Next Article in Journal
Investigation of the Effect of the Skull in Transcranial Photoacoustic Imaging: A Preliminary Ex Vivo Study
Next Article in Special Issue
NO2 Sensing with SWCNT Decorated by Nanoparticles in Temperature Pulsed Mode: Modeling and Characterization
Previous Article in Journal
Disturbance Recognition and Collision Detection of Manipulator Based on Momentum Observer
Previous Article in Special Issue
Gas Sensing Properties of Cobalt Titanate with Multiscale Pore Structure: Experiment and Simulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Letter

Adsorption Properties of Pd3-Modified Double-Vacancy Defect Graphene toward SF6 Decomposition Products

School of Electrical Engineering, Xi’an Jiaotong University, Xi’an 710049, China
*
Author to whom correspondence should be addressed.
Sensors 2020, 20(15), 4188; https://doi.org/10.3390/s20154188
Submission received: 30 June 2020 / Revised: 22 July 2020 / Accepted: 22 July 2020 / Published: 28 July 2020
(This article belongs to the Special Issue Gas Sensing Materials)

Abstract

:
In this study, we investigate Pd3-cluster-modified 555–777 graphene (Pd3-graphene) as a novel resistor-type gas sensor to detect SF6 decomposition products based on density functional theory calculations. We obtained and minutely analyzed the relevant parameters of each most stable adsorption configuration to explore the microscopic mechanism during gas adsorption. Theoretical results reveal that Pd3-graphene shows great adsorption capacity and sensitivity toward those decompositions. High adsorption energies and abundant charge transfer amounts could guarantee a stable adsorption structure of decomposition gases on Pd3-graphene surface. The complex change of density of states verifies a strong chemical reaction between the gases and the surface. Moreover, the conductivity of Pd3-graphene would improve due to the decrease of energy gap, and the sensitivity was calculated as SOF2 > H2S > SO2 > SO2 F2. This work provides an effective method to evaluate the operation status of SF6 gas-insulated equipment.

1. Introduction

Owing to its outstanding insulation and arc-extinguishing properties, SF6 has been widely applied in gas-insulated switchgear (GIS) [1,2]. However, in the long-term operation of GIS, several inevitable insulation defects may cause partial discharge and partial overheat, which will lead to the initial decomposition of SF6 [3,4]. Simultaneously, SF6 will ultimately decompose to several characteristic products with the reaction of trace O2 and H2O in SF6-insulated equipment: H2S, SO2, SOF2 and SO2 F2 gas [5,6,7]. These decomposition products can corrode the metal parts inside the equipment, accelerate the aging of the insulation medium—and even result in the sudden failure of GIS [8,9]. Therefore, discovering SF6 decomposition products and dealing with insulation defects in a timely manner have great importance. To date, several methods such as gas chromatography, mass spectrometry and Fourier transform infrared spectroscopy, are suggested to detect SF6 products [10,11]. However, these methods are either inaccurate or complicated, so none of them are used in detecting SF6 decomposition products.
In recent years, the gas sensor method has been used in various fields due to its advantages such as high sensitivity, rapid response and small size. Graphene is a 2D material with a unique 2D monoatomic layer structure and an electronic energy band structure. Its excellent characteristics—including high electron mobility, high thermal conductivity, brilliant mechanical properties and large specific surface area—make it a promising gas sensor material [12,13,14,15,16,17]. However, in actual production, several intrinsic defects in graphene may be observed, such as single defects, double-vacancy defects and Stone–Wales defects [18]. Lattice defects may cause local charge traps on the graphene surface, which has a substantial effect on improving the electronic structure and the adsorption capacity of graphene. Liu et al. anchored Sc, Zn, Mo, Ru, Rh, Pd and Ag atoms on defective graphene for N2 reduction reaction and found that a single Mo embedded on nitrogen-doped 555–777 graphene shows eminent catalytic performance [19]. Arokiyanathan et al. calculated the adsorption properties of small Li clusters on graphene with Stone–Wales defect. The results showed that Li clusters have a strong interaction with the defect region [20]. Ma et al. reported hydrogen adsorption on Co-4-doped defective graphene and discovered that point defects in graphene can effectively improve the hydrogen storage capacity of Co-4 [21]. Despite much research on defective graphene [22,23,24,25], few relevant research on SF6 decomposition product detection have been reported.
555–777 graphene is a double-vacancy defect graphene sheet [18]. The absence of two carbon atoms does not destroy the original SP2 hybrid orbital network but forms a stable topological hole. Thus, 555–777 graphene has a more stable electronic structure compared with other defective structures. Previous studies showed that doping transition metal elements on graphene surface can improve its electrical conductivity and enhance its capacity to adsorb gas molecules [26,27,28,29,30]. Liu et al. calculated the adsorption of SF6 decompositions on Pd (111) surface and found that Pd (111) would have great adsorption capacity on absorbing those products. Thus, doping Pd element on 555–777 graphene may be a potential method to improve its adsorption properties [30]. In this study, Pd3-doped 555–777 graphene (Pd3-graphene) is carried out to detect SF6 decomposition products (H2S, SO2, SOF2 and SO2 F2). Different adsorption structures of each gas are obtained to find the most stable adsorption structure. Furthermore, adsorption energy, charge transfer, density of states (DOS) and deformation charge density (DCD) of Pd3-graphene surface before and after gas adsorption are calculated to reveal the gas response mechanism. Overall, our work aims to provide some theoretical basis for developing a novel graphene sensor.

2. Computational Details and Models

In this work, all calculations are carried out using Dmol3 package based on density functional theory (DFT) [31]. Generalized gradient approximation and Perdew–Burke–Ernzerhof function are utilized to deal with the electron exchange-correlation energy and interaction effect of electrons [32,33]. DFT semi-core pseudopots are set to recoup the relativistic effect of Pd atoms. The Brillouin zone is sampled with 6 × 6 × 1 k-points in the Monkhorst–Pack grid [34]. Double numeric plus polarization basis sets are used to obtain a higher calculation precision of hydrogen bond [35]. The smearing is set as 0.001 Ha to ensure the accuracy of the DFT calculation. The SCF tolerance is 1 × 10−6 Ha and spin polarization is considered due to the magnetic properties of Pd atoms. Additionally, total energy of convergence tolerance, maximum force and maximum displacement are set as 1 × 10−5 Ha, 2 × 10−3 Ha/Å and 5 × 10−3 Å, respectively.
Intrinsic 555–777 graphene surface consists of a 6 × 6 supercell, and the lattice parameter is 14.760 Å × 14.760 Å. The vacuum area is set as 20 Å to avoid the interaction of neighboring slabs. The Pd3 cluster is a triangular structure composed of three Pd atoms. Through DFT calculations, various parameters of stable adsorption structures can be acquired, including adsorption energy (Ead), charge transfer amount (QT), adsorption distance (D) and energy gap (Eg). The definition of these parameters are the same as the previous study [36].

3. Results and Discussions

3.1. Geometric Optimization

We apply geometric optimization to minimize their energy and obtain the most stable status of H2S, SO2, SOF2, SO2 F2 and Pd3-graphene. The optimized structures are shown in Figure 1 (adapted from [36]). Figure 1a shows the stable structure of intrinsic 555–777 graphene with three pentagons and heptagons at the center of the graphene sheet. Considering that the charge trap in the center of the hole has a strong binding ability to electrons and the special D electronic structure of Pd metal, we dope the Pd3 cluster at the defect center of graphene to form a stable anchored structure. Figure 1b shows that the Pd3 cluster is parallel to the graphene surface, indicating that the force of the crystal plane on the three Pd atoms is almost equal. Bond lengths between the Pd atoms are 2.714, 2.714 and 2.812 Å. Millikan charge analysis shows that 0.16 electrons transfer from the Pd3 cluster to the graphene sheet after geometric optimization, showing an evident orbital hybridization between C and Pd atoms. The electronic structure of the graphene surface will be greatly changed, which may substantially improve gas adsorption capacity.

3.2. Adsorption Systems

3.2.1. H2S Adsorption on Pd3-Graphene

The H2S molecule is placed in different distances and angles to approach the Pd3-graphene surface and obtain the most stable adsorption structure of H2S gas. Considering the symmetry of H2S molecule, its two typical stable adsorption structures are acquired through H and S atoms approaching the Pd3-graphene surface, as shown in Figure 2. Ead, QT and D are listed in Table 1.
Figure 2a shows the adsorption configuration of system M1. The H2S molecule approaches the surface by the S atom and forms a stable interaction between the S and Pd atoms. The D of M1 is 2.338 Å, revealing that the reaction is relatively strong. When the adsorption energy is greater than 0.6 eV, the adsorption is generally considered chemical. The corresponding Ead of M1 is −1.211 eV; hence, the adsorption is chemical, and the process is spontaneous. The positive value of QT of M1 (0.286) demonstrates that 0.286 electrons transfer from the H2S molecule to the Pd3-graphene.
In Figure 2b, the H2S molecule approaches the Pd3-graphene by H atom, forming a new ionic bond with the Pd atom. The length of the H–Pd bond is 1.868 Å, indicating a strong chemical interaction between H and Pd atoms. Similar to system M1, 0.279 electrons transfer from the H2S molecule to the sheet during optimization. The Ead of system M2 (−1.185 eV) is smaller than that of M1 (−1.211 eV), which means that system M1 releases more energy than M2 after H2S adsorption. Above all, according to the greater Ead and QT of configuration M1, M1 is the most stable adsorption structure of H2S on Pd3-graphene surface.
DOS and partial density of states (PDOS) of the most stable adsorption structure (M1) are calculated to analyze the mechanism of H2S adsorption further, as shown in Figure 3. In Figure 3a, the black and red lines represent the DOS before and after H2S adsorption, respectively. Several distinct changes are observed in DOS after H2S adsorption, which mainly reflect in the increase of DOS from −8 eV to −4 eV. The DOS at the Fermi level rises slightly, revealing that electrons can more easily transfer from the valence band to the conduction band, which may have a substantial effect on improving its conductivity. The PDOS in Figure 3b shows that the increased area of DOS from −8 eV to −4 eV mainly consists of the 3p orbit of S atom. The large overlapped area from −7 eV to −3 eV between the 4d orbit of Pd atom and the 3p orbit of S atom manifests that the chemical reaction between these atoms is strong, which is consistent with the above analysis.

3.2.2. SO2 Adsorption on Pd3-Graphene

For the adsorption of SO2 on Pd3-graphene surface, various initial positions of SO2 molecule are set to investigate the most stable adsorption system. After geometric optimization, two typical stable adsorption configurations are obtained. Figure 4 shows the adsorption structures through S and O atoms approaching the sheet. Figure 4a shows that the SO2 molecule forms a stable adsorption structure by building a chemical bond between S and Pd atoms. The length of the S–Pd bond is 2.192 Å, declaring a strong interaction between them. The two O atoms are far from the sheet; thus, no direct interaction may exist between them and the crystal plane. In Figure 4b the SO2 molecule approaches the Pd3 cluster by O atoms, forming two chemical bonds with the Pd atoms. The length of both O–Pd bonds is 2.131 Å, which is shorter than that of the S–Pd bond in system M3. Thus, the chemical reaction in system M4 may be stronger than that in M3.
Table 2 presents the related parameters of configuration M3 and M4. Both their Ead are negative, manifesting that the adsorption is exothermic and proceeds spontaneously. By contrast, the Ead of system M4 is greater than that of system M3, implying that the SO2 molecule absorbed on the Pd3-graphene surface through two O atoms is the most stable structure. Given the strong reaction between O and Pd atoms in system M4, 0.283 electrons transfer from the sheet to SO2 molecule. For system M3, the small QT means that the ionic S–Pd bond is relatively easy to break.
The DOS and PDOS of M4 are intensively calculated to analyze the adsorption mechanism because M4 is the most stable adsorption configuration of SO2 on Pd3-graphene. Figure 5a presents that the DOS of M4 exhibits three new peaks after SO2 adsorption at around −8, −6 and −4 eV. PDOS shows that the new peaks mainly result from the 2p orbit of O atom and the 3p orbit of S atom. The small increase of DOS at fermi level demonstrates an increase of the conductivity of Pd3-graphene. In addition, a large overlapped area between the 2p orbit of O atom and the 4d orbit of Pd atom ranges from −5 eV to −3 eV, reflecting a complex hybridization of those orbits. Therefore, the chemical reaction between O atoms and Pd atom is strong enough to form a stable adsorption configuration.

3.2.3. SOF2 Adsorption on Pd3-Graphene

Subsequently, we place the SOF2 molecule in different directions and distances approaching the sheet to explore the most stable adsorption structure. Figure 6 shows three typical adsorption configurations, in which the SOF2 molecule is absorbed on Pd3-graphene surface by S, F and O atoms. Related parameters are listed in Table 3.
Figure 6a shows that the SOF2 molecule approaches the surface by the S atom and forms a stable adsorption structure ultimately. The S atom is trapped by the Pd atom and builds a new chemical bond with a bond length of 2.170 Å. The Ead of system M5 listed in Table 3 is −1.230 eV, which indicates that the adsorption is chemical, and the structure is stable enough. During adsorption, only 0.038 electrons transfer from SOF2 molecule to the sheet, so their electronic structure may change slightly. For system M6 shown in Figure 6b, SOF2 molecule absorbs on the Pd3-graphene surface by forming an ionic bond length of 2.124 Å between F and Pd atoms. The Ead of M6 (−0.284 eV) is smaller than that of M5 (−1.230 eV), revealing that M5 is more stable than M6. In system M7, the SOF2 molecule absorbs directly above the hollow position of Pd3 cluster. The length of the O–Pd bond is 2.299 Å, which is longer than that of M5 and M6. However, the specific Ead of M7 (−1.282 eV) is greater than that of M5 (−1.230 eV), implying that system M7 is more stable. Opposite to system M5, 0.132 electrons transfer from the sheet to SOF2 molecule after absorbing on Pd3-graphene. Hence, the SOF2 molecule tends to be absorbed on the Pd3-graphene surface in the form of M7, establishing an adsorption system with the lowest energy.
Figure 7 shows the DOS and PDOS of the most stable adsorption structure of SOF2 on the Pd3-graphene surface. The DOS of M7 structure increases distinctly in the range from −9 eV to −7.5 eV and −6 eV to −3 eV after SOF2 adsorption. These changes are mainly caused by the 2p orbit of O atom and the 2p orbit of F atom. A slight increase of DOS near the fermi level is observed, signifying that SOF2 adsorption can improve the conductivity of Pd3-graphene. The PDOS displaces the hybridization of individual atoms during adsorption. The 2p orbit of O atom, the 2p of F atom and the 4d orbit of Pd atom overlap from −6 eV to −1 eV, suggesting that the chemical reaction between them is considered strong, and the electronic structure of the SOF2 molecule is active.

3.2.4. SO2 F2 Adsorption on Pd3-Graphene

Eventually, the SO2 F2 molecule approaches the Pd3-graphene surface in various directions and distances to find the most stable adsorption structure. After adsorption structure optimization, two typical adsorption systems are obtained, as shown in Figure 8. Correlative parameters of adsorption systems are listed in Table 4.
Figure 8a presents the adsorption configuration of SO2 F2 molecule approaching the Pd3-graphene surface by F atom. The Pd atom builds a chemical connection with the F atom during the approaching process and forms an F–Pd bond with length of 2.014 Å ultimately. The S–F bond is stretched to 2.40 Å due to the radial force of the Pd atom. In system M9, the SO2 F2 molecule approaches the sheet by O atom. Similarly, the O atom forms a chemical bond with the Pd atom, but the length of the O–Pd bond (2.181 Å) is longer than that of the F–Pd bond (2.014 Å), implying a stronger interaction in configuration M8. The Ead of M8 (−0.920 eV) is greater than that of M9 (−0.804 eV), demonstrating that the M8 structure is more stable than the M9 structure. Additionally, 0.705 electrons transfer from the sheet to the SO2 F2 molecule in M8, which is 2.4 times of M9, confirming a stronger oxidation of the F atom. Thus, configuration M8 is supposed to be the most stable adsorption structure of SO2 F2 on the Pd3-graphene surface.
Figure 9a shows that DOS increases substantially in the range of −7 to 0.5 eV. PDOS reveals that the 2p orbit of O atom and the 2p orbit of F atom mainly contribute the increased area, that is, the DOS near the fermi level rises slightly. Thus, the conductivity of Pd3-graphene is expected to improve after SO2 F2 adsorption. According to the PDOS, a large hybridization is observed between the 2p orbit of F atom and the 4d orbit of Pd atom from −5 eV to −1 eV, implying that these orbits are relatively active, while SO2 F2 approaches the surface.

3.2.5. SF6 Adsorption on Pd3-Graphene

In order to ensure that the Pd3-graphene sensor can be utilized in GIS, we further calculated the adsorption of SF6 gas on Pd3-graphene surface to eliminate the interference of SF6 gas. Same as other gases, SF6 approaches the graphene surface at different angles and distances. After the geometric optimization, we obtained the most stable adsorption structure of SF6 on Pd3-graphene surface shown in Figure 10. As presented, SF6 molecule approaches the Pd3-graphene surface by F atom and forms a weak interaction with the Pd atom. The adsorption distance of SF6 is 4.060 Å, which is longer than other gases. Moreover, the Ead of SF6 is calculated to be–0.124 eV, reflecting that the adsorption the adsorption system is not stable enough. During the adsorption, 0.162 electrons transfer from the Pd3-graphene to SF6 molecule. In conclusion, the Pd3-graphene can be a potential sensor applied in the SF6 gas-insulated equipment.

3.3. Electronic Properties

The deformation charge density of pure Pd3-graphene and the four most stable adsorption systems is calculated to investigate the difference in electronic structure before and after modifying or gas adsorption. As presented in Figure 11, the red region means an increase of charge density after adsorption, whereas the blue region means a decrease. In Figure 11a, the charge density of Pd3 cluster decreases, revealing that the defect of 555–777 graphene has a strong oxidation during the doping process. Although the absence of carbon atoms do not give rise to dangling bonds, the reconstruction of the graphene structure results in the change of bond lengths. Meanwhile, the defect would cause a rehybridization of the sigma and pi orbitals of carbon atoms, which would enhance the electron activity near the defect. When the transition metal approaches the defect, its special d electrons form a strong chemical interaction with the defect. Figure 11b shows that the charge density of the two Pd atoms decreases, while the S atom receives electrons from the Pd3 cluster after adsorption. Therefore, the adsorption reaction is concentrated between the S atom and the Pd atom. In Figure 11c, electrons transfer from the Pd atoms to the O atom due to the strong electronegativity of the O atom, indicating that the two O–Pd bonds are stable enough to support the adsorption structure. The O atom also receives electrons from the S atom, such that the electronic structure inside the SO2 gas changes prominently. As for SOF2 gas, the charge density neighboring the O atom increases to a certain degree, and the charge density of the Pd atom decreases. In principle, the O atom performs as an electron acceptor during SOF2 adsorption. In Figure 11e, the DCD of system M8 shows that the F atom receives electrons from the Pd atom and the nearby the S atom, verifying the fairly strong oxidability of the F atom. In conclusion, the Pd3 cluster behaves as an electron donator during gas adsorption and leads to a violent chemical reaction in the vicinity, certifying that the Pd3 dopant can substantially enhance the adsorption capacity of intrinsic 555–777 graphene.
According to the evident effect on the electronic structure of Pd3-graphene after SF6 product adsorption, the sensitivity and selectivity for application of chemical gas sensor must be further investigated. As a resistor-type gas sensor, the change in conductivity is an important factor in detecting SF6 decomposition products. The conductivity ( σ ) of Pd3-graphene gas sensor could be evaluated by the following formula [37]:
σ     e ( E g 2 k T ) ,
where k, T and Eg represent the Boltzmann constant, temperature and HOMO–LUMO energy gap, respectively. Therefore, under a certain temperature condition, σ is an exponential function of the Eg and a smaller energy gap would determine a higher conductivity. We perform frontier molecular orbital theory to calculate the energy of HOMO (EH) and LUMO (EL) and obtain the specific Eg, which the difference between them.
Figure 12 intuitively presents the EH, EL and Eg of pristine Pd3-graphene and the four most stable adsorption systems, where the black number represents the Eg. For pristine Pd3-graphene, the Eg is 0.086 eV, signifying that the basic conductivity is relatively high, which probably results from the Pd3 dopant. After gas adsorption, the Eg of different systems decreases without exception. While H2S and SO2 absorbing on the Pd3-graphene surface in the most stable configuration, the Eg decreases by 13.95% and 9.30%, respectively, manifesting an increase in conductivity of the Pd3-graphene sensor. For the SOF2 adsorption system, the Eg of Pd3-graphene decreases by 38.37%, which is the largest among the four adsorption systems, indicating that the electronic structure is more continuous, and the electronic transition is effortless. However, for the SO2 F2 adsorption system, Eg reduces with the smallest degree, only 5.81%, which means the conductivity of the Pd3-graphene sensor has a slight rise. In conclusion, the conductivity of Pd3-graphene sensor increases after gas adsorption, which agrees with the analysis in the DOS part.
As for N3&Ni doped 555–777 graphene reported in [36], N3&Ni-graphene has a high sensitivity on absorbing H2S and SO2 gas while the Eg decreases from 0.426 eV to 0.052 eV and 0.138 eV, respectively. Hence, the sensitivity of N3&Ni-graphene may be higher than Pd3-graphene while absorbing H2S and SO2 gas. However, it is toilsome for N3&Ni-graphene to distinguish SOF2 molecule and SO2 F2 due to their similar Eg. The problem can be solved by using Pd3-graphene gas sensor for its obviously different Eg of various adsorption systems. As a result, Pd3-graphene can be a better gas sensor material on detecting SF6 decomposition product.

4. Conclusions

In this work, we propose Pd3-cluster-doped 555–777 graphene as a novel resistor-type gas sensor to detect SF6 decomposition products (H2S, SO2, SOF2 and SO2 F2), and the calculation results are as follows:
(I) Adsorption energies of the most stable adsorption systems of H2S, SO2, SOF2 and SO2 F2 are −1.211, −1.591, −1.282 and −0.920 eV, respectively, demonstrating that the four gases could be absorbed on Pd3-graphene surface by chemisorption. Charge transfer makes various atoms form ionic bonds during adsorption. In addition, the change of DOS and PDOS verifies high hybridizations of different atomic orbits.
(II) The conductivity of Pd3-graphene would be improved without exception due to the decrease of Eg after gas adsorption. SOF2 gas adsorption leads to the largest increase, whereas SO2 F2 is the smallest. According to the reduced value of Eg, the sensitivity of the four gases follows the order SOF2 > H2S > SO2 > SO2 F2.
Thus, we confirm that Pd3-graphene can absorb SF6 decomposition products, and the Pd3-graphene sensor can be employed to evaluate the insulation condition of GIS by detecting these decompositions.

Author Contributions

Conceptualization and methodology, J.L. and L.P.; software, formal analysis, data curation and writing—original draft preparation, J.L.; writing—review and editing, F.C., X.Y. and F.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

We gratefully acknowledge the financial support from the National Key Research and Development Program of China under Grant 2017YFB0902500 and the Science and Technology Project of SGCC through the Key Technology of Environment Friendly Gas-Insulated Transmission Line.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Knobloch, H. The Comparison of Arc-Extinguishing Capability of Sulfur Hexafluoride (SF6) with Alternative Gases in High-Voltage Circuit-Breakers. In Gaseous Dielectrics VIII; Springer: Boston, MA, USA, 1998; pp. 565–571. [Google Scholar]
  2. Chee-Hing, D.J.; Srivastava, K.D. Insulation Performance of Dielectric-Coated Electrodes in Sulphur Hexafluoride Gas. IEEE Trans. Electr. Insul. 1975, EI-10, 119–124. [Google Scholar] [CrossRef]
  3. Sauers, I. By-product formation in spark breakdown of SF6/O2 mixtures. Plasma Chem. Plasma Process. 1988, 8, 247–262. [Google Scholar] [CrossRef]
  4. Casanovas, A.M.; Casanovas, J.; Lagarde, F.; Belarbi, A. Study of the decomposition of SF6 under dc negative polarity corona discharges (point-to-plane geometry): Influence of the metal constituting the plane electrode. J. Appl. Phys. 1992, 72, 3344–3354. [Google Scholar] [CrossRef]
  5. Zhang, X.; Gui, Y.; Dai, Z. Adsorption of gases from SF6 decomposition on aluminum-doped SWCNTs: A density functional theory study. Eur. Phys. J. D 2015, 69, 185. [Google Scholar] [CrossRef]
  6. Ju, T.; Fan, L.; 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]
  7. Zhang, X.; Chen, Q.; Hu, W.; Zhang, J. A DFT study of SF6 decomposed gas adsorption on an anatase (101) surface. Appl. Surf. Sci. 2013, 286, 47–53. [Google Scholar] [CrossRef]
  8. Ju, T.; Zeng, F.; Pan, J.; Zhang, X.; Qiang, Y.; He, J.; Hou, X. Correlation analysis between formation process of SF6 decomposed components and partial discharge qualities. IEEE Trans. Dielectr. Electr. Insul. 2013, 20, 864–875. [Google Scholar]
  9. Fridman, A.; Chirokov, A.; Gutsol, A. Non-Thermal Atmospheric Pressure Discharges. J. Phys. D Appl. Phys. 2005, 38, R1. [Google Scholar] [CrossRef]
  10. Kóréh, O.; Rikker, T.; Molnár, G.; Mahara, B.M.; Torkos, K.; Borossay, J. Study of decomposition of sulphur hexafluoride by gas chromatography/mass spectrometry. Rapid Commun. Mass Spectrom. 1997, 11, 1643–1648. [Google Scholar] [CrossRef]
  11. Zhang, X.X.; Wang, Z.; Hu, Y.G.; Ren, J.B. Recognition of SF6 decomposition components by two-dimensional infrared spectroscopy. Gaodianya Jishu/High Volt. Eng. 2010, 36, 1475–1479. [Google Scholar]
  12. Yuan, W.; Shi, G. Graphene-based gas sensors. J. Mater. Chem. A 2013, 1, 10078–10091. [Google Scholar] [CrossRef]
  13. Liu, J. Adsorption of DNA onto gold nanoparticles and graphene oxide: Surface science and applications. Phys. Chem. Chem. Phys. 2012, 14, 10485–10496. [Google Scholar] [CrossRef] [Green Version]
  14. Wei, W.; Nong, J.; Zhang, G.; Tang, L.; Zhu, Y. Graphene-Based Long-Period Fiber Grating Surface Plasmon Resonance Sensor for High-Sensitivity Gas Sensing. Sensors 2016, 17, 2. [Google Scholar] [CrossRef] [PubMed]
  15. Singh, G.; Choudhary, A.; Haranath, D.; Joshi, A.G.; Singh, N.; Singh, S.; Pasricha, R. ZnO decorated luminescent graphene as a potential gas sensor at room temperature. Carbon 2012, 50, 385–394. [Google Scholar] [CrossRef]
  16. Cen, C.; Zhang, Y.; Chen, X.; Yang, H.; Wu, P. A dual-band metamaterial absorber for graphene surface plasmon resonance at terahertz frequency. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 117, 113840. [Google Scholar] [CrossRef]
  17. Wu, P.; Chen, Z.; Xu, D.; Zhang, C.; Jian, R. A Narrow Dual-Band Monolayer Unpatterned Graphene-Based Perfect Absorber with Critical Coupling in the Near Infrared. Micromachines 2020, 11, 58. [Google Scholar] [CrossRef] [Green Version]
  18. Banhart, F.; Kotakoski, J.; Krasheninnikov, A.V. Structural Defects in Graphene. ACS Nano 2011, 5, 26–41. [Google Scholar] [CrossRef] [Green Version]
  19. Liu, P.; Fu, C.; Li, Y.; Wei, H. Theoretical screening of single atom anchored on defective graphene for electrocatalytic N2 reduction reactions: A DFT study. Phys. Chem. Chem. Phys. 2020, 22, 9322–9329. [Google Scholar] [CrossRef]
  20. Arokiyanathan, A.L.; Panjulingam, N.; Lakshmipathi, S. Chemical Properties of Lithium Cluster (Lix, x = 2–8) on Stone–Wales Defect Graphene Sheet: A DFT Study. J. Phys. Chem. C 2020, 124, 7229–7237. [Google Scholar] [CrossRef]
  21. Ma, S.; Chen, J.; Wang, L.; Jiao, Z. First-principles insight into hydrogen adsorption over Co-4 anchored on defective graphene. Appl. Surf. Sci. 2020, 504, 144413. [Google Scholar] [CrossRef]
  22. Olsson, E.; Hussain, T.; Karton, A.; Cai, Q. The adsorption and migration behavior of divalent metals (Mg, Ca, and Zn) on pristine and defective graphene. Carbon 2020, 163, 276–287. [Google Scholar] [CrossRef]
  23. Guan, Z.Y.; Ni, S.; Hu, S. First-Principles Study of 3d Transition-Metal-Atom Adsorption onto Graphene Embedded with the Extended Line Defect. ACS Chem. Biol. 2020, 5, 5900–5910. [Google Scholar] [CrossRef] [Green Version]
  24. Zhang, J.; Deng, Y.; Cai, X.; Chen, Y.; Peng, M.; Jia, Z.; Jiang, Z.; Ren, P.; Yao, S.; Xie, J. Tin Assisted Fully Exposed Platinum Clusters Stabilized on Defect-Rich Graphene for Dehydrogenation Reaction. ACS Catal. 2019, 9, 5998–6005. [Google Scholar] [CrossRef]
  25. Olsson, E.; Chai, G.; Dove, M.; Cai, Q. Adsorption and migration of alkali metals (Li, Na, and K) on pristine and defective graphene surfaces. Nanoscale 2019, 11, 5274–5284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Gui, Y.; Peng, X.; Liu, K.; Ding, Z. Adsorption of C2H2, CH4 and CO on Mn-doped graphene: Atomic, electronic, and gas-sensing properties. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 119, 113959. [Google Scholar] [CrossRef]
  27. Gao, X.; Zhou, Q.; Wang, J.; Xu, L.; Zeng, W. Adsorption of SO2 molecule on Ni-doped and Pd-doped graphene based on first-principle study. Appl. Surf. Sci. 2020, 517, 146180. [Google Scholar] [CrossRef]
  28. Zhang, X.; Cui, H.; Gui, Y. Synthesis of Graphene-Based Sensors and Application on Detecting SF6 Decomposing Products: A Review. Sensors 2017, 17, 363. [Google Scholar] [CrossRef]
  29. Bae, G.; Jeon, I.S.; Jang, M.; Song, W.; Myung, S.; Lim, J.; Lee, S.S.; Jung, H.-K.; Park, C.-Y.; An, K.-S. Complementary Dual-Channel Gas Sensor Devices Based on a Role-Allocated ZnO/Graphene Hybrid Heterostructure. ACS Appl. Mater. Interfaces 2019, 11, 16830–16837. [Google Scholar] [CrossRef]
  30. 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. [Google Scholar] [CrossRef]
  31. Delley, B.J. From Molecules to Solids with the DMol3 Approach. J. Chem. Phys. 2000, 113, 7756–7764. [Google Scholar] [CrossRef]
  32. White, J.A.; Bird, D.M. Implementation of gradient-corrected exchange-correlation potentials in Car-Parrinello total-energy calculations. Phys. Rev. B Condens. Matter 1994, 50, 4954–4957. [Google Scholar] [CrossRef] [PubMed]
  33. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  34. Monkhorstand, H.J.; Pack, J.D. Special points for Brillouin-zoneintegrations. Phys. Rev. B 1977, 16, 5188–5192. [Google Scholar]
  35. Delley, B. An all-electron numerical method for solving the local density functional for polyatomic molecules. J. Chem. Phys. 1990, 92, 508. [Google Scholar] [CrossRef]
  36. Li, J.; Pang, L.; He, K.; Zhang, L. Theoretical Study of The Gas Sensing Mechanism of N3&Ni Doped Double Vacancies Defect Graphene upon SF6 Decompositions. IEEE Access 2019, 7, 145567–145573. [Google Scholar]
  37. Li, S.S. Semiconductor Physical Electronics; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2006. [Google Scholar]
Figure 1. Optimized structures of intrinsic 555–777 graphene, Pd3-graphene and gas molecules (distance in Å). © 2020 IEEE.
Figure 1. Optimized structures of intrinsic 555–777 graphene, Pd3-graphene and gas molecules (distance in Å). © 2020 IEEE.
Sensors 20 04188 g001
Figure 2. Adsorption structures of H2S on Pd3-graphene surface (distance in Å).
Figure 2. Adsorption structures of H2S on Pd3-graphene surface (distance in Å).
Sensors 20 04188 g002
Figure 3. Density of states (DOS) and partial density of states (PDOS) of system M1. Dashed line represents the Fermi level.
Figure 3. Density of states (DOS) and partial density of states (PDOS) of system M1. Dashed line represents the Fermi level.
Sensors 20 04188 g003
Figure 4. Adsorption structures of SO2 on Pd3-graphene surface (distance in Å).
Figure 4. Adsorption structures of SO2 on Pd3-graphene surface (distance in Å).
Sensors 20 04188 g004
Figure 5. Density of states (DOS) and partial density of states (PDOS) of system M4. Dashed line represents the Fermi level.
Figure 5. Density of states (DOS) and partial density of states (PDOS) of system M4. Dashed line represents the Fermi level.
Sensors 20 04188 g005
Figure 6. Adsorption structures of SOF2 on Pd3-graphene surface (distance in Å).
Figure 6. Adsorption structures of SOF2 on Pd3-graphene surface (distance in Å).
Sensors 20 04188 g006
Figure 7. Density of states (DOS) and partial density of states (PDOS) of system M7. Dashed line represents the Fermi level.
Figure 7. Density of states (DOS) and partial density of states (PDOS) of system M7. Dashed line represents the Fermi level.
Sensors 20 04188 g007
Figure 8. Adsorption structures of SO2 F2 on Pd3-graphene surface (distance in Å).
Figure 8. Adsorption structures of SO2 F2 on Pd3-graphene surface (distance in Å).
Sensors 20 04188 g008
Figure 9. Density of states (DOS) and partial density of states (PDOS) of system M8. Dashed line represents the Fermi level.
Figure 9. Density of states (DOS) and partial density of states (PDOS) of system M8. Dashed line represents the Fermi level.
Sensors 20 04188 g009
Figure 10. The most adsorption structure of SF6 on Pd3-graphene surface (distance in Å).
Figure 10. The most adsorption structure of SF6 on Pd3-graphene surface (distance in Å).
Sensors 20 04188 g010
Figure 11. Deformation charge density (DCD) of pure Pd3-graphene and the most stable adsorption structures. (a) Pure Pd3-graphene; (b) H2S; (c) SO2; (d) SOF2; (e) SO2 F2.
Figure 11. Deformation charge density (DCD) of pure Pd3-graphene and the most stable adsorption structures. (a) Pure Pd3-graphene; (b) H2S; (c) SO2; (d) SOF2; (e) SO2 F2.
Sensors 20 04188 g011
Figure 12. EL, EH and Eg of pristine Pd3-graphene and various most stable adsorption systems. From the left to the right are pristine Pd3-graphene, M1, M4, M7 and M8 systems, respectively.
Figure 12. EL, EH and Eg of pristine Pd3-graphene and various most stable adsorption systems. From the left to the right are pristine Pd3-graphene, M1, M4, M7 and M8 systems, respectively.
Sensors 20 04188 g012
Table 1. Structural parameters of H2S adsorption systems on Pd3-graphene surface. Position represents the approaching way, for instance, S–Pd means the H2S molecule approaches the Pd atom by the S atom.
Table 1. Structural parameters of H2S adsorption systems on Pd3-graphene surface. Position represents the approaching way, for instance, S–Pd means the H2S molecule approaches the Pd atom by the S atom.
SystemPositionEad (eV)QT (e)D (Å)
M1S–Pd−1.2110.2862.338
M2H–Pd−1.1850.2791.868
Table 2. Structural parameters of SO2 adsorption systems on Pd3-graphene surface. Position represents the approaching way, for instance, S–Pd means the SO2 molecule approaches the Pd atom by the S atom.
Table 2. Structural parameters of SO2 adsorption systems on Pd3-graphene surface. Position represents the approaching way, for instance, S–Pd means the SO2 molecule approaches the Pd atom by the S atom.
SystemPositionEad (eV)QT (e)D (Å)
M3S–Pd−1.534−0.0232.192
M4O–Pd−1.591−0.2832.131
Table 3. Structural parameters of SOF2 adsorption systems on Pd3-graphene surface. Position represents the approaching way, for instance, S–Pd means the SOF2 molecule approaches the Pd atom by the S atom.
Table 3. Structural parameters of SOF2 adsorption systems on Pd3-graphene surface. Position represents the approaching way, for instance, S–Pd means the SOF2 molecule approaches the Pd atom by the S atom.
SystemPositionEad (eV)QT (e)D (Å)
M5S–Pd−1.2300.0382.170
M6F–Pd−0.284−0.2552.124
M7O–Pd−1.282−0.1322.299
Table 4. Structural parameters of SO2 F2 adsorption systems on Pd3-graphene surface. Position represents the approaching way, for instance, F–Pd means the SO2 F2 molecule approaches the Pd atom by F atom.
Table 4. Structural parameters of SO2 F2 adsorption systems on Pd3-graphene surface. Position represents the approaching way, for instance, F–Pd means the SO2 F2 molecule approaches the Pd atom by F atom.
SystemPositionEad (eV)QT (e)D (Å)
M8F–Pd−0.920−0.7052.014
M9O–Pd−0.804−0.2942.181

Share and Cite

MDPI and ACS Style

Li, J.; Pang, L.; Cai, F.; Yuan, X.; Kong, F. Adsorption Properties of Pd3-Modified Double-Vacancy Defect Graphene toward SF6 Decomposition Products. Sensors 2020, 20, 4188. https://doi.org/10.3390/s20154188

AMA Style

Li J, Pang L, Cai F, Yuan X, Kong F. Adsorption Properties of Pd3-Modified Double-Vacancy Defect Graphene toward SF6 Decomposition Products. Sensors. 2020; 20(15):4188. https://doi.org/10.3390/s20154188

Chicago/Turabian Style

Li, Jie, Lei Pang, Fuwei Cai, Xieyu Yuan, and Fanyu Kong. 2020. "Adsorption Properties of Pd3-Modified Double-Vacancy Defect Graphene toward SF6 Decomposition Products" Sensors 20, no. 15: 4188. https://doi.org/10.3390/s20154188

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