sensor. In general, Mo2C gas-sensitive materials are not suitable for resistive gas sensors. The next section will explore other applications in terms of low-power material properties.

3.4. Work Function (WF) Analysis

WF represents the minimum energy required for the transfer of electrons from the surface of the adsorbed substrate material to the vacuum [36]. This minimum energy determines the contact potential of the band arrangement between the gas molecules and the gas-sensitive material during the whole gas adsorption process [37]. This section will discuss the possibility of Mo2C as a field-effect transistor sensor.
Figure 6 shows the WF of Mo2C and the four adsorption systems, from which it can be seen that the trend of the work function changes differently for the adsorption of different gases [38]. In more detail, the WF of Mo2C decreases to 4.56 eV after the adsorption of H2S, and the WF of the system decreases to 4.71 after the adsorption of SO2, indicating that it is easier for electrons to escape from the Mo2C surface to the vacuum layer after the adsorption of these two gases [39]. In contrast, the WF of the system increases to 4.96 eV and 5.04 eV after the adsorption of SOF2 and SO2F2, respectively, which increases the difficulty of electron escape from the Mo2C surface. The change in WF caused by gas adsorption shows that Mo2C has the possibility of gas detection using Kelvin probe microscopy [40]. Since Mo2C has two different WF trends for the four gases, this provides the possibility to explore Mo2C as a field-effect transistor sensor for selective detection of SF6 decomposition gases [41,42].

4. Conclusions

First principles were employed to investigate the adsorption behavior of Mo2C on H2S, SO2, SOF2, and SO2F2 for sensing applications in the detection of SF6 decomposition products. The whole research work is divided into three parts: firstly, the structure and properties of Mo2C material are introduced; secondly, the adsorption behavior of four gas molecules on the Mo2C surface is investigated; and finally, the sensing performance of Mo2C for these four gases is comprehensively analyzed considering industrial applications. The results showed that H2S, SO2, SOF2, and SO2F2 were successfully adsorbed on the Mo2C surface with adsorption energies of −1.560 eV, −2.538 eV, −3.548 eV, and −4.944 eV, respectively, and their adsorption properties were determined by combining CDD and DOS analyses. The possibility of Mo2C as a field-effect transistor sensor with good selectivity was found from WF calculations. Overall, the whole research work provides data to support the industrial application of Mo2C as a microfluidic gas-sensitive sensor.

Author Contributions

L.L. carried out the calculations and wrote this manuscript, G.Z. conceived and designed the research, Z.W., J.Y. and S.T. helped analyze the data, Y.L. review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Key Research and Development Pro gram of Hubei Province, China (No. 2020BAA02), Natural Science Foundation of Hubei Province (No. 2020CFB166).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tang, J.; Rao, X.; Zeng, F.; Cai, W.; Cheng, L.; Zhang, C. Influence Mechanisms of Trace H2O on the Generating Process of SF6 Spark Discharge Decomposition Components. Plasma Chem. Plasma Processing 2017, 37, 325–340. [Google Scholar] [CrossRef]
  2. Wu, Y.; Ding, D.; Wang, Y.; Zhou, C.; Lu, H.; Zhang, X. Defect recognition and condition assessment of epoxy insulators in gas insulated switchgear based on multi-information fusion. Measurement 2022, 190, 110701. [Google Scholar] [CrossRef]
  3. Wang, Y.; Ding, D.; Zhang, Y.; Yuan, Z.; Tian, S.; Zhang, X. Research on infrared spectrum characteristics and detection technology of environmental-friendly insulating medium C5F10O. Vib. Spectrosc. 2022, 118, 103336. [Google Scholar] [CrossRef]
  4. Zhang, X.; Liu, L.; Wang, J.; Wang, Z. Detection of SF6 decomposition components by pristine and Cr-doped GaN based on the first-principles theory. Comput. Theor. Chem. 2021, 1205, 113431. [Google Scholar] [CrossRef]
  5. Wang, J.; Zhang, X.; Liu, L.; Wang, Z. Adsorption of SF6 Decomposition Products by the S Vacancy Structure and Edge Structure of SnS2: A Density Functional Theory Study. ACS Omega 2021, 6, 28131–28139. [Google Scholar] [CrossRef] [PubMed]
  6. Tang, J.; Liu, F.; Zhang, X.; Meng, Q.; Zhou, J. Partial discharge recognition through an analysis of SF6 decomposition products part 1: Decomposition characteristics of SF6 under four different partial discharges. IEEE Trans. Dielectr. Electr. Insul. 2012, 19, 29–36. [Google Scholar] [CrossRef]
  7. Tang, J.; Liu, F.; Meng, Q.; Zhang, X.; Tao, J. Partial discharge recognition through an analysis of SF6 decomposition products part 2: Feature extraction and decision tree-based pattern recognition. IEEE Trans. Dielectr. Electr. Insul. 2012, 19, 37–44. [Google Scholar] [CrossRef]
  8. Li, Y.; Zhang, X.; Chen, D.; Xiao, S.; Tang, J. Adsorption behavior of COF2 and CF4 gas on the MoS2 monolayer doped with Ni: A first-principles study. Appl. Surf. Sci. 2018, 443, 274–279. [Google Scholar] [CrossRef]
  9. Ding, K.; Lin, Y.; Huang, M. The enhancement of NO detection by doping strategies on monolayer MoS2. Vac. Technol. Appl. Ion Phys. Int. J. Abstr. Serv. Vac. Sci. Technol. 2016, 130, 146–153. [Google Scholar]
  10. Wang, J.; Zhang, X.; Liu, L.; Wang, Z. Dissolved gas analysis in transformer oil using Ni-Doped GaN monolayer: A DFT study. Superlattices Microstruct. 2021, 159, 107055. [Google Scholar] [CrossRef]
  11. Kadioglu, Y.; Ersan, F.; Kecik, D.; Aktürk, O.; Aktürk, E.; Ciraci, S. Chemical and substitutional doping, and anti-site and vacancy formation in monolayer AlN and GaN. Phys. Chem. Chem. Phys. 2018, 20, 16077–16091. [Google Scholar] [CrossRef] [PubMed]
  12. Feng, C.; Qin, H.; Yang, D.; Zhang, G. First-Principles Investigation of the Adsorption Behaviors of CH2O on BN, AlN, GaN, InN, BP, and P Monolayers. Materials 2019, 12, 676. [Google Scholar] [CrossRef] [PubMed]
  13. Meng, R.; Lu, X.; Ingebrandt, S.; Chen, X. Adsorption of Gas Molecules on Graphene-Like ZnO Nanosheets: The Roles of Gas Concentration, Layer Number, and Heterolayer. Adv. Mater. Interfaces 2017, 4, 1700647. [Google Scholar] [CrossRef]
  14. Gönüllü, Y.; Haidry, A.A.; Saruhan-Brings, B.J.S. Nanotubular Cr-doped TiO2 for use as high-temperature NO2-selective gas sensor. Sens. Actuators B Chem. 2015, 217, 78–87. [Google Scholar] [CrossRef]
  15. Cho, S.-Y.; Kim, J.Y.; Kwon, O.; Kim, J.; Jung, H.-T. Molybdenum carbide chemical sensors with ultrahigh signal-to-noise ratios and ambient stability. J. Mater. Chem. A 2018, 6, 23408–23416. [Google Scholar] [CrossRef]
  16. Wang, T.; Tian, X.; Yang, Y.; Li, Y.-W.; Wang, J.; Beller, M.; Jiao, H. Coverage dependent adsorption and co-adsorption of CO and H2 on the CdI2-antitype metallic Mo2C(001) surface. Phys. Chem. Chem. Phys. 2014, 17, 1907–1917. [Google Scholar] [CrossRef]
  17. Hassan, A.; Ilyas, S.Z.; Ahmed, S.; Niaz, F.; Jalil, A.; Khan, Q. Ab-initio study of molybdenum carbide (Mo2C) as an adsorption-based filter. Phys. Lett. A 2021, 392, 127119. [Google Scholar] [CrossRef]
  18. Zhang, D.; Tong, J.; Xia, B. Humidity-sensing properties of chemically reduced graphene oxide/polymer nanocomposite film sensor based on layer-by-layer nano self-assembly. Sens. Actuators B Chem. 2014, 197, 66–72. [Google Scholar] [CrossRef]
  19. Zhang, D.; Sun, Y.E.; Li, P.; Zhang, Y. Facile fabrication of MoS2-modified SnO2 hybrid nanocomposite for ultrasensitive humidity sensing. ACS Appl. Mater. Interfaces 2016, 8, 14142–14149. [Google Scholar] [CrossRef]
  20. Zhang, X.; Yu, L.; Wu, X.; Hu, W. Experimental Sensing and Density Functional Theory Study of H2S and SOF2Adsorption on Au-Modified Graphene. Adv. Sci. 2015, 2, 1500101. [Google Scholar] [CrossRef]
  21. Yang, Q.-L.; Ban, Y.-L.; Lian, J.-W.; Yu, Z.-F.; Wu, B. SIW Butler Matrix with Modified Hybrid Coupler for Slot Antenna Array. IEEE Access 2016, 4, 9561–9569. [Google Scholar] [CrossRef]
  22. Cui, H.; Zheng, K.; Tao, L.; Yu, J.; Zhu, X.; Li, X.; Chen, X. Monolayer Tellurene-Based Gas Sensor to Detect SF6 Decompositions: A First-Principles Study. IEEE Electron Device Lett. 2019, 40, 1522–1525. [Google Scholar] [CrossRef]
  23. Wang, Y.; Zhang, D.; Han, J.; Yang, Y.; Guo, Y.; Bai, Z.; Cheng, J.; Chu, P.K.; Pang, H.; Luo, Y. Porous Mo2C-Mo3N2 heterostructure/rGO with synergistic functions as polysulfides regulator for high-performance lithium sulfur batteries. Chem. Eng. J. 2021, 433, 133629. [Google Scholar] [CrossRef]
  24. Zhang, D.; Yu, S.; Wang, X.; Huang, J.; Pan, W.; Zhang, J.; Meteku, B.E.; Zeng, J. UV illumination-enhanced ultrasensitive ammonia gas sensor based on (001)TiO2/MXene heterostructure for food spoilage detection. J. Hazard. Mater. 2022, 423, 127160. [Google Scholar] [CrossRef]
  25. Zhang, Y.-H.; Han, L.-F.; Xiao, Y.-H.; Jia, D.-Z.; Guo, Z.-H.; Li, F. Understanding dopant and defect effect on H2S sensing performances of graphene: A first-principles study. Comput. Mater. Sci. 2013, 69, 222–228. [Google Scholar] [CrossRef]
  26. Liu, Y.; Zhou, Q.; Mi, H.; Wang, J.; Zeng, W. Gas-sensing mechanism of Cr doped SnP3 monolayer to SF6 partial discharge decomposition components. Appl. Surf. Sci. 2021, 546, 149084. [Google Scholar] [CrossRef]
  27. Zhang, X.; Wang, J.; Chen, D.; Liu, L. The adsorption performance of harmful gas on Cu doped WS2: A first-principle study. Mater. Today Commun. 2021, 28, 102488. [Google Scholar] [CrossRef]
  28. Sun, X.; Yang, Q.; Meng, R.; Tan, C.; Liang, Q.; Jiang, J.; Ye, H.; Chen, X. Adsorption of gas molecules on graphene-like InN monolayer: A first-principle study. Appl. Surf. Sci. 2017, 404, 291–299. [Google Scholar] [CrossRef]
  29. Ma, S.; Li, D.; Rao, X.; Xia, X.; Su, Y.; Lu, Y. Pd-doped h-BN monolayer: A promising gas scavenger for SF6 insulation devices. Adsorption 2020, 26, 619–626. [Google Scholar] [CrossRef]
  30. Chen, D.; Zhang, X.; Tang, J.; Cui, H.; Li, Y. Noble metal (Pt or Au)-doped monolayer MoS2 as a promising adsorbent and gas-sensing material to SO2, SOF2 and SO2F2: A DFT study. Appl. Phys. A 2018, 124, 194. [Google Scholar] [CrossRef]
  31. Gui, Y.; Shi, J.; Yang, P.; Li, T.; Tang, C.; Xu, L. Platinum modified MoS2 monolayer for adsorption and gas sensing of SF6 decomposition products: A DFT study. High Volt. 2020, 5, 454–462. [Google Scholar] [CrossRef]
  32. Cui, H.; Chen, D.; Zhang, Y.; Zhang, X. Dissolved gas analysis in transformer oil using Pd catalyst decorated MoSe2 monolayer: A first-principles theory. Sustain. Mater. Technol. 2019, 20, e00094. [Google Scholar] [CrossRef]
  33. Liu, X.; Ni, Y.X.; Wang, H.Y.; Wang, H. Tuning structural, electronic, and magnetic properties of black-AsP monolayer by adatom adsorptions: A first principles study. Chin. J. Chem. Phys. 2020, 33, 311–318. [Google Scholar] [CrossRef]
  34. Cui, H.; Zhang, X.; Chen, D.; Tang, J. Pt & Pd decorated CNT as a workable media for SOF2 sensing: A DFT study. Appl. Surf. Sci. 2018, 471, 335–341. [Google Scholar] [CrossRef]
  35. 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. 2021, 581, 152365. [Google Scholar] [CrossRef]
  36. Zhu, H.; Cui, H.; He, D.; Cui, Z.; Wang, X. Rh-doped MoTe2 Monolayer as a Promising Candidate for Sensing and Scavenging SF6 Decomposed Species: A DFT Study. Nanoscale Res. Lett. 2020, 15, 129. [Google Scholar] [CrossRef]
  37. Sun, H.; Tao, L.-Q.; Li, T.; Gao, X.; Wang, G.; Peng, Z.; Zhu, C.; Zou, S.; Gui, Y.; Xia, S.-Y.; et al. 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]
  38. Guo, H.; Zhang, F.; Qiu, J.; Wu, L.; Zhu, B.; Chen, X.; Yu, J. PbSnS₂-Based Gas Sensor to Detect SF₆ Decompositions: DFT and NEGF Calculations. IEEE Trans. Electron Devices 2021, 68, 5322–5325. [Google Scholar] [CrossRef]
  39. Cui, H.; Zhang, G.; Zhang, X.; Tang, J. Rh-doped MoSe2 as a toxic gas scavenger: A first-principles study. Nanoscale Adv. 2018, 1, 772–780. [Google Scholar] [CrossRef]
  40. Liu, W.; Qiu, X.; Song, Y.; Zhang, X.; Tian, S.; Liu, L. Adsorption behaviour of CF4 and COF2 gas on the GaN monolayer doped with Pt catalytic: A first-principles study. Surf. Sci. 2022, 719, 122032. [Google Scholar] [CrossRef]
  41. Zhang, G.Z.; Wang, Z.T.; Zhang, X.X. Theoretical screening into Ru-doped MoS2 monolayer as a promising gas sensor upon SO2 and SOF2 in SF6 insulation devices. Mol. Phys. 2022, 120, 8. [Google Scholar] [CrossRef]
  42. Kou, L.; Tang, C.; Zhang, Y.; Heine, T.; Chen, C.; Frauenheim, T. Tuning Magnetism and Electronic Phase Transitions by Strain and Electric Field in Zigzag MoS2 Nanoribbons. J. Phys. Chem. Lett. 2012, 3, 2934–2941. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) most stable conformation, (b) band structure of Mo2C and (c) four molecular models.
Figure 1. (a) most stable conformation, (b) band structure of Mo2C and (c) four molecular models.
Chemosensors 10 00368 g001
Figure 2. Most stable conformation of H2S (a), SO2 (b), SOF2 (c) and SO2F2 (d) systems.
Figure 2. Most stable conformation of H2S (a), SO2 (b), SOF2 (c) and SO2F2 (d) systems.
Chemosensors 10 00368 g002
Figure 3. CDD of H2S (a), SO2 (b), SOF2 (c) and SO2F2 (d) systems. In CDD, the blue area is electron accumulation and the yellow area is electron depletion, with an isosurface of 0.05 e/Å3.
Figure 3. CDD of H2S (a), SO2 (b), SOF2 (c) and SO2F2 (d) systems. In CDD, the blue area is electron accumulation and the yellow area is electron depletion, with an isosurface of 0.05 e/Å3.
Chemosensors 10 00368 g003
Figure 4. Band structures of H2S (a), SO2 (b), SOF2 (c) and SO2F2 (d) systems. The dash line is the Fermi level.
Figure 4. Band structures of H2S (a), SO2 (b), SOF2 (c) and SO2F2 (d) systems. The dash line is the Fermi level.
Chemosensors 10 00368 g004
Figure 5. DOS of H2S (a1,a2), SO2 (b1,b2), SOF2 (c1,c2) and SO2F2 (d1,d2) systems. The dash line is the Fermi level.
Figure 5. DOS of H2S (a1,a2), SO2 (b1,b2), SOF2 (c1,c2) and SO2F2 (d1,d2) systems. The dash line is the Fermi level.
Chemosensors 10 00368 g005
Figure 6. WF of adsorption systems.
Figure 6. WF of adsorption systems.
Chemosensors 10 00368 g006
Table 1. Adsorption energy (Ead), adsorption distance (D), and transferred charge (Q) for the best adsorption configuration of four gas molecules on Mo2C. D (Å) is the distance between the gas molecule and the nearest atom of the adsorption substrate; Q (e) is the charge transfer between the gas molecule and the adsorption substrate.
Table 1. Adsorption energy (Ead), adsorption distance (D), and transferred charge (Q) for the best adsorption configuration of four gas molecules on Mo2C. D (Å) is the distance between the gas molecule and the nearest atom of the adsorption substrate; Q (e) is the charge transfer between the gas molecule and the adsorption substrate.
MoleculeEad (eV)D (Å)Q (e)
H 2 S −1.5602.5430.125
SO 2 −2.5382.152−0.224
SOF 2 −3.4581.888−0.643
SO 2 F 2 −4.9441.511−0.488
Table 2. HOMO/LUMO energies of the four gas molecules.
Table 2. HOMO/LUMO energies of the four gas molecules.
MoleculeHOMO (eV)LUMO (eV)Energy Gap
H 2 S −0.226−0.0690.157
SO 2 −0.301−0.1800.121
SOF 2 −0.319−0.1210.198
SO 2 F 2 −0.342−0.1080.234
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, L.; Zhang, G.; Wang, Z.; Yuan, J.; Tan, S.; Li, Y. Mo2C-Based Microfluidic Gas Sensor Detects SF6 Decomposition Components: A First-Principles Study. Chemosensors 2022, 10, 368. https://doi.org/10.3390/chemosensors10090368

AMA Style

Liu L, Zhang G, Wang Z, Yuan J, Tan S, Li Y. Mo2C-Based Microfluidic Gas Sensor Detects SF6 Decomposition Components: A First-Principles Study. Chemosensors. 2022; 10(9):368. https://doi.org/10.3390/chemosensors10090368

Chicago/Turabian Style

Liu, Li, Guozhi Zhang, Zengting Wang, Jiawei Yuan, Senyuan Tan, and Yi Li. 2022. "Mo2C-Based Microfluidic Gas Sensor Detects SF6 Decomposition Components: A First-Principles Study" Chemosensors 10, no. 9: 368. https://doi.org/10.3390/chemosensors10090368

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