Next Article in Journal
The Advances of Hydrosol–Gel Transition-Based Sensors
Next Article in Special Issue
Ru-Doped PtTe2 Monolayer as a Promising Exhaled Breath Sensor for Early Diagnosis of Lung Cancer: A First-Principles Study
Previous Article in Journal
Low-Tech Test for Mercury Detection: A New Option for Water Quality Assessment
Previous Article in Special Issue
Mo2C-Based Microfluidic Gas Sensor Detects SF6 Decomposition Components: A First-Principles Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption and Sensing of CO2, CH4 and N2O Molecules by Ti-Doped HfSe2 Monolayer Based on the First-Principle

1
State Grid HBEPC. Ltd. Economic & Technology Research Institute, Wuhan 430077, 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, Wuhan University, Wuhan 430072, China
*
Author to whom correspondence should be addressed.
Submission received: 19 September 2022 / Revised: 4 October 2022 / Accepted: 6 October 2022 / Published: 12 October 2022
(This article belongs to the Special Issue Application and Advance of Gas Sensors)

Abstract

:
With the continuous emission of greenhouse gases, the greenhouse effect is becoming more and more serious. CO2, CH4, and N2O are three typical greenhouse gases, and in order to limit their emissions, it is imperative that they are accurately monitored. In this paper, the doping behavior of Ti on the surface of HfSe2 is investigated, based on the first-nature principle. Additionally, the parameters of adsorption energy and the transfer charges of Ti−HfSe2 for CO2, CH4, and N2O are calculated and compared, while the sensing characteristics of Ti−HfSe2 are analyzed. The results show that the structure is most stable when Ti is located above the lower-layer Se atom. The CO2 and N2O adsorption systems with large adsorption energies and transfer charges are a chemical adsorption, while the CH4 system is a physical adsorption with small adsorption energies and transfer charges. In addition, Ti−HfSe2 has a good sensitivity and recovery time for CO2 at 298 K, which is feasible for industrial application. All the contents of this paper provide theoretical guidance for the implementation of Ti−HfSe2 as a gas-sensitive material for the detection of greenhouse gas components.

1. Introduction

With the development of the industry, the environmental impact of greenhouse gases is increasing and the global extreme climate problems caused by greenhouse gases have attracted the attention of countries around the world [1,2,3]. Due to the economy and population growth, anthropogenic greenhouse gas emission has been rising since the beginning of the industrial era. Meanwhile, the concentrations of carbon dioxide (CO2), methane (CH4), and nitrous oxide (N2O) have reached their current highest levels [4]. In order to limit carbon emissions, it is necessary to monitor each emission source.
Currently, greenhouse gases are mainly monitored by means of gas sensors, which include electrochemical gas sensors [5], semiconductor gas sensors [6], infrared gas sensors [7], and new nano-gas sensors. Compared with other gas sensors, the new nano-gas sensors have a specific surface area close to the theoretical extremes, excellent semiconductor properties, and abundant active sites, which lead to the advantages of smaller size and higher accuracy and broaden the application fields of sensors greatly [8,9]. Gas-sensitive material is at the core of nano gas sensors, which mainly includes metal oxide semiconductors [10,11,12], carbon nanotubes [13,14,15,16], graphene [17,18,19], and transition metal dichalcogenides (TMDs) [20,21,22,23]. Among them, TMDs are a new two-dimensional layered nanomaterial with the chemical formula MX2, where M stands for transition metal elements (Mo, W, Ti, Hf, etc.) and X stands for sulfur group elements (S, Se, and Te). Its structure includes three layers, in which two layers of X atoms are sandwiching a single layer of M atoms in between [24]. Compared with graphene, TMDs have greater potential for gas adsorption and sensing because of their certain band gap [25]. However, since TMD sheets tend to form a dense stacking structure during the formation of the conductive network, their sensitivity and recovery time at room temperature still need to be improved. Some studies have shown that doping the surface of adsorbent materials with metals such as Ti, Mn, Ag, and Pt can improve the adsorption performance of adsorbent materials [26,27]. As a representative material in TMDs, MoS2 has received much attention for its application in gas sensors [28]. Li et al. studied the adsorption properties of Mn-doped MoS2 for dissolved gases in transformer oil [29], and Zhang et al. investigated the adsorption properties of Ti-doped MoS2 on CF4, C2F6, and COF2 [30].
However, there are other TMDs with good adsorption properties that have rarely been investigated. HfSe2 has a larger work function than MoS2, which makes HfSe2 less prone to losing electrons. At the same time, HfSe2 experiences a more violent reaction when exposed to air, indicating a strong interaction between HfSe2 and air. Above all, HfSe2 is very suitable for gas sensors [31]. Cui et al. studied the adsorption properties of Pt-doped HfSe2 on SO2 and SOF2, as well as Pd-doped HfSe2 on SO2 and NO2 [32,33]; the results indicate that HfSe2 has great adsorption and sensing properties. As a commonly used doping metal, Ti is often used to dope the surface of nanomaterials to improve their properties [34,35,36]. At present, the research for Ti−doped HfSe2 is not clear enough, so this paper establishes the Ti−doped HfSe2 model based on the first-principle to obtain the most stable structure of Ti−HfSe2 and investigates the adsorption performance of Ti−HfSe2 by analyzing the adsorption energy, transfer of electrons, and other adsorption parameters of CO2, CH4, and N2O adsorption systems. Meanwhile, this paper also calculates the sensitivity and recovery time of Ti−HfSe2 for the three gases to evaluate the feasibility of Ti−HfSe2 as a gas-sensitive material. This study can provide a theoretical basis for the application of Ti−doped HfSe2 nanomaterials in gas sensors.

2. Computation Details

Density functional theory (DFT) is a method for studying multi-electron systems that uses the electron density as the fundamental physical quantity to describe the system [37]. In the field of gas sensing, DFT is also a very effective research tool [38]. The calculations in this paper were all performed in the DMol3 module of Materials Studio based on DFT [39]. In this paper, a 3×3 monolayer HfSe2 supercell containing 16 Hf atoms and 18 Se atoms was constructed. A vacuum layer with a thickness of 15 Å was built to prevent the interaction of periodic HfSe2 crystals in the upper layer. The Perdew–Burke–Ernzerhof (PBE) functional within generalized gradient approximation (GGA) was used to deal with the exchange correlation interaction between electrons, while the Tkatchenko and Scheffler (TS) method was used to describe the van der Waals interactions, using Double Numerical Polarization (DNP) as the atomic orbital basis group [40], and the global orbital cut-off radius was 5 Å. To simplify the process, a DFT semi-core pseudopotential (DSPP) containing relativistic effects was used. The Monkhorst-Pack k-point mesh of 3 × 3 × 1 was used to geometrically optimize the structure in the paper, and the convergence criteria were set as follows: the tolerance accuracy, the maximum force, and the maximum displacement were 10−5 Ha, 0.002 Ha/Å, and 0.005 Å, respectively.

3. Results and Discussions

Firstly, in this paper, the geometry of the gas molecules is optimized. The optimized gas molecules are shown in Figure 1. In Figure 1, both CO2 and N2O have linear structures, where the C-O, N-N, and N-O bond lengths are 1.178 Å, 1.141 Å, and 1.197 Å, respectively, and the structure of CH4 is a regular tetrahedron with a C-H bond length of 1.097 Å.

3.1. Analysis of Ti−HfSe2 Monolayer

For the structure of Ti doped on the surface of HfSe2, three possible doping sites are considered: the sites above the upper-layer Se atom, the Hf atom, and the lower-layer Se atom. The doping sites are shown in Figure 2.
The binding energy Ebind of the Ti atom doped on the surface of HfSe2 is calculated by Equation (1) [41]:
Ebind = ETi−HfSe2ETiEHfSe2
in which ETi−HfSe2, ETi, and EHfSe2 denote the total energy of Ti−HfSe2, the energy of Ti atoms, and the total energy of monolayer HfSe2, respectively. Obviously, the lowest binding energy corresponds to the most stable configuration (MSC). Table 1 indicates the parameters of Ti atoms at different doping sites. Combined with the results of geometry optimization of the adsorption system, it can be found that the geometrical optimization results for the doping site above the upper-layer Se layer is consistent with the doping site above Hf, and both doping sites are not the MSC. When the doping site is above the lower-layer Se, the binding energy is biggest, reaching −5.001 eV, indicating that Ti forms the MSC on the HfSe2 surface. The structure of MSC is shown in Figure 3. The huge binding energy also makes Ti bond with the upper Se layer. Figure 4 shows the electron density difference (EDD) of Ti-HfSe2, where the blue area indicates electron accumulation, and the yellow area indicates electron depletion. One can see that the electron depletion area is mainly concentrated around the Ti atom, and the electron accumulation area is mainly concentrated on the Ti-Se bond, which indicates that Ti transfers 1.034e of electrons to HfSe2.
The density of states (DOS) and partial density of states (PDOS) before and after Ti doping are shown in Figure 5. From Figure 5a, the DOS of the adsorbed material shifts to a lower energy overall, and the shape of the DOS of the adsorbed material is partially changed, with new peaks near −14eV and −5eV after Ti doping. Additionally, the Ti atoms have a great effect on the DOS in the part above the Fermi level, which leads to a higher peak of Ti−HfSe2 at −1eV to 2eV than the peak of HfSe2 pristine. Figure 5b shows the PDOS of the adsorbed material, where Ti 3d orbitals hybridize highly in the range of −6eV to 2eV, with Hf 6s, 5p, and 5d orbitals and Se 4p orbitals producing a strong hybridization. This also indicates that Ti forms a very stable structure with the HfSe2 surface, and the strong electronic orbital hybridization also leads to the formation of chemical bonds between Ti and Se, which is consistent with the previous analysis.

3.2. Adsorption Properties of Ti−HfSe2 Monolayer

The CO2, CH4, and N2O molecules are placed on the Ti−HfSe2 surface in different ways, and these structures are geometrically optimized to finally find the MSC. Through several attempts, it can be found that the MSC exists when the gas molecule is located above the Ti atom. For the CO2 gas molecule, two adsorption structures are selected: parallel and perpendicular close to the Ti−HfSe2 surface. For the CH4 molecule, two adsorption configurations are chosen: C atom perpendicular to the Ti atom and H atom perpendicular to the Ti atom. Due to the asymmetric structure of the N2O gas molecule, there are three adsorption structures: parallel close to the Ti−HfSe2 surface and perpendicular to the Ti−HfSe2 surface with N and O atoms, respectively. Combined with the simulation results, the lowest energy adsorption structure can be obtained as shown in Figure 6.
In order to further analyze the adsorption characteristics of the gas on the Ti−HfSe2 monolayer, the energy change in the adsorption process of the gas molecules on the Ti−HfSe2 surface is taken as the adsorption energy in this paper. The adsorption energy Ead of the gas adsorption process is shown in Equation (2) [42]:
Ead = ETi−HfSe2/gasETi−HfSe2Egas
in which ETi−HfSe2/gas, ETi−HfSe2, and Egas denote the total energy of the system after gas adsorption, the energy of the layer Ti−HfSe2, and the total energy of the gas molecules, respectively.
The charge transfer ΔQ of the adsorption system is calculated by the Mulliken method, and the expression is as shown in Equation (3):
ΔQ = Q1Q2
in which Q1 and Q2 denote the charge of gas molecules after and before adsorption. If ΔQ > 0, this indicates that the gas molecule loses electrons during the adsorption process; if ΔQ < 0, this indicates that the gas molecule gains electrons during the adsorption process. The calculation results are shown in Table 2.
The equation for the EDD is shown in Equation (4):
Δρ = ρgas/Ti−HfSe2ρgasρTi−HfSe2
in which ρgas/Ti−HfSe2, ρgas, and ρTi−HfSe2 denote the electron density of the adsorption system after adsorption of gas molecules, individual gas molecules, and Ti−HfSe2, respectively. The EDD is shown in Figure 7 based on the calculated results, where the blue area indicates electron accumulation, and the yellow area indicates electron depletion.
For the CO2 adsorption system, the large adsorption energy leads to a significant change in the structure of CO2, with the C-O bond length elongated from 1.178 Å to 1.207 Å and 1.350 Å, and with the angle changed to 132.734°. The Ti-Se bond in Ti−HfSe2 increased by about 0.126 Å, following adsorption. In the CH4 adsorption system, the C-H bond in CH4 near Ti−HfSe2 increased by 0.009 Å, and the Ti-Se bond length changed by 0.001 Å. The whole system is almost unchanged, which also indicates that CH4 adsorption on Ti−HfSe2 surface is not effective. There are no bonds formed during the adsorption process, indicating that the adsorption process is a physical adsorption and the main force is Van der Waals (VDW). From the EDD of the CH4 adsorption system, the electron depletion region is concentrated on the surface of H atoms in CH4, and the electron accumulation region is concentrated near the C atoms. The trace electron transferring amount of 0.107e also confirms the weak interaction between CH4 and Ti−HfSe2. The N2O adsorption system changes significantly during the adsorption process, with the Ti-Se bond elongating by 0.032 Å to 0.130 Å, attracted by N2O. From the EDD, it can be seen that a large electron accumulation region is formed near N and O atoms, and a large electron depletion region exists near Ti atoms, indicating that there is a large amount of charge transfer during the adsorption process, further verifying the strong adsorption performance of Ti−HfSe2 on N2O. The CO2 and N2O adsorption systems have large adsorption energies and transfer charges, and both create new bonds during the adsorption process, indicating that these adsorption processes are chemical adsorptions, rebuilding the electron arrangement.

3.3. DOS of Adsorption System

To further analyze the interaction between gas molecules and the surface of the adsorbent material, the DOS and PDOS of three gas adsorption systems are investigated in this section, and the results are shown in Figure 8.
From Figure 8, the DOS and PDOS change to different degrees after the adsorption of gas by Ti−HfSe2. The CO2 adsorption system is shown in Figure 8a. The structure of CO2 is changed after adsorption and the positions of the different peaks of the DOS of CO2 are changed, which leads to three new peaks at −21 eV, −8.5 eV, and −6.5 eV for CO2/Ti−HfSe2. In addition, in the range from −1eV to 0eV, the Ti 3d orbitals, and the C 2p orbitals and O 2p orbitals overlap highly, forming a strong hybridization. This also confirms the bonding between Ti atoms and C and O atoms, respectively, making the structure of CO2 gas molecules change. The adsorption effect is obvious; the adsorption is a chemical adsorption. Figure 8b shows the DOS and PDOS of the CH4 adsorption system. It can be seen that the DOS of CH4 is changed in position and peaks during the adsorption process, which leads to two new peaks in the DOS of the adsorption system at −15 eV and −7.5 eV. It can also be seen that the other parts almost overlap with the DOS of before adsorption, indicating that the interaction between CH4 molecules and the Ti−HfSe2 surface is weak, and only a small amount of charge is transferred during the adsorption process. Meanwhile, the Ti 3d orbitals only have a small overlap with the C 2p and H 1s orbitals. The trace amount of hybridization also indicates that Ti−HfSe2 is less effective in the adsorption of CH4. In the N2O adsorption system, the DOS of the system changed significantly after adsorption, generating new peaks at −20 eV, −10 eV, −7.5 eV, −6.5 eV, and −3 eV, and the newly generated peaks were caused by the adsorbed N2O. There is an obvious overlap between the Ti 3d orbitals and the N 2p and O 2p orbitals near the Fermi level, indicating a strong hybridization between these orbitals and a strong adsorption of N2O by Ti−HfSe2, an adsorption which is sufficient to change the structure of N2O. It also indicates that Ti is bonded to N and O atoms. The adsorption is chemisorption, and the main force is an intermolecular force.

3.4. Gas sensor Analyzation of Ti−HfSe2

The sensitivity of this material and the recovery time after adsorption of each gas are analyzed to evaluate the possibility of Ti−HfSe2 as a resistive gas sensor.
The sensitivity of gas-sensitive materials can be obtained by detecting the change in conductivity of the material before and after adsorption of gas, and the band gap is the main factor that determines the conductivity of the material. Furthermore, there is a quantitative relationship between conductivity (σ) and band gap (Bg), as in Equation (5) [43,44]:
σ = A × e B g 2 k T
in which A is a constant, k is Boltzmann constant, and T(K) denotes the thermodynamic temperature.
The relationship between sensitivity (S) and conductivity (σ) of the gas-sensitive material is shown in Equation (6):
S = 1 σ g a s 1 σ p u r e 1 σ p u r e
in which σgas and σpure denote the conductivity of the adsorption system and Ti−HfSe2, respectively. The change in conductivity of Ti−HfSe2 after adsorbing different gases can be approximated by calculating the change in band gap of each system. When the band gap of the material increases, it is more difficult to excite electrons in the material from the valence band to the conduction band, leading to a decrease in the conductivity of the material; conversely, the conductivity of the material increases.
The band structures of CO2, CH4, and N2O after adsorption are shown in Figure 9. From Equations (5) and (6), the sensitivities of Ti−HfSe2 for CO2, CH4, and N2O at room temperature are 73.34, 123.63, and 1013.97, respectively, which indicates that the conductivity of Ti−HfSe2 has changed significantly after adsorbing three gases above with a high sensitivity. Therefore, Ti−HfSe2 can meet the application requirements of gas detection.
In addition, the desorption performance is also an important index for evaluating gas sensors and is often expressed in terms of the recovery time (τ), which is the time for the gas to desorb from the material surface, and is expressed as shown in Equation (7) [45,46]:
τ = A 1 × e E g k T
in which A is the attempt frequency with a value of 1012s−1, k is the Boltzmann constant, T(K) denotes the thermodynamic temperature, and Eg(eV) denotes the potential barrier; the value of the potential barrier of the desorption process can be considered as the absolute value of the adsorption energy Ead. When the potential barrier is large, the desorption process on the material surface becomes very difficult, and heating is often used to speed up the desorption process.
The recovery time for different gases at different temperatures is shown in Table 3. The recovery times for CO2, CH4, and N2O at room temperature are 38.024 s, 4.852 × 10−4 s, and 1.312 × 1011 s, respectively. CO2 has a fast recovery time at room temperature, which allows Ti−HfSe2 to be recycled quickly to meet the practical use requirements of gas sensors. Therefore, Ti−HfSe2 has significant advantages in terms of gas-sensitive properties and practicality. For the CH4 adsorption system, the recovery time reaches milliseconds at room temperature due to its weaker adsorption energy, and the recovery time is too fast for the gas sensor to detect the gas. The recovery time of N2O at room temperature is longer, at 1.312 × 1011 s. Even when heated to 398 K, the recovery time is still long, indicating that N2O is strongly adsorbed and Ti−HfSe2 can be used as an adsorbent for this gas to reduce its emission.

4. Conclusions

In this paper, the doping model of Ti on the HfSe2 surface is constructed and CO2, CH4, and N2O are adsorbed by Ti−HfSe2. The results show that the MSC of Ti−HfSe2 is when Ti is located above the upper-layer Se atoms of HfSe2 and that Ti−HfSe2 has a good adsorption performance for CO2 and N2O. Furthermore, the sensitivity and recovery time of Ti−HfSe2 to CO2 are great, which allows Ti−HfSe2 to be used to detect CO2. Conversely, Ti−HfSe2 is not suitable for detecting CH4, due to its fast recovery time. Due to the excellent adsorption and excessive recovery time of N2O, Ti−HfSe2 can be considered as an adsorbent for N2O.

Author Contributions

Conceptualization, Y.W. and B.L.; methodology, R.F.; software, B.L.; formal analysis, Y.W. and P.W.; data curation, L.J. and P.W.; writing—original draft preparation, Y.W. and S.T.; writing—review and editing, B.L. and S.T.; visualization, L.J.; supervision, Y.W.; project administration, P.W.; funding acquisition, S.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lv, Y.; Wang, Y.; Sun, Y.; Li, N.; Wang, X.; Hao, R. Status analysis of domestic and foreign strategies for climate change health adaptation. J. Environ. Health 2022, 12, 241–253. [Google Scholar]
  2. 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]
  3. 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, 0263–2241. [Google Scholar] [CrossRef]
  4. Fan, Z.; Song, C.; Qi, X.; Zeng, L.; Wu, F. Accounting of greenhouse gas emissions in the Chinese agricultural system from 1980 to 2020. Sheng Tai Xue Bao 2022, 42, 1–13. [Google Scholar]
  5. Chen, Z.; Li, J.; Zhang, X.; Jiang, F. Progress on Construction and Research of Sensing Films of Molecularly Imprinted Electrochemical Sensors. J. Instrum. 2010, 29, 97–104. [Google Scholar]
  6. Li, W.; Huang, S.; Chen, W. Recent developments of gas sensors for methane detection. J. Fujian Univ. Technol. 2006, 4, 4–8. [Google Scholar]
  7. Zhang, Y.; Zhang, X.; Fu, M.; Wang, D.; Wang, T.; Tian, S. Quantitative Analysis Method of C4F7N and Its Decomposition Products Based on Infrared Spectroscopy Technology. High Volt. Eng. 2022, 48, 1836–1845. [Google Scholar]
  8. Ren, H.; He, J.; Liang, R.; Tian, Z. Research advances of gas nanosensors. Microelectron. Eng. 2003, 40, 16. [Google Scholar]
  9. Liu, K.; Zou, D.; Lian, W.; Ma, L.; Ma, L.; Chen, Z. Research and Application of Nanosensor. Instrum. Exp. Tech. 2008, 1, 3. [Google Scholar]
  10. Zhang, X.; Wu, F.; Tie, J.; Xiong, H.; Jiang, H.; Tang, J. TiO2 Nanotube Array Sensor for Detecting SF6 Decomposition Component SO2F2. High Volt. Eng. 2014, 40, 3396–3402. [Google Scholar]
  11. Zhang, X.; Zhang, J.; Jia, Y.; Xiao, P.; Tang, J. TiO2 nanotube array sensor for detecting SF6 decomposition component SO2. Sensors 2012, 12, 3302–3313. [Google Scholar] [CrossRef] [PubMed]
  12. Tian, Y.; Feng, Q.; Zhou, K.; Liu, P. Effect of optical gas sensing properties rutile phase oxide TiO2, SnO2 and GeO2 surface oxidation with HCl molecules. Chin. Sci. Bull. 2017, 62, 1729–1737. [Google Scholar] [CrossRef]
  13. Zhang, X.; Ren, J.; Xiao, P.; Tang, J.; Yao, Y. Multi-wall Carbon Nanotube Films Sensor Applied to SF6 PD Detection. Proc. CSEE 2009, 29, 114–118. [Google Scholar]
  14. Wang, X.; Zhang, X.; Sun, C.; Yang, B. Surface Modification of Multi-wall Carbon Nanotubes by Dielectric Barrier Discharge in Atmospheric Pressure and the Analysis on Gas Sensitive Characteristics. High Volt. Eng. 2012, 38, 223–228. [Google Scholar]
  15. Zhang, X.; Zhang, J.; Tang, J.; Meng, F.; Liu, W. Ni-doped Carbon Nanotube Sensor for Detecting Dissolved Gases in Transformer Oil. Proc. CSEE 2011, 31, 119–124. [Google Scholar]
  16. Zhang, X.; Meng, F.; Ren, J.; Tang, J.; Yang, B. Simulation on the B-doped Single-walled Carbon Nanotubes Detecting the Partial Discharge of SF6. High Volt. Eng. 2011, 37, 1689–1694. [Google Scholar]
  17. Pi, S.; Zhang, X.; Fang, J.; Chen, D. Different PtNPs Doped Graphene Sensor for Detecting Gas Sensitive Property of SF6 Decomposition Components. High Volt. Appar. 2022, 58, 67–72. [Google Scholar]
  18. Ding, Y.; Diao, Q.; Liu, D.; Liu, A.; Jiao, M.; Zhu, G. Synthesis of Graphene Quantum Dots and Application in Gas Sensing. Chin. J. Anal. Chem. 2022, 50, 495–505. [Google Scholar]
  19. Zhang, S.; Tang, M.; Deng, L.; Wang, D. Research Progress of CO and CO2 Gas Sensors Based on Graphene. Microelectronics 2022, 52, 71–76. [Google Scholar]
  20. 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]
  21. Li, H.; Liu, S.; Feng, Q. Research and Application Progress of Electrochemical Sensors Based on Two-dimensional Layered Semiconductor Materials. Mater. Rep. 2022, 36, 20080298. [Google Scholar]
  22. Yang, Z.; Li, B.; Han, Y.; Su, C.; Chen, W.; Zhou, Z. Gas sensors based on two-dimensional transition metal dichalcogenide nanoheterojunctions. Chin. Sci. Bull. 2019, 64, 3699–3716. [Google Scholar]
  23. Wang, C.; Wang, L.; Xiang, S.; Wang, H.; Wan, S.; Wu, J. Study on adsorption characteristics of Si-modified MoS2 for typicical decomposition components of g3 environmental protection gas. J. At. Mol. Phys. 2022, 40, 051005. [Google Scholar]
  24. Zhang, Y.; Wei, X.; Yu, Y.; Zhang, H. Research Progress on the Application of Transition-metal Dichalcogenides in FET. J. Synth. Cryst. 2017, 46, 868–873. [Google Scholar]
  25. Pham, V.P.; Yeom, G.Y. Recent Advances in Doping of Molybdenum Disulfide: Industrial Applications and Future Prospects. Adv. Mater. 2016, 28, 9024–9059. [Google Scholar] [CrossRef]
  26. Fan, Y.; Zhang, J.; Qiu, Y.; Zhu, J.; Zhang, Y.; Hu, G. A DFT study of transition metal (Fe, Co, Ni, Cu, Ag, Au, Rh, Pd, Pt and Ir)-embedded monolayer MoS2 for gas adsorption. Comp. Mater. Sci. 2017, 138, 255–266. [Google Scholar] [CrossRef]
  27. Gui, Y.; He, X.; Ding, Z. Adsorption of SF6 decomposition components on Pt-doped graphyne monolayer: A DFT study. IEEE Access 2019, 99, 1. [Google Scholar]
  28. Mleczko, M.J.; Zhang, C.; Lee, H.R. Atomically-thin HfSe2 transistors with native metal oxides. In Proceedings of the 74th Annual Device Research Conference, Newark, DE, USA, 19–22 June 2016. [Google Scholar]
  29. Li, Y.; Liu, Q.; Xia, Y.; Hu, H.; Li, G.; Liu, S. Calculation based on the first-principle on the adsorption performance of Mn doped MoS2 for dissolved gases in oil. J. At. Mol. Phys. 2022, 40, 011005. [Google Scholar]
  30. Zhang, X.; Wang, J.; Chen, D.; Tian, S.; Zhang, G.; Wang, L.; Wu, T. Calculation of Adsorption Performance Ti Dopped with MoS2 to CF4, C2F6, and COF2 Based on the First-principle. High Volt. Appar. 2021, 57, 89–96. [Google Scholar]
  31. Mirabelli, G.; Mcgeough, C.; Schmidt, M. Air sensitivity of MoS2, MoSe2, MoTe2, HfS2, and HfSe2. J. Appl. Phys. 2016, 120, 125102. [Google Scholar] [CrossRef] [Green Version]
  32. Cui, H.; Zhu, H.; Jia, P. Adsorption and sensing of SO2 and SOF2 molecule by Pt-doped HfSe2 monolayer: A first-principles study. Appl. Surf. Sci. 2020, 530, 147242. [Google Scholar] [CrossRef]
  33. Cui, H.; Jia, P.; Peng, X. Adsorption of SO2 and NO2 molecule on intrinsic and Pd-doped HfSe2 monolayer: A first-principles study. Appl. Surf. Sci. 2020, 513, 145863. [Google Scholar] [CrossRef]
  34. Zhou, H.; Zheng, J.; Wen, Q.P. The effect of Ti content on the structural and mechanical properties of MoS2-Ti composite coatings deposited by unbalanced magnetron sputtering system. Phys. Procedia 2011, 18, 234–239. [Google Scholar]
  35. Renevier, N.M.; Oosterling, H.; KoNig, U. Performance and limitations of MoS2/Ti composite coated inserts. Surf. Coat. Technol. 2003, 172, 13–23. [Google Scholar] [CrossRef]
  36. Zhang, Y.; Han, H.; Zhang, M.; Zhai, C.; Wang, H.; Chen, Y. Synthesis of NaP Zeolite Doped with Titanium and Its Adsorption Performance. J. Chin. Chem. Soc. 2020, 48, 1966–1975. [Google Scholar]
  37. Wang, D.; Chen, D.; Pi, S.; Zhang, X.; Tang, J. Density Functional Theory Study of SF6 Decomposed Products Over ZnO(0001) with Gas Sensing Properties. CES TEMS 2020, 35, 1592–1602. [Google Scholar]
  38. Zhang, D.; Cao, Y.; Yang, Z.; Wu, J. Nanoheterostructure construction and DFT study of Ni-doped In2O3 nanocubes/WS2 hexagon nanosheets for formaldehyde sensing at room temperature. ACS Appl. Mater. 2020, 12, 11979–11989. [Google Scholar] [CrossRef]
  39. Liu, L.; Yao, H.; Jiang, Z. Theoretical study of methanol synthesis form CO2 hydrogenation on PdCu3(111) Surfaces. Appl. Surf. Sci. 2018, 451, 333–345. [Google Scholar] [CrossRef]
  40. Xiao, S.; Zhang, J.; Zhang, X.; Chen, D.; Fu, M.; Tang, J.; Li, Y. Adsorption Characteristics of γ-Al2O3 for the Environment-friendly Insulation Medium C3F7CN/N2 and Its Decomposition Products. High Volt. Eng. 2018, 44, 3135–3140. [Google Scholar]
  41. Li, D.; Rao, X.; Lei, Z.; Zhang, Y.; Ma, S.; Chen, L.; Yu, Z. First-Principle Insight into the Ru-Doped PtSe2 Monolayer for Detection of H2 and C2H2 in Transformer Oil. ACS Omega 2020, 5, 31872–31879. [Google Scholar] [CrossRef]
  42. Kang, S.G. First-principles prediction of NO2 and SO2 adsorption on MgO/(Mg0.5Ni0.5)O/MgO(100). Appl. Surf. Sci. 2021, 1, 150650. [Google Scholar] [CrossRef]
  43. Wu, H.; Zhang, B.; Li, X.; Hu, X. First-principles screening upon Pd-doped HfSe2 monolayer as an outstanding gas sensor for DGA in transformers. Comput. Theor. Chem. 2022, 1, 1208. [Google Scholar] [CrossRef]
  44. Sharma, A.; Anu; Khan, M.S. Sensing of CO and NO on Cu-doped MoS2 Monolayer Based Single Electron Transistor: A First Principles Study. IEEE Sens. J. 2018, 18, 2853–2860. [Google Scholar] [CrossRef]
  45. Cui, H.; Zhang, X.; Zhang, G.; Tang, J. Pd-doped MoS2 monolayer: A promising candidate for DGA in transformer oil based on DFT method. Appl. Surf. Sci. 2019, 470, 1035. [Google Scholar] [CrossRef]
  46. Zhang, D.; Yang, Z.; Li, P.; Pang, M.; Xue, Q. Flexible self-powered high-performance ammonia sensor based on Au-decorated MoSe2 nanoflowers driven by single layer MoS2-flake piezoelectric nanogenerator. Nano Energy 2019, 65, 103974. [Google Scholar] [CrossRef]
Figure 1. Gas molecular structure. (a) CO2; (b) CH4; (c) N2O.
Figure 1. Gas molecular structure. (a) CO2; (b) CH4; (c) N2O.
Chemosensors 10 00414 g001
Figure 2. HfSe2 pristine and three types of Ti doping. (a) HfSe2 pristine; (b) above the upper-layer Se; (c) above the Hf; (d) above the lower-layer Se.
Figure 2. HfSe2 pristine and three types of Ti doping. (a) HfSe2 pristine; (b) above the upper-layer Se; (c) above the Hf; (d) above the lower-layer Se.
Chemosensors 10 00414 g002
Figure 3. MSC of Ti−HfSe2. (a) Top view; (b) front view.
Figure 3. MSC of Ti−HfSe2. (a) Top view; (b) front view.
Chemosensors 10 00414 g003
Figure 4. EDD of Ti−HfSe2 with isosurface of 0.12e/Å3. (a) Top view; (b) front view.
Figure 4. EDD of Ti−HfSe2 with isosurface of 0.12e/Å3. (a) Top view; (b) front view.
Chemosensors 10 00414 g004
Figure 5. DOS and PDOS of adsorption system. The dash line is Fermi level. (a) DOS of Ti−HfSe2, Ti and HfSe2; (b) PDOS of Ti−HfSe2.
Figure 5. DOS and PDOS of adsorption system. The dash line is Fermi level. (a) DOS of Ti−HfSe2, Ti and HfSe2; (b) PDOS of Ti−HfSe2.
Chemosensors 10 00414 g005
Figure 6. MSC of each adsorption system. (a,b) CO2 adsorption system; (c,d) CH4 adsorption system; (e,f) N2O adsorption system.
Figure 6. MSC of each adsorption system. (a,b) CO2 adsorption system; (c,d) CH4 adsorption system; (e,f) N2O adsorption system.
Chemosensors 10 00414 g006
Figure 7. EDD of adsorption system. (a,b) EDD of CO2 with isosurface of 0.04e/Å3; (c,d) EDD of CH4 with isosurface of 0.02e/Å3; (e,f) EDD of N2O with isosurface of 0.04e/Å3.
Figure 7. EDD of adsorption system. (a,b) EDD of CO2 with isosurface of 0.04e/Å3; (c,d) EDD of CH4 with isosurface of 0.02e/Å3; (e,f) EDD of N2O with isosurface of 0.04e/Å3.
Chemosensors 10 00414 g007
Figure 8. DOS and PDOS of adsorption system. (a) CO2 adsorption system; (b) CH4 adsorption system (b); (c) N2O adsorption system.
Figure 8. DOS and PDOS of adsorption system. (a) CO2 adsorption system; (b) CH4 adsorption system (b); (c) N2O adsorption system.
Chemosensors 10 00414 g008aChemosensors 10 00414 g008b
Figure 9. Band structures of Ti−HfSe2 and each adsorption system. (a) Ti−HfSe2; (b) CO2; (c) CH4; (d) N2O.
Figure 9. Band structures of Ti−HfSe2 and each adsorption system. (a) Ti−HfSe2; (b) CO2; (c) CH4; (d) N2O.
Chemosensors 10 00414 g009
Table 1. Binding energy and distance for different doping positions.
Table 1. Binding energy and distance for different doping positions.
Doped SiteEbind/eVDistance/Å
Above the upper Se−4.730Ti-Se 2.450 Ti-Hf 2.838
Above the Hf−4.730Ti-Se 2.450 Ti-Hf 2.840
Above the lower Se−5.001Ti-Se 2.424 Ti-Hf 3.216
Table 2. Calculation result of CO2, CH4, and N2O adsorption system.
Table 2. Calculation result of CO2, CH4, and N2O adsorption system.
GasEad/eVΔQ/eDistance/Å
CO2−0.813−0.483C-Ti 2.057
CH4−0.5200.107H-Ti 2.357
N2O−1.384−0.468N-Ti 1.916
Table 3. Recovery time of gas at different temperatures.
Table 3. Recovery time of gas at different temperatures.
GasRecovery Time/s
298 K398 K498 K
CO238.0240.0151.626 × 10−4
CH44.852 × 10−42.834 × 10−61.786 × 10−7
N2O1.312 × 10111.490 × 10595.109
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, Y.; Liu, B.; Fang, R.; Jing, L.; Wu, P.; Tian, S. Adsorption and Sensing of CO2, CH4 and N2O Molecules by Ti-Doped HfSe2 Monolayer Based on the First-Principle. Chemosensors 2022, 10, 414. https://0-doi-org.brum.beds.ac.uk/10.3390/chemosensors10100414

AMA Style

Wang Y, Liu B, Fang R, Jing L, Wu P, Tian S. Adsorption and Sensing of CO2, CH4 and N2O Molecules by Ti-Doped HfSe2 Monolayer Based on the First-Principle. Chemosensors. 2022; 10(10):414. https://0-doi-org.brum.beds.ac.uk/10.3390/chemosensors10100414

Chicago/Turabian Style

Wang, Yingxiang, Benli Liu, Rengcun Fang, Lin Jing, Peng Wu, and Shuangshuang Tian. 2022. "Adsorption and Sensing of CO2, CH4 and N2O Molecules by Ti-Doped HfSe2 Monolayer Based on the First-Principle" Chemosensors 10, no. 10: 414. https://0-doi-org.brum.beds.ac.uk/10.3390/chemosensors10100414

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