Skip to main content
ACS Omega logoLink to ACS Omega
. 2020 Mar 27;5(13):7722–7728. doi: 10.1021/acsomega.0c00922

Adsorption and Gas Sensing Properties of the Pt3-MoSe2 Monolayer to SOF2 and SO2F2

Hai Qian 1, Jun Deng 1,*, Zhicheng Xie 1, Zhicheng Pan 1, Jinyin Zhang 1, Haibin Zhou 1
PMCID: PMC7144166  PMID: 32280916

Abstract

graphic file with name ao0c00922_0005.jpg

SF6 acts as an insulation gas in gas-insulated switchgear (GIS), which inevitably decomposes under partial discharge caused by insulation defects. This work is devoted to finding a new gas-sensing material for detecting two characteristic SF6 decomposition products: SOF2 and SO2F2. The platinum-cluster-modified molybdenum diselenide (Pt3-MoSe2) monolayer has been proposed as a gas sensing material. Based on first-principles calculations, the adsorption properties and the mechanism were studied by analyzing the adsorption structures, adsorption energy, charge transfer, density of states, and molecular orbitals. The adsorption ability of Pt3-MoSe2 to SO2F2 is stronger than that to SOF2 due to its chemisorption property. The obvious change of conductivity of the adsorption system during the gas adsorption process shows that Pt3-MoSe2 is sensitive to both of the gas molecules. In addition, the modest adsorption energy signifies that the gas adsorption process can be reversible, which confirms the feasibility of Pt3-MoSe2-based gas sensors. Our calculation suggests that Pt3-MoSe2-based gas sensors can be employed in GIS for partial discharge detection.

1. Introduction

Gas-insulated switchgear (GIS) plays a key role in modern electric systems, which is mostly filled with SF6 gas, a kind of colorless and nontoxic inert gas with high chemical stability and excellent arc extinguishing performance.13 However, electric discharge inevitably occurs in GIS during the long-term running process. Among them, partial discharge is one of the most serious electric discharges because of its strong concealment property.4,5 Under the high energy density of partial discharge, SF6 breaks to various unstable fluorides.6,7 Then, these fluorides quickly react with micro-water and micro-oxygen that existed in GIS to some characteristic decomposition products, such as SOF2 and SO2F2. On the one side, the insulation strength of these decomposition products is obviously weaker than that of SF6. On the other side, these decomposition products can corrode the GIS under the micro-water condition, which decreases the stability of GIS. Partial discharge may ultimately lead to the breakdown of GIS. Currently, online detection of the partial discharge in GIS has attracted great concerns.8,9

Based on the high detection sensitivity, small volume, and low price of the gas sensor, it has been widely used in industry, the environment, and medical fields.10,11 By detecting the SOF2 and SO2F2 products based on appropriate gas sensors, it may be an effective method to online-detect the partial discharge. However, there is still no ideal gas sensitive material for SOF2 and SO2F2 detection at the current stage. Therefore, it is of great significance to explore a new type of gas sensing material for the detection of SF6 decomposition products.

As a transition metal semiconductor material, molybdenum diselenide (MoSe2) has been widely used in various application fields.1214 It is a layered structure with inactive chemical property, which shows high-temperature resistance and strong acid and alkali resistance.15,16 Accompanying its large specific surface area, it has been applied to gas adsorption and sensing studies. Thus, MoSe2 was proposed as a gas sensing material for the detection of SF6 decomposition products. However, intrinsic MoSe2 shows limitations on SOF2 and SO2F2 adsorption.17 Studies showed that transition metal cluster modification could effectively improve the adsorption effect of MoSe2 by enhancing its surface activity. Platinum is one of the most commonly used transition metal for surface modification, which shows excellent surficial activity for gas adsorption.18,19 Thus, platinum-cluster-modified MoSe2 may be a candidate gas sensing material employed in SOF2 and SO2F2 detection.

Sensor performance could change dramatically according to a different number of modification platinum atoms. Studies demonstrated that Pt3 modification could significantly improve the adsorption activity of the MoSe2 surface.2,20,21 In addition, Pt3-MoSe2 shows weak adsorption to SF6 according to our calculation, which makes it compatible with the SF6 environment. Therefore, Pt3-MoSe2 was adopted as a gas sensing material for SOF2 and SO2F2 detection. Based on density functional theory (DFT), a detailed study of the entire adsorption process, adsorption energy, charge transfer, and molecular orbital analysis was conducted and analyzed. A comparative analysis of the adsorption mechanism to SOF2 and SO2F2 gases confirms that Pt3-MoSe2 is an ideal material for these two gases. This study not only analyzes the adsorption mechanism of Pt3-MoSe2 to SOF2 and SO2F2 but also provides a theoretical basis to develop Pt3-MoSe2 sensors in experiment.

2. Results and Discussion

2.1. Structures and Electronic Properties of SOF2, SO2F2, and Pt3-MoSe2

The structures of SOF2, SO2F2, and Pt3-MoSe2 monolayer were optimized, respectively. As shown in Figure 1a–e, both SOF2 and SO2F2 belong to the three-dimensional spatial structure. The bond lengths of the S–F and the S–O in the SOF2 molecule are 1.67 and 1.46 Å, respectively. The tetrahedral structure and high electronegativity make the SOF2 molecule extremely stable. SO2F2 shows similar physicochemical properties to SOF2. However, more oxygen atoms in SO2F2 make it slightly more active, and the effect of orbital hybridization reduces the bond length of the S–F bond. Intrinsic and Pt3-modified MoSe2 are presented in Figure 1d,e. The two-dimensional layered structure makes a great contribution to the large specific surface area. After Pt3 modification, it possesses a more brilliant feature, three Pt atoms form a triangular ring that stands parallel to the MoSe2 surface, which increases the metal activity and adsorption performance of MoSe2.

Figure 1.

Figure 1

Structures of (a) SOF2, (b) SO2F2, (c) MoSe2, and (d, e) Pt3-MoSe2. The distance is in angstroms.

Figure 2 shows the band structure and total density of states (TDOS) of MoSe2 and Pt3-MoSe2, which facilitate the analysis of the change of the physical properties. For intrinsic MoSe2, its band gap is 1.75 eV, which is consistent with its semiconductor characteristics. After Pt cluster doping, the band structure changes dramatically, as the band gap reduces to only 0.16 eV as shown in Figure 2b. As a result, the conductivity of the system rises significantly, which is of great significance for gas sensing. As for the TDOS of MoSe2 and Pt3-MoSe2, TDOS distribution tends to move left. There is more electron filling at the Fermi level, indicating that the conductivity of the system enhances. From the analysis of the projected density of states (PDOS), the 5d orbital of the Pt atom contributes the most to the entire electron distribution.

Figure 2.

Figure 2

(a) Band structure of MoSe2, (b) band structure of Pt3-MoSe2, (c) DOS and PDOS of MoSe2 and Pt3-MoSe2.

2.2. Adsorption of the SOF2 Molecule on the Pt3-MoSe2 Monolayer

Figure 3 shows the different approaching methods of the SOF2 molecule to the Pt3-MoSe2 surface after geometry optimization. There are significant differences in these adsorption structures, and the maximum adsorption energy is up to −1.73 eV by S atom approaching, which simultaneously has the shortest adsorption distance due to the stronger oxidizing ability of S compared with O and F elements. Throughout the various adsorption process, the adsorption distances are 2.12, 2.17, and 3.71 Å, respectively. Almost no charge transfer occurs, and the interaction between SOF2 and Pt3-MoSe2 bends the Pt3 plane. The structure of the SOF2 molecule nearly does not change during the interaction. It can be concluded that the M1 system belongs to chemisorption, and M2 and N3 systems belong to physisorption. According to the huge difference in the adsorption energy among different approaching methods (from −0.46 to −1.73 eV), M1 is the most stable adsorption structure, which has been chosen for further analyzing its adsorption and sensing effects (Table 1).

Figure 3.

Figure 3

Adsorption of the SOF2 molecule on Pt3-MoSe2.

Table 1. Eads, QT, and Adsorption Distance (d) of SOF2 on Pt3-MoSe2.

parameter Eads (eV) QT (e) d (Å)
M1 system –1.73 –0.05 2.12
M2 system –0.51 –0.02 2.17
M3 system –0.46 –0.04 3.71

As shown in Figure 4a, the most stable structure (M1) is adopted for the analysis of physical properties, including TDOS, PDOS, and differential charge density. Once the target gas molecule adsorbs on the surface of Pt3-MoSe2, the electronic properties of the entire system could obviously change, thus the corresponding gas sensor can be prepared based on the different change rule of conductivity.

Figure 4.

Figure 4

TDOS, PDOS, and differential charge density of the SOF2 adsorption system.

Figure 4b shows the TDOS changes before and after gas adsorption. TDOS has a minor change after the SOF2 molecule adsorbs on the Pt3-MoSe2 surface, mainly reflecting in less electron filling at the Fermi level. The redistribution of charge decreases the conductivity of the entire system. According to PDOS, the decrease in TDOS at the Fermi level is mainly caused by the 5d orbital of Pt. In the energy range from −6 to 0 eV, the 3p and 3s orbits of S, the 5d orbital of Pt have a certain degree of orbital hybridization, which causes a slight increase of electron filling in the valence band. Two small peaks aroused by the 3p orbital of S occur at −9 and −7.5 eV energy levels. The change of TDOS shows that the gas molecule adsorption has a great influence on the electronic redistribution in the valence band, and the decline in the conductivity of the entire system depends on the gas itself and the interaction with Pt3-MoSe2.

To intuitively display the change of charge redistribution after gas adsorption, differential charge density was studied as shown in Figure 4c, where blue and red areas represent the loss and reception of electrons, respectively. It can be seen that there is no obvious charge transfer in the adsorption process, because there is only a little charge redistribution according to the red- and blue-color-marked areas. This conclusion is also consistent with the amount of QT analysis above. In conclusion, Pt3-MoSe2 shows chemisorption to SOF2 molecules with suitable adsorption energy.

2.3. Adsorption of the SO2F2 Molecule on the Pt3-MoSe2 Monolayer

As shown in Figure 5, three potential adsorption structures of SO2F2 on the Pt3-MoSe2 surface were obtained, labeled N1, N2, and N3. Table 2 shows the corresponding parameters of adsorption structures. It is evident that the interaction between SO2F2 and the Pt3-MoSe2 monolayer is much stronger than SOF2 adsorption, especially when approaching the surface by the F atom in the N3 system. Dissociation occurs during the adsorption process, and the free F atom and the SO2 group bond to the Pt atom with bond lengths of 2.07 and 2.30 Å by S–Pt and F–Pt, respectively. The adsorption energy for SO2F2 adsorption is up to −3.51 eV with a charge transfer of 0.91e, and the reflection belongs to strong chemisorption. When SO2F2 approaches Pt3-MoSe2 by S and O atoms in N1 and N2 systems, the adsorption energy is −1.61 and −0.44 eV, and the amount of charge transfer is significantly lower than that in the N3 system. Especially in the N2 system, all atoms of SO2F2 are far away from the doped Pt cluster and the MoSe2 surface, which leads to the lowest adsorption energy. According to the adsorption energy and adsorption distance, it can be concluded that strong chemisorption occurs between SO2F2 and Pt3-MoSe2.

Figure 5.

Figure 5

Adsorption of the SO2F2 molecule on Pt3-MoSe2.

Table 2. Eads, QT, and Adsorption Distance (d) of SO2F2 on Pt3-MoSe2.

parameter Eads (eV) QT (e) d (Å)
N1 system –1.61 –0.42 S–Pt: 2.07
      O–Pt: 2.30
N2 system –0.44 –0.01 O–Pt: 2.76
N3 system –3.51 –0.91 F–Pt: 1.95
      S–Pt: 2.17

To analyze the adsorption mechanism of SO2F2 on Pt3-MoSe2, the adsorption properties of were further studied as shown in Figure 6. The approaching method of SO2F2 to Pt3-MoSe2 in N1 and N3 systems are considered in this work as shown in Figure 6a. From the distribution of TDOS in Figure 6b, only a small change of TDOS occurs during SO2F2 adsorption when approaching the N1 system. The electron filling reduces at the Fermi level, which increases the number of electron fills in the valence band. In addition, some peaks of PDOS appear near the −9 eV level, which is caused by the 3p orbital of S, the 2p orbital of O, and the 5d orbital of Pt. The PDOS distribution shows obvious orbital hybridization.

Figure 6.

Figure 6

TDOS, PDOS, and differential charge density of the SO2F2 adsorption system.

A huge change in TDOS and PDOS is obtained in the N3 system in Figure 6b. TDOS moves to the right side as a whole with some new peaks appearing at different energy levels. The 5d orbital of Pt and the 2p orbital of F show strong hybridization, which contributes a lot to TDOS distribution. The formation of a Pt–F chemical bond is a good explanation to this huge change. It is certain that there is clear chemisorption between the SO2F2 molecule and the Pt3-MoSe2 surface. At the Fermi level, there is minimal electron filling after SO2F2 adsorption, and the conductivity drops obviously, which shows a significant meaning for gas sensing. From the perspective of PDOS, The 3p orbital peak of S appears at −6 eV energy level, and it causes a little effect on the electrical conductivity of the whole system. The 2p orbital of F and the 5d orbital of Pt have a large overlap between −4 and 0 eV, which greatly changes the electron distribution at the Fermi level.

Similar to the above analysis, to more intuitively display the charge distribution after gas adsorption, the differential charge density was analyzed as shown in Figure 6c. When approaching the N1 system, Pt atoms lose a lot of electrons to the SO2F2 molecule. Therefore, the charge distribution near the S atom becomes dense, corresponding to the previous charge transfer. In the case of the N3 system, the charge transfer is more distinct, especially near the F–Pt bond; a chemical bond plays a key role in this case. In summary, the surface activity of MoSe2 is significantly enhanced after Pt3 modification, and it shows good sensitivity to the SO2F2 gas.

2.4. Molecular Orbital Analysis for Gas Molecule Adsorption on the Pt3-MoSe2 Monolayer

Molecular orbital analysis was done for SOF2 and SO2F2 gas adsorption in M1 and N3 systems. Based on molecular orbital theory, the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) distributions of the adsorption system were calculated as shown in Figure 7. The Eg between the HOMO and the LUMO was calculated to evaluate conductivity changes, as presented in Table 3.

Figure 7.

Figure 7

HOMO and LUMO distribution of Pt3-MoSe2 before and after gas adsorption.

Table 3. Molecular Orbits and Energy Gaps of Individual Molecules Before and after Adsorption.

structure EHOMO (eV) ELUMO (eV) Eg (eV)
Pt3-MoSe2 –3.92 –3.92 0
M1 system –4.17 –4.05 0.12
N3 system –4.53 –4.31 0.22

As the distribution of the HOMO and the LUMO of the Pt3-MoSe2 monolayer before and after gas adsorption presented in Figure 7, the distribution of the HOMO and the LUMO mainly locates around the Pt atom and Mo atoms, which is related to their unique metal activity. After Pt3 modification, both of the HOMO and the LUMO are −3.92 eV, namely zero gap, reflecting that the Pt3-MoSe2 material shows good electrical conductivity. In the case of SOF2 gas adsorption in the M1 system, a part of LUMO transfers to the gas molecule, and the Eg increases to 0.12 eV. Therefore, the conductivity of the system declines to some extent. As for SO2F2 gas adsorption in the N3 system, almost all LUMO on the Se atom transfers to the Pt atom and the SO2F2 molecule. The Eg has reached 0.22 eV after SO2F2 adsorption. Compared with the adsorption of the SOF2 molecule, the bigger Eg leads to a larger decrease of conductivity of the entire system. In summary, after adsorbing SOF2 and SO2F2 gas molecules, the surface conductivity of the Pt3-MoSe2 material decreases in different ranges. As a new type of two-dimensional material, the Pt3-MoSe2 monolayer provides excellent adsorption properties to SOF2 and SO2F2.

3. Conclusions

The platinum-cluster-modified MoSe2 (Pt3-MoSe2) monolayer has been proposed as a gas sensing material to detect SOF2 and SO2F2 based on first-principles calculations. Both of SOF2 and SO2F2 adsorption on the Pt3-MoSe2 molecule belongs to chemisorption. The adsorption energies of their optimized adsorption structures are −1.73 and −3.51 eV, respectively. From the analysis of DOS and molecular orbital theory, Pt cluster modification improves the surface activity and conductivity of MoSe2-based materials. After the adsorption of SF6 decomposition products on Pt3-MoSe2, the conductivity of the system obviously declines to some extent, verifying that Pt3-MoSe2 is sensitive to both of the gases. The simulation results show that Pt3-MoSe2 could be a very suitable material for SOF2 and SO2F2 adsorption and gas sensing, which lays the foundation for designing Pt3-MoSe2 monolayer-based gas sensors. In summary, our calculations suggest that Pt3-MoSe2-based gas sensors can be employed in GIS for partial discharge detection.

4. Computational Methods

The entire computations in this work were conducted in DMol3 based on DFT. A periodic boundary model with a supercell of 4 × 4 × 1 and a vacuum slab of 20 Å was used to avoid the influence of the adjacent supercell.22,23 Generalized gradient approximation (GGA) of the Perdew–Burke–Ernzerhof (PBE) exchange–correlation functional was taken in this work.2427 The Brillouin-zone was carried out by the Monkhorst–Pack scheme with a k-point set to 7 × 7 × 1, which presents a good approximation for Pt3-MoSe2.28,29 The double numerical plus polarization basis set (DNP) and the DFT semicore pseudopotential (DSSP) method have been applied.30,31 The displacement, maximum stress, and energy convergence were set to 5 × 10–3 Ha, 2 × 10–3 Ha/Å, and 1 × 10–5 Ha, respectively. For the stationary electronic structure, a convergence criterion of 1.0 × 10–6 Ha for self-consistent field tolerance was employed.22,32

To analyze the adsorption properties of the Pt3-MoSe2 surface, the geometries of SOF2, SO2F2, and intrinsic MoSe2 surfaces were first optimized. Then, Pt3 cluster modification was considered to obtain improved adsorption performance. Adsorption energy, charge transfer, and molecular orbital analysis were analyzed.

The adsorption energy (Eads) shows the change of total energy during the adsorption process, which can be calculated by eq 1

4. 1

Esurf/gas shows the total energy of the system after gas molecule adsorption, Esurf indicates the total energy of the Pt3-MoSe2 monolayer before adsorption, and Egas represents the total energy of the isolated gas molecules before adsorption. If Eads < 0, the adsorption process is exothermic and occurs spontaneously.

To analyze the charge transfer in the adsorption process, Mulliken population analysis was used to obtain the charge redistribution and calculate the charge transfer amount. As shown in eq 2, the charge transfer (QT) was used to evaluate the amount of charge transfer between the Pt3-MoSe2 monolayer and gas molecules during gas adsorption, where Qa and Qb represent the total net carried charge of gas molecules after and before gas adsorption.

4. 2

As shown in eq 3, the energy gap (Eg) of the molecular orbital between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) was calculated to evaluate the change of conductivity, where EHOMO and ELUMO represent the energy of the HOMO and the LUMO, respectively.

4. 3

Acknowledgments

This study was supported by the National Key Research and Development Program of China (2017YFB0902701).

The authors declare no competing financial interest.

References

  1. Gui Y.; Tang C.; Zhou Q.; Xu L.; Zhao Z.; Zhang X. The sensing mechanism of N-doped SWCNTs toward SF6 decomposition products: A first-principle study. Appl. Surf. Sci. 2018, 440, 846–852. 10.1016/j.apsusc.2018.01.247. [DOI] [Google Scholar]
  2. Wang Y.; Gui Y.; Ji C.; Tang C.; Zhou Q.; Li J.; Zhang X. Adsorption of SF6 decomposition components on Pt3-TiO2(101) surface: A DFT study. Appl. Surf. Sci. 2018, 459, 242–248. 10.1016/j.apsusc.2018.07.219. [DOI] [Google Scholar]
  3. Liu D.; Gui Y.; Ji C.; Tang C.; Zhou Q.; Li J.; Zhang X. Adsorption of SF6 decomposition components over Pd(111): A density functional theory study. Appl. Surf. Sci. 2019, 465, 172–179. 10.1016/j.apsusc.2018.09.147. [DOI] [Google Scholar]
  4. Zhang X.; Gui Y.; Xiao H.; Zhang Y. Analysis of adsorption properties of typical partial discharge gases on Ni-SWCNTs using density functional theory. Appl. Surf. Sci. 2016, 379, 47–54. 10.1016/j.apsusc.2016.04.048. [DOI] [Google Scholar]
  5. Zhang X.; Yu L.; Gui Y.; Hu W. First-principles study of SF6 decomposed gas adsorbed on Au-decorated graphene. Appl. Surf. Sci. 2016, 367, 259–269. 10.1016/j.apsusc.2016.01.168. [DOI] [Google Scholar]
  6. Zhang X.; Yu L.; Wu X.; Hu W. Experimental sensing and density functional theory study of H2S and SOF2 adsorption on Au-modified graphene. Adv. Sci. 2015, 2, 1500101 10.1002/advs.201500101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. 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: Mater. Sci. Process. 2018, 124, 194 10.1007/s00339-018-1629-y. [DOI] [Google Scholar]
  8. He X.; Gui Y.; Xie J.; Liu X.; Wang Q.; Tang C. A DFT study of dissolved gas (C2H2, H2, CH4) detection in oil on CuO-modified BNNT. Appl. Surf. Sci. 2020, 500, 144030 10.1016/j.apsusc.2019.144030. [DOI] [Google Scholar]
  9. Gui Y.; Li W.; He X.; Ding Z.; Tang C.; Xu L. Adsorption properties of pristine and Co-doped TiO2(1 0 1) toward dissolved gas analysis in transformer oil. Appl. Surf. Sci. 2020, 507, 145163 10.1016/j.apsusc.2019.145163. [DOI] [Google Scholar]
  10. Li T.; Wu Y.; Huang J.; Zhang S. Gas sensors based on membrane diffusion for environmental monitoring. Sens. Actuators, B 2017, 243, 566–578. 10.1016/j.snb.2016.12.026. [DOI] [Google Scholar]
  11. Alrammouz R.; Podlecki J.; Abboud P.; Sorli B.; Habchi R. A review on flexible gas sensors: From materials to devices. Sens. Actuators, A 2018, 284, 209–231. 10.1016/j.sna.2018.10.036. [DOI] [Google Scholar]
  12. Sorkin V.; Pan H.; Shi H.; Quek S. Y.; Zhang Y. W. Nanoscale transition metal dichalcogenides: Structures, properties, and applications.. Crit. Rev. Solid State Mater. Sci. 2014, 39, 319–367. 10.1080/10408436.2013.863176. [DOI] [Google Scholar]
  13. Zhang S.; Yan Z.; Li Y.; Chen Z.; Zeng H. Atomically thin arsenene and antimonene: Semimetal-semiconductor and indirect-direct band-gap transitions. Angew. Chem., Int. Ed. 2015, 54, 3112–3115. 10.1002/anie.201411246. [DOI] [PubMed] [Google Scholar]
  14. Tan C.; Lai Z.; Zhang H. Ultrathin two-dimensional multinary layered metal chalcogenide nanomaterials. Adv. Mater. 2017, 29, 1701392 10.1002/adma.201701392. [DOI] [PubMed] [Google Scholar]
  15. Zhao Y.; Wang W.; Li C.; Xu Y.; He L. Tuning the magnetic properties of the monolayer MoSe2 by nonmetal doping: First-principles study. Solid State Commun. 2018, 281, 6–11. 10.1016/j.ssc.2018.06.010. [DOI] [Google Scholar]
  16. Liu G.; Dai S.; Zhu B.; Li P.; Wu Z.; Gu Y. Third-order nonlinear optical properties of MoSe2/graphene composite materials. Opt. Laser Technol. 2019, 120, 105746 10.1016/j.optlastec.2019.105746. [DOI] [Google Scholar]
  17. Qian H.; Deng J.; Zhou H.; Yang X.; Chen W. First-principles study of Pd-MoSe2 as sensing material for characteristic SF6 decomposition components. AIP Adv. 2019, 9, 125013 10.1063/1.5128831. [DOI] [Google Scholar]
  18. Okubo K.; Ohyama J.; Satsuma A. Surface modification of Pt nanoparticles with other metals boosting the alkaline hydrogen oxidation reaction. Chem. Commun. 2019, 55, 3101–3104. 10.1039/C9CC00582J. [DOI] [PubMed] [Google Scholar]
  19. Xing L.-L.; Wang Q.; Xue X.-Y. One-step synthesis of Pt-ZnO nanoflowers and their enhanced photocatalytic activity. Sci. Adv. Mater. 2016, 8, 1275–1279. 10.1166/sam.2016.2683. [DOI] [Google Scholar]
  20. Baruah S. D.; Gour N. K.; Sarma P. J.; Deka R. C. DFT insight on oxygen adsorbed platinum trimer cluster (Pt3) for CO oxidation. Comput. Theor. Chem. 2017, 1114, 1–7. 10.1016/j.comptc.2017.05.023. [DOI] [Google Scholar]
  21. Jacob T.; Goddard W. A. Chemisorption of (CHx and C2Hy) hydrocarbons on Pt(111) clusters and surfaces from DFT studies. J. Phys. Chem. B 2005, 109, 297–311. 10.1021/jp0463868. [DOI] [PubMed] [Google Scholar]
  22. Wang J.; Zhou Q.; Zeng W. Competitive adsorption of SF6 decompositions on Ni-doped ZnO (100) surface: Computational and experimental study. Appl. Surf. Sci. 2019, 479, 185–197. 10.1016/j.apsusc.2019.01.255. [DOI] [Google Scholar]
  23. Zhou Q.; Chen W.; Xu L.; Kumar R.; Gui Y.; Zhao Z.; Tang C.; Zhu S. Highly sensitive carbon monoxide (CO) gas sensors based on Ni and Zn doped SnO2 nanomaterials. Ceram. Int. 2018, 44, 4392–4399. 10.1016/j.ceramint.2017.12.038. [DOI] [Google Scholar]
  24. Perdew J. P.; Burke K.; Ernzerhof M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
  25. Perdew J. P.; Chevary J. A.; Vosko S. H.; Jackson K. A.; Pederson M. R.; Singh D. J.; Fiolhais C. Atoms, molecules, solids, and surfaces - applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687. 10.1103/PhysRevB.46.6671. [DOI] [PubMed] [Google Scholar]
  26. Hammer B.; Hansen L. B.; Norskov J. K. Improved adsorption energetics within density-functional theory using revised Perdew-Burke-Ernzerhof functionals. Phys. Rev. B 1999, 59, 7413–7421. 10.1103/PhysRevB.59.7413. [DOI] [Google Scholar]
  27. Ernzerhof M.; Scuseria G. E. Assessment of the Perdew-Burke-Ernzerhof exchange-correlation functional. J. Chem. Phys. 1999, 110, 5029–5036. 10.1063/1.478401. [DOI] [PubMed] [Google Scholar]
  28. Monkhorst H. J.; Pack J. D. Special points for brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. 10.1103/PhysRevB.13.5188. [DOI] [Google Scholar]
  29. Blöchl P. E.; Jepsen O.; Andersen O. K. Improved tetrahedron method for brillouin-zone integrations. Phys. Rev. B 1994, 49, 16223–16233. 10.1103/PhysRevB.49.16223. [DOI] [PubMed] [Google Scholar]
  30. Inada Y.; Orita H. Efficiency of numerical basis sets for predicting the binding energies of hydrogen bonded complexes: Evidence of small basis set superposition error compared to Gaussian basis sets. J. Comput. Chem. 2008, 29, 225–232. 10.1002/jcc.20782. [DOI] [PubMed] [Google Scholar]
  31. Lin T. T.; Liu X. Y.; He C. A DFT study on poly(lactic acid) polymorphs. Polymer 2010, 51, 2779–2785. 10.1016/j.polymer.2010.03.062. [DOI] [Google Scholar]
  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. Sustainable Mater. Technol. 2019, 20, e00094 10.1016/j.susmat.2019.e00094. [DOI] [Google Scholar]

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES