Next Article in Journal
Alleviation of Ultrafiltration Membrane Fouling by ClO2 Pre-Oxidation: Fouling Mechanism and Interface Characteristics
Next Article in Special Issue
Recent Advances in Mixed-Matrix Membranes for Light Hydrocarbon (C1–C3) Separation
Previous Article in Journal
Influence of Reverse Osmosis Process in Different Operating Conditions on Phenolic Profile and Antioxidant Activity of Conventional and Ecological Cabernet Sauvignon Red Wine
Previous Article in Special Issue
Preparation of Al-Containing ZSM-58 Zeolite Membranes Using Rapid Thermal Processing for CO2/CH4 Mixture Separation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Density Functional Theory Study of B, N, and Si Doped Penta-Graphene as the Potential Gas Sensors for NH3 Detection

1
Faculty of Materials, Metallurgy and Chemistry, Jiangxi University of Science and Technology, Ganzhou 34100, China
2
Jiangxi Advanced Copper Industry Research Institute, Yingtan 335000, China
*
Authors to whom correspondence should be addressed.
Membranes 2022, 12(1), 77; https://doi.org/10.3390/membranes12010077
Submission received: 15 December 2021 / Revised: 31 December 2021 / Accepted: 6 January 2022 / Published: 8 January 2022
(This article belongs to the Special Issue Membranes for Gas Separation and Purification Processes)

Abstract

:
Designing a high-performance gas sensor to efficiently detect the hazardous NH3 molecule is beneficial to air monitoring and pollution control. In this work, the first-principles calculations were employed to investigate the adsorption structures, electronic characteristics, and gas sensing properties of the pristine and B-, N-, P-, Al-, and Si-doped penta-graphene (PG) toward the NH3, H2S, and SO2 molecules. The results indicate that the pristine PG is insensitive to those toxic gases due to the weak adsorption strength and long adsorption distance. Nevertheless, the doping of B, N, Al, and Si (B and Al) results in the transition of NH3 (H2S and SO2) adsorption from physisorption to chemisorption, which is primarily ascribed to the large charge transfer and strong orbital hybridizations between gas molecules and doping atoms. In addition, NH3 adsorption leads to the remarkable variation of electrical conductivity for the B-, N-, and Si-doped PG, and the adsorption strength of NH3 on the B-, N-, and Si-doped PG is larger than that of H2S and SO2. Moreover, the chemically adsorbed NH3 molecule on the N-, B-, and Si-doped PG can be effectively desorbed by injecting electrons into the systems. Those results shed light on the potential application of PG-based nanosheets as reusable gas sensors for NH3 detection.

1. Introduction

In recent years, the excessive emission of poisonous ammonia (NH3) has brought a serious threat to the ecological environment and human health [1]. Long-term exposure to NH3 can cause bronchitis, lung swelling, and even death [2]. In this regard, the highly efficient detection of trace NH3 gas is regarded as an effective method to restrain its negative effects. Nowadays, many researchers have focused on the fields of seeking the high-performance sensing materials for NH3 detection [3,4,5]. Among them, the two-dimensional (2D) nanomaterials have great potential application in gas senors due to the good stability, large surface-to-volume ratio, and massive active sites [6,7,8,9]. For instance, C-vacancy defected Ti2CO2 monolayer was predicted by the first-principles method to be a reusable gas sensor for NH3 detection [10]. In addition, NH3 adsorption based on borophene [11], C2N monolayer [12], phosphorus carbide monolayer [13], and Ni-doped InN monolayer [14] has been theoretically studied to evaluate the possibility of those materials as an NH3 sensor. However, some intrinsic shortcomings, including the insufficient sensitivity, excessively large adsorption strength, and long recovery time, restrict their industrial utilization. Thus, it is necessary to explore the high-performance sensor material for NH3 detection.
Penta-graphene (PG), a new 2D allotrope of graphene, was recently proposed and predicted to have excellent dynamical and mechanical stability up to 1000K [15]. Moreover, many theoretical investigations showed that PG has a great potential application in gas capture and sensing [16,17,18], hydrogen storage [19], and lithium-ion batteries [20] due to its good mechanical, electrical, and optical properties [21,22]. Cheng et al. [23] investigated the electronic and transport properties of NH3, NO, and NO2 adsorbed on PG monolayers by using the first-principles and non-equilibrium Green’s function methods, and revealed that PG was highly sensitive and selective to NO and NO2 molecules, while it showed poor NH3 sensing performance because of the weak physisorption. As known, the heteroatom dopant method has been proven to significantly improve the interactions between gas molecules and substrate [24,25]. For example, the adsorption strength, charge transfer, and sensing properties of CO and CO2 on Fe-doped PG were higher than those on the pristine PG [26]. By using the density functional theory (DFT) calculations, Chen et al. [17] found that the doping of B and N into PG could obviously enhance the adsorption strength of SF6 decomposed products and demonstrated the B-doped PG could be a good sensing material for H2S molecules. In addition, the N-doped PG exhibited high CO2 selectivity from the mixtures of H2, N2, and CH4 in an external electric field [16]. In addition, the Pt-doped PG was reported to have good detection ability of H2 molecules [27]. Nevertheless, the investigations into the sensing properties of PG with different dopants toward the poisonous NH3 are still lacking.
Generally, the gaseous pollutants of NH3, H2S, and SO2 are simultaneously released from the combustion of fuel and industrial production. Therefore, the first-principles calculations based DFT method were performed to study the adsorption behaviors of these toxic gases on the pristine and B-, N-, P-, Al-, and Si-doped PG monolayers, and the corresponding sensing characteristics were discussed in this work. Meanwhile, the adsorption structure and electronic properties (e.g., energy band, density of states, difference charge density, and charge transfer) were also systematically analyzed. Our results indicate that the B-, N-, and Si-doped PG nanosheets could be promising NH3 sensing materials with high selectivity and sensitivity, which provide the fundamental basis for the development of novel PG-based gas sensor.

2. Calculation Method and Details

All the gas sensing simulations on pristine penta-graphene (PG), doped-PG, and charged PG were performed using the Dmol3 module [28] of Materials studio software (Accelrys, California, USA). The generalized gradient approximation in the form of the Perdew–Burke–Ernzerhof [29] and the double numerical plus polarization (DNP) basis set [30] were chosen for all the spin calculations. The Grimme approach [31] and DFT semi-core pseudopotential (DSSP) method [32] were used to describe the van der Waals interactions and core treatment, respectively. The 3 × 3 ×1 supercell of PG (54 atoms) with 20 Å along the z direction was adopted during the simulations, and the Monkhorst−Pack meshes of 4 × 4 ×1 (10 × 10 ×1) were employed to conduct the geometry optimizations (electronic properties calculations), and the Hirshfeld charge was calculated to analyze the charge transfer between gas molecules and substrates. In addition, the global orbital cutoff radius was set as 4.6 Å, and the convergence tolerances were 1.0 × 10−5 Ha, 0.002 Ha/Å, and 0.005 Å for total energy, maximum force, and maximum displacement, respectively. In doped PG systems, the formation energy (Efor) of B, N, P, Al, or Si atoms could be determined by the following equation [17]:
E for = E doped - PG     E PG + E C     E doped - atom
where E PG and E doped - PG are the total energies of pristine and doped PG, and E C and E doped - atom are the chemical potential of C and doped atoms, respectively. The chemical potentials of C (B, P, Al, and, Si) were obtained from the average energy for every C (B, P, Al, and Si) atom in bulk diamond (B, P, Al, and Si), while the chemical potential of the N atom was calculated by the N2 molecule.
The adsorption characteristics of NH3, H2S, and SO2 over pristine and doped PG were studied in this work, and the adsorption energy (Eads) could be obtained by [33,34]
  E ads = E gas + substrate     E substrate     E gas
where E gas + substrate and E substrate are the total energies of pristine PG (doped PG) with and without gas adsorptions, and E gas is the energy of a single gas molecule. As known, a more negative Eads means a stronger gas–substrate interaction.

3. Results and Discussion

3.1. Electronic Properties and Stability of Doped PG

The optimized lattice constant of PG with P-421M symmetry is 3.63 Å, and the calculated band gap is 2.43 eV (Figure 1c), both of which match well with previous reported results [35]. There are two types of carbon atoms, i.e., the sp2 hybridized carbon atom (C1 site) and sp3 hybridized carbon atom (C2 site), as shown in Figure 1a. Thus, the C1 and C2 sites substituted by B, N, P, Al, or Si atoms are both taken into account, and the obtained formation energy of different doped PG is listed Table 1. The difference between our results and other studies may be ascribed to the different supercell size and dopant concentration. One can see that N and P substitutions at the C1 site are the exothermic process, while the other doping cases are the endothermic process. Additionally, all the dopant atoms prefer to substitute the C1 atom due to the smaller formation energy, which is consistent with previous calculated results [17,21]. Therefore, the cases of B, N, P, Al, and Si substituting the C1 site are discussed in the following sections.

3.2. Adsorption of NH3, H2S and SO2 on Pristine PG

In order to evaluate the detection ability of PG, the adsorption behaviors of NH3, H2S, and SO2 on the pristine PG were investigated in this section. Considering the structure symmetry of PG, six possible adsorption sites, including two top sites (labeled as A and B), three bridge sites (labeled as C, D, and E), and one hollow site (labeled as F) were taken into account, as displayed in Figure 1a. For convenience, one gas molecule adsorption on the PG is defined as “gas@PG”, for instance, NH3@PG means the NH3 adsorption on PG. The calculated adsorption energy, adsorption distance, charge transfer, and band gap of three toxic gases adsorbed on PG are presented in Table 2. It is found that the adsorption energies of the NH3, H2S, and SO2 molecules are −0.298 eV∼−0.382 eV, and the adsorption distances are in the range of 2.661 Å to 3.169 Å (Figure 2). This indicates that all the toxic gases belong to the representative physisorption, which can be also demonstrated by the small Hirshfeld charge transferred from PG to gases. Moreover, compared with the pristine PG, the band gap of PG after gas molecule adsorptions shows little change. Therefore, the pristine PG exhibits poor detection ability toward the NH3, H2S, and SO2 molecules due to weak interactions and negligible resistance variation.
To further comprehend the micro-interaction mechanism between gas molecules and PG, the density of states (DOSs) and charge density difference (CDD) of various adsorption systems were calculated, as shown in Figure 3 and Figure 4. For the NH3@PG, H2S@PG, and SO2@PG, the gas–substrate interactions are mainly contributed by the weak hybridizations between PG and p orbitals of gas molecules. By contrast, NH3-s, H2S-s, and SO2-s orbitals show the weaker interactions with the penta-graphene due to the small peak value of DOSs (see Figure 3). For the NH3@PG and H2S@PG, the partial electrons of NH3 or H2S are transferred to PG (Figure 4a,b), indicating the NH3 and H2S molecules act as electron donors, which agrees well with the results of the Hirshfeld charge (Table 1). In addition, compared with the case of H2S, NH3 loses more electrons and thus has the lower adsorption energy. For the SO2@PG, a part of the charge of PG is transferred to the SO2 molecule, as shown in Fig. 4c, which suggests the SO2 molecule acts as the electron acceptor. However, the small transfer charge between gas molecules and PG still results in weak interactions, and thus the pristine PG shows poor detection ability toward these toxic gases.

3.3. Adsorption of NH3, H2S, and SO2 on Doped PG

The adsorption behaviors of NH3, H2S, and SO2 on M-doped penta-graphene (M-PG, M=B, N, Al, and Si) were also investigated, and the obtained calculation results and corresponding lowest energy configurations are displayed in Table 2 and Figure 5. Compared with the NH3 on pristine PG, doping B, N, Al, or Si atoms into PG nanosheets can obviously enhance the NH3 adsorption strength with the Eads of −1.06 eV∼−2.46 eV, resulting in the transition of NH3 adsorption from physisorption to chemisorption. In addition, the adsorption distances are significantly shortened (Figure 5a–d), and the charge number transferred from NH3 to the doped PG is remarkably increased in comparison with the case of NH3 adsorption on the pristine PG. By contrast, only the doping of Al and B atoms can strengthen the adsorption of PG toward H2S and SO2, the rest of the dopants show a slight change (see Table 2). In particular, the H2S is dissociation-chemisorbed on the Al-PG surface with one H bonding to the nearest neighbor C atom, as shown in Figure 5f. Therefore, the doping of B, N, P, Al, and Si atoms may be able to improve the detection ability of the NH3, H2S, and SO2 molecules.
The variation of electric conductivity (σ) could be utilized to evaluate the sensitivity of a material before and after molecular adsorption, which is defined as σ ∝ exp(−Eg/2KBT), where Eg, KB, and T are the band gap, Boltzmann constant, and temperature, respectively. As mentioned above, NH3 and H2S adsorptions lead to a little charge in the band gap of the pristine PG, which demonstrates that PG has poor sensitivity to NH3 and H2S. However, the band gap of B-, N-, and Si-PG undergo great variation after NH3 adsorption, as shown in Figure 6e,f,h, implying that NH3 adsorption has a huge impact on the resistivity of B-, N-, and Si-doped PG, thus these doped PG exhibits high NH3 sensitivity. Furthermore, the adsorption strength of NH3 on the B-, N-, and Si-PG surface is much larger than that of H2S and SO2 on the same substrates (see Table 3). Those results indicate that B-, N-, and Si-PG show excellent NH3 selectivity and sensitivity and could be a promising gas sensor for NH3 detection. In addition, compared with the Al-PG, the B-PG also has good H2S sensitivity due to the obvious change of band gap, as displayed in Figure 6i. Nonetheless, the band gap of Al-PG and B-PG remains unchanged in spite of SO2 chemisorption (Figure 6k,l), which suggests that Al- and B-PG could not be the SO2 sensing materials.
Figure 7 displays the difference charge density of NH3, H2S, and SO2 adsorbed the doped PG. For the NH3 adsorption, one sees that the charge accumulation primarily appears around the N, B, Al, and Si atoms, while the charge depletion occurs in the vicinity of the NH3 molecule (Figure 7a–d). The Hirshfeld charge analysis indicates the NH3 acts as an electron donor with a large charge transfer from NH3 to N-, B-, Al-, and Si-doped PG of 0.465 e, 0.406 e, 0.312 e, and 0.354 e, respectively. Those results also demonstrate that NH3 is chemically adsorbed on N-, B-, Al-, and Si-doped PG. The CDD plot of H2S@B-PG shows similar results as displayed in Figure 7e. However, for the SO2@B-PG and SO2@Al-PG systems, the charges are mainly accumulated around the SO2 molecule, while the charges are depleted near the B and Al dopants, thus the SO2 acts as an electron acceptor and obtains the charge from B-PG and Al-PG by 0.105 e- and 0.203 e (see Table 3). Considering the large adsorption strength and charge transfer, the SO2 adsorbed on the B- and Al-doped PG also belongs to chemisorption.
In order to further understand the micro-interactions of the gas-substrate, we calculated the DOSs of the N (S) atom of NH3 (H2S/SO2) and its nearest neighbor dopant atoms for various adsorption systems, as displayed in Figure 8. For the NH3@N-PG system, the NH3 is adsorbed on top of its nearest neighbor C atom, and the obvious orbital hybridizations between C-2p and N-2p of NH3 can be observed in the range of −12.5 eV to −10 eV, while the relatively weak interactions between C-sp and N-2s are found at −21.05 eV as shown in Figure 8a. For the NH3@B-PG, NH3@Al-PG, and NH3@Si-PG systems (Figure 8b–d), there are significant hybridizations between N-2p of NH3 and B-, Al-, and Si-sp orbitals ranging from −10.0 eV to −5.0 eV. Moreover, there are also weak hybridizations of N-2s and B-, Al-, and Si-2s around −20.0 eV. In summary, the strong interactions between the N atom of NH3 and its nearest neighbor atoms result in NH3 chemisorption. As for the H2S@B-PG and H2S@Al-PG, the strong adsorption strength of the H2S molecule is primarily contributed by the orbital interactions between S-2p of H2S and B- and Al-sp in the range of −10.0 eV to 0.0 eV (Figure 8e–f). By contrast, the strong orbital hybridizations between S-2p of SO2 and B- and Al-sp occur in the larger range of −10.0 eV to 2.5 eV for the SO2@B-PG and SO2@Al-PG systems, which leads to the SO2 molecule being chemically adsorbed over the B- and Al-doped PG.

3.4. Recovery of Doped PG after Sensing Toxic Gases

The short recovery time at room temperature is also of great importance for a good sensing material. Consequently, the feasibility of the toxic gases (NH3, H2S, and SO2) desorbed from the N-, B-, Si-, and Al-doped PG was explored in this section. As displayed in Table 3, the adsorption energies of NH3, H2S, and SO2 on the doped PG with chemisorption are all lower than −0.97 eV. As known, the gas molecule can be released from a solid material when the corresponding Eads is higher than −0.50 eV [36]. Accordingly, those toxic gases cannot be desorbed immediately after adsorption, and the recovery time becomes very long at room temperature. Recently, the introduction of a negative charge could effectively modulate the adsorption strength of a gas molecule on the substrate [37,38,39]. In this regard, the adsorption energies of NH3, H2S, and SO2 on the doped PG with a different negative charge were calculated, and the results are shown in Figure 9. With the increase in injected negative charge, the adsorption strength of NH3 on B-, Al-, and Si-doped PG is gradually decreased, while that of H2S and SO2 on B- and Al-doped PG is gradually increased. This indicates that the H2S or SO2 removal from the B-PG and Al-PG surface becomes more difficult after the introduction of electrons due to the stronger adsorption strength. Nevertheless, the adsorption energies of NH3@B-PG and NH3@Si-PG are increased from −2.46 eV to −0.38 eV and from −1.11 eV to −0.29 eV after adding 2.0 e into the systems. For the NH3@N-PG system, the adsorption energy is also remarkably changed from −1.06 eV to −0.24 eV (−0.35 eV) after injecting 0.5 e (1.0 e) into the N-PG substrate. Thus, we can conclude that the chemically adsorbed NH3 molecule on the N-, B-, and Si-doped PG could be easily desorbed by controlling the number of injected electrons into the systems. Considering the N-, B-, and Si-doped PG also exhibit good selectivity and sensitivity toward the NH3 molecule as discussed above, we can therefore say that the N-, B-, and Si-doped PG could be a promising sensing material for NH3 detection.

4. Conclusions

In summary, the adsorption behaviors and gas sensing properties of the pristine and N-, B-, P-, Al-, and Si-doped PG toward three toxics gases (NH3, H2S, and SO2) were studied by DFT calculations in this work. In addition, the adsorption characteristics, electronic properties, and the sensing mechanism were also discussed. The results of formation energy show that the substitution of the C1 atom by N, B, P, Al, or Si is energetically favorable. The pristine PG exhibits poor sensing ability of the NH3, H2S, and SO2 molecules due to weak adsorption strength. In contrast, doping B, N, Al, or Si atoms into PG nanosheets can obviously improve the NH3 adsorption strength, while only the doping of the B and Al atoms enhance the interactions between the H2S and SO2 molecules and PG, which was further verified by the analysis of DOSs, CDD, and the charge transfer. Additionally, the adsorption strength of NH3 on B-, N-, and Si-doped PG is much larger than that of H2S and SO2 on the same substrates, demonstrating that the B-, N-, and Si-doped PG have good NH3 selectivity from the mixtures of H2S and SO2. In addition, doped PG exhibits high NH3 sensitivity due to its great variation of resistivity after NH3 adsorption. More interestingly, the chemically adsorbed NH3 molecule on the N-, B-, and Si-doped PG could be effectively desorbed by controlling the number of injected electrons into the systems. Therefore, the N-, B-, and Si-doped PG can be the potential materials to develop NH3 sensors with high sensitivity and selectivity.

Author Contributions

G.C.: Conceptualization, validation, formal analysis, and visualization; H.Z.: methodology, writing—review and editing, and project administration; H.X.: software, investigation, supervision, writing—review and editing, and project administration; L.G.: data curation, writing—original draft preparation, and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant number 52161001, 52074135), the Key project of Natural Science Foundation of Jiangxi Province (20202ACBL214012), the Postdoctoral Research Foundation of China (2020M682115), and independent project of Jiangxi Advanced Copper Industry Research Institute (grant number ZL-202012).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data available upon request.

Acknowledgments

We thank Jiangxi Advanced Copper Industry Research Institute for the use of the supercomputing center.

Conflicts of Interest

The authors declare no conflict of interest. The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Zhang, Y.; Zhang, J.; Jiang, Y.; Duan, Z.; Liu, B.; Zhao, Q.; Wang, S.; Yuan, Z.; Tai, H. Ultrasensitive flexible NH3 gas sensor based on polyaniline/SrGe4O9 nanocomposite with ppt-level detection ability at room temperature. Sens. Actuators B Chem. 2020, 319, 128293. [Google Scholar] [CrossRef]
  2. Deng, Z.; Meng, G.; Fang, X.; Dong, W.; Shao, J.; Wang, S.; Tong, B. A novel ammonia gas sensors based on p-type delafossite AgAlO2. J. Alloy. Compd. 2019, 777, 52–58. [Google Scholar] [CrossRef]
  3. Xiao, B.; Li, Y.-C.; Yu, X.-F.; Cheng, J.-B. MXenes: Reusable materials for NH3 sensor or capturer by controlling the charge injection. Sens. Actuators B Chem. 2016, 235, 103–109. [Google Scholar] [CrossRef]
  4. Khakbaz, P.; Moshayedi, M.; Hajian, S.; Soleimani, M.; Narakathu, B.B.; Bazuin, B.J.; Pourfath, M.; Atashbar, M.Z. Titanium car-bide MXene as NH3 sensor: Realistic first-principles study. J. Phys. Chem. C 2019, 123, 29794–29803. [Google Scholar] [CrossRef]
  5. Xu, Y.; Meng, R.; Xiong, D.; Xiang, S.; Wang, S.; Jing, X.; Chen, X. Monolayer ZnS as a promising candidate for NH3 sensor: A first-principle study. IEEE Sens. J. 2017, 17, 6515. [Google Scholar] [CrossRef]
  6. Xiong, H.; Liu, B.; Zhang, H.; Qin, J. Theoretical insight into two-dimensional M-Pc monolayer as an excellent material for formaldehyde and phosgene sensing. Appl. Surf. Sci. 2021, 543, 148805. [Google Scholar] [CrossRef]
  7. Wang, D.; Lan, T.; Pan, J.; Liu, Z.; Yang, A.; Yang, M.; Chu, J.; Yuan, H.; Wang, X.; Li, Y.; et al. Janus MoSSe monolayer: A highly strain-sensitive gas sensing material to detect SF6 decompositions. Sens. Actuators A Phys. 2020, 311, 112049. [Google Scholar] [CrossRef]
  8. Salih, E.; Ayesh, A.I. First principle investigation of H2Se, H2Te and PH3 sensing based on graphene oxide. Phys. Lett. A 2020, 384, 126775. [Google Scholar] [CrossRef]
  9. Jia, N.; Liu, L.; Wang, C.; Guo, G.; Wang, R. Adsorption of gas molecules on 2D Na3Bi monolayer: A first-principles study. Phys. Lett. A 2021, 399, 127280. [Google Scholar] [CrossRef]
  10. Hong, L.X.; Shan, L.S.; Liang, Y.Y.; Zhou, Z.R. Adsorption of NH3 onto vacancy-defected Ti2CO2 monolayer by first-principles calculations. Appl. Surf. Sci. 2020, 504, 144325. [Google Scholar] [CrossRef]
  11. Huang, C.-S.; Murat, A.; Babar, V.; Montes, E.; Schwingenschlögl, U. Adsorption of the gas molecules NH3, NO, NO2, and CO on borophene. J. Phys. Chem. C 2018, 122, 14665–14670. [Google Scholar] [CrossRef]
  12. Yong, Y.; Cui, H.; Zhou, Q.; Su, X.; Kuang, Y.; Li, X. C2N monolayer as NH3 and NO sensors: A DFT study. Appl. Surf. Sci. 2019, 487, 488–495. [Google Scholar] [CrossRef]
  13. Zhang, J.; Yang, G.; Yuan, D.; Tian, J.; Ma, D. A first-principles study of doped black phosphorus carbide monolayers as NO2 and NH3 sensors. J. Appl. Phys. 2019, 125, 074501. [Google Scholar] [CrossRef]
  14. Cui, H.; Zhang, X.; Li, Y.; Chen, D.; Zhang, Y. First-principles insight into Ni-doped InN monolayer as a noxious gases scavenger. Appl. Surf. Sci. 2019, 494, 859–866. [Google Scholar] [CrossRef]
  15. Zhang, S.; Zhou, J.; Wang, Q.; Chen, X.; Kawazoe, Y.; Jena, P. Penta-graphene: A new carbon allotrope. Proc. Natl. Acad. Sci. USA 2015, 112, 2372–2377. [Google Scholar] [CrossRef] [Green Version]
  16. Sathishkumar, N.; Wu, S.Y.; Chen, H.T. Charge-modulated/electric-field controlled reversible CO2/H2 capture and storage on metal-free N-doped penta-graphene. Chem. Eng. J. 2019, 391, 123577. [Google Scholar] [CrossRef]
  17. Chen, D.; Zhang, X.; Tang, J.; Cui, H.; Chen, Z.; Li, Y. Different doping of penta-graphene as adsorbent and gas sensing material for scavenging and detecting SF6 decomposed species. Sustain. Mater. Technol. 2019, 21, e00100. [Google Scholar] [CrossRef]
  18. Qin, H.; Feng, C.; Luan, X.; Yang, D. First-principles investigation of adsorption behaviors of small molecules on penta-graphene. Nanoscale Res. Lett. 2018, 13, 264. [Google Scholar] [CrossRef] [PubMed]
  19. Sathishkumar, N.; Wu, S.Y.; Chen, H.T. Boron- and nitrogen-doped penta-graphene as a promising material for hydrogen storage: A computational study. Int. J. Energy Res. 2019, 43, 4867–4878. [Google Scholar] [CrossRef]
  20. Xiao, B.; Li, Y.-C.; Yu, X.-F.; Cheng, J.-B. Penta-graphene: A promising anode material as the Li/Na-ion battery with both extremely high theoretical capacity and fast charge/discharge rate. ACS Appl. Mater. Interfaces 2016, 8, 35342–35352. [Google Scholar] [CrossRef]
  21. Dai, X.S.; Shen, T.; Feng, Y.; Liu, H.C. Structure, electronic and optical properties of Al, Si, P doped penta-graphene: A first-principles study. Phys. B Condens. Matter 2019, 574, 411660. [Google Scholar] [CrossRef]
  22. Wang, J.; Wang, Z.; Zhang, R.J.; Zheng, Y.X.; Chen, L.Y.; Wang, S.Y.; Tsoo, C.C.; Huang, H.J.; Su, W.S. A first-principles study of the electrically tunable band gap in few-layer penta-graphene. Phys. Chem. Chem. Phys. 2018, 20, 18110–18116. [Google Scholar] [CrossRef] [PubMed]
  23. Cheng, M.-Q.; Chen, Q.; Yang, K.; Huang, W.-Q.; Hu, W.-Y.; Huang, G.-F. Penta-graphene as a potential gas sensor for NOx detection. Nanoscale Res. Lett. 2019, 14, 306. [Google Scholar] [CrossRef] [Green Version]
  24. Tang, S.; Xu, L.; Peng, B.; Xiong, F.; Chen, T.; Luo, X.; Huang, X.; Li, H.; Zeng, J.; Ma, Z.; et al. Ga-doped Pd/CeO2 model catalysts for CO oxidation reactivity: A density functional theory study. Appl. Surf. Sci. 2021, 575, 151655. [Google Scholar] [CrossRef]
  25. Xiong, H.; Xie, J.; Dong, J. Insight into rare earth yttrium and nitrogen co-decorated graphene as a promising material for NOx detection. Phys. Lett. A 2020, 384, 126910–126918. [Google Scholar] [CrossRef]
  26. Zhang, C.-P.; Li, B.; Shao, Z.-G. First-principle investigation of CO and CO2 adsorption on Fe-doped penta-graphene. Appl. Surf. Sci. 2019, 469, 641–646. [Google Scholar] [CrossRef]
  27. Habibpour, R.; Ahmadi, A.; Faghihnasiri, M.; Amani, P. A comparative investigation of penta-graphene and Pt single atom@penta-graphene in H2 and O2 detection: DFT study with assessment of the van der Waals density functionals. Appl. Surf. Sci. 2020, 528, 147043–147055. [Google Scholar] [CrossRef]
  28. Delley, B. From molecules to solids with the DMol3 approach. J. Chem. Phys. 2000, 113, 7756–7764. [Google Scholar] [CrossRef]
  29. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Delley, B. Hardness conserving semilocal pseudopotentials. Phys. Rev. B 2002, 66, 155125. [Google Scholar] [CrossRef]
  31. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef] [PubMed]
  32. Vozzi, C.; Negro, M.; Calegari, F.; Sansone, G.; Nisoli, M.; De Silvestri, S.; Stagira, S. Generalized molecular orbital tomography. Nat. Phys. 2011, 7, 822–826. [Google Scholar] [CrossRef]
  33. Yang, S.; Zhang, B.; Zheng, X.; Chen, G.; Ju, Y.; Chen, B.Z. Insight into the adsorption mechanisms of CH4, CO2, and H2O mol-ecules on illite (001) surfaces: A first-principles study. Surf. Interfaces 2021, 23, 101039–101047. [Google Scholar] [CrossRef]
  34. Song, M.; Chen, Y.; Liu, X.; Xu, W.; Zhao, Y.; Zhang, M.; Zhang, C. A first-principles study of gas molecule adsorption on hydrogen-substituted graphdiyne. Phys. Lett. A 2020, 384, 126332–126338. [Google Scholar] [CrossRef]
  35. Zhao, K.; Guo, Y.; Wang, Q. Contact properties of a vdW heterostructure composed of penta-graphene and penta-BN2 sheets. J. Appl. Phys. 2018, 124, 165103. [Google Scholar] [CrossRef] [Green Version]
  36. Guo, H.; Zhang, W.; Lu, N.; Zhuo, Z.; Zeng, X.C.; Wu, X.; Yang, J. CO2 Capture on h-BN Sheet with High Selectivity Controlled by External Electric Field. J. Phys. Chem. C 2015, 119, 6912–6917. [Google Scholar] [CrossRef]
  37. Sun, Q.; Li, Z.; Searles, D.J.; Chen, Y.; Lu, G. (Max); Du, A. charge-controlled switchable CO2 Capture on boron nitride nanomaterials. J. Am. Chem. Soc. 2013, 135, 8246–8253. [Google Scholar] [CrossRef]
  38. Bruno, G.; Losurdo, M.; Kim, T.-H.; Brown, A. Adsorption and desorption kinetics of Ga on GaN(0001): Application of Wolkenstein theory. Phys. Rev. B 2010, 82, 075326. [Google Scholar] [CrossRef] [Green Version]
  39. Wolkenstein, T. Electronic Processes on Semiconductor Surfaces during Chemisorption; Springer: Boston, MA, USA, 1991; pp. 83–124. [Google Scholar]
Figure 1. Top view (a), side view (b), and band structure (c) of pristine penta-graphene after full relaxation. The blue dashed square is the primitive cell, and the symbols of A∼F represent the possible adsorption sites.
Figure 1. Top view (a), side view (b), and band structure (c) of pristine penta-graphene after full relaxation. The blue dashed square is the primitive cell, and the symbols of A∼F represent the possible adsorption sites.
Membranes 12 00077 g001
Figure 2. The lowest-energy structures of gas molecules adsorbed on pristine PG: (a) NH3, (b) H2S, and (c) SO2. The light blue, purple, dark blue, yellow, and red balls are C, H, N, S, and O atoms, respectively.
Figure 2. The lowest-energy structures of gas molecules adsorbed on pristine PG: (a) NH3, (b) H2S, and (c) SO2. The light blue, purple, dark blue, yellow, and red balls are C, H, N, S, and O atoms, respectively.
Membranes 12 00077 g002
Figure 3. Density of states of NH3, H2S, and SO2 adsorbed on the pristine PG. The dashed line is the Fermi level.
Figure 3. Density of states of NH3, H2S, and SO2 adsorbed on the pristine PG. The dashed line is the Fermi level.
Membranes 12 00077 g003
Figure 4. Charge density difference of different adsorption systems (a) NH3@PG, (b) H2S@PG, and (c) SO2@PG. The green (red) region is electron accumulation (depletion), and the isosurface is ±0.01 e/Å−3.
Figure 4. Charge density difference of different adsorption systems (a) NH3@PG, (b) H2S@PG, and (c) SO2@PG. The green (red) region is electron accumulation (depletion), and the isosurface is ±0.01 e/Å−3.
Membranes 12 00077 g004
Figure 5. Fully-relaxed structures of different adsorption systems. NH3 adsorbed on N-PG (a), B-PG (b), Al-PG (c), and Si-PG (d); H2S adsorbed on B-PG (e) and Al-PG (f); SO2 adsorbed on B-PG (g) and Al-PG (h).
Figure 5. Fully-relaxed structures of different adsorption systems. NH3 adsorbed on N-PG (a), B-PG (b), Al-PG (c), and Si-PG (d); H2S adsorbed on B-PG (e) and Al-PG (f); SO2 adsorbed on B-PG (g) and Al-PG (h).
Membranes 12 00077 g005
Figure 6. Band structures of the N-, B-, Al-, and Si-doped PG without gas adsorptions (ad) and with NH3, H2S, and SO2 adsorptions (el). The dashed lines are the Fermi level.
Figure 6. Band structures of the N-, B-, Al-, and Si-doped PG without gas adsorptions (ad) and with NH3, H2S, and SO2 adsorptions (el). The dashed lines are the Fermi level.
Membranes 12 00077 g006
Figure 7. Difference charge density of different adsorption systems (a) NH3@N-PG, (b) NH3@B-PG, (c) NH3@Al-PG, (d) NH3@Si-PG, (e) H2S@B-PG, (f) H2S@Al-PG, (g) SO2@B-PG, and (h) SO2@Al-PG. The green (red) region is electron accumulation (depletion), and the isosurface is ±0.02 e/Å−3.
Figure 7. Difference charge density of different adsorption systems (a) NH3@N-PG, (b) NH3@B-PG, (c) NH3@Al-PG, (d) NH3@Si-PG, (e) H2S@B-PG, (f) H2S@Al-PG, (g) SO2@B-PG, and (h) SO2@Al-PG. The green (red) region is electron accumulation (depletion), and the isosurface is ±0.02 e/Å−3.
Membranes 12 00077 g007
Figure 8. The DOSs of the N (S) atom of NH3 (H2S/SO2) and its nearest neighbor dopant atoms for various adsorption systems, (a) NH3@N-PG, (b) NH3@B-PG, (c) NH3@Al-PG, (d) NH3@Si-PG, (e) H2S@B-PG, (f) H2S@Al-PG, (g) SO2@B-PG, and (h) SO2@Al-PG. The dashed line is the Fermi level.
Figure 8. The DOSs of the N (S) atom of NH3 (H2S/SO2) and its nearest neighbor dopant atoms for various adsorption systems, (a) NH3@N-PG, (b) NH3@B-PG, (c) NH3@Al-PG, (d) NH3@Si-PG, (e) H2S@B-PG, (f) H2S@Al-PG, (g) SO2@B-PG, and (h) SO2@Al-PG. The dashed line is the Fermi level.
Membranes 12 00077 g008
Figure 9. Adsorption energies of NH3, H2S, and SO2 over doped PG as a function of the injected negative charges.
Figure 9. Adsorption energies of NH3, H2S, and SO2 over doped PG as a function of the injected negative charges.
Membranes 12 00077 g009
Table 1. The formation energy of penta-graphene with different dopants.
Table 1. The formation energy of penta-graphene with different dopants.
StructuresEfor/eV
sp2 Hybridized Carbon Atom (C1 Site)sp3 Hybridized Carbon Atom (C2 Site)
B-doped PG+0.92 (+0.67) [17]+1.00 (+0.74) [17]
N-doped PG−0.10 (−0.01) [17]+1.80 (+1.90) [17]
P-doped PG−0.14+0.95
Si-doped PG+0.87+0.96
Al-doped PG+1.51+3.23
Table 2. Adsorption energy (Eads), the Hirshfeld charge transferred from PG to gases (QT), adsorption distance (d), and band gap (Eg) of different gases adsorbed on pristine PG.
Table 2. Adsorption energy (Eads), the Hirshfeld charge transferred from PG to gases (QT), adsorption distance (d), and band gap (Eg) of different gases adsorbed on pristine PG.
Gas MoleculesEads/eVQT/ed/ÅEg/eV
NH3−0.3820.0582.970 (N-C)2.35
H2S−0.2980.0403.169 (S-C)2.22
SO2−0.328−0.0722.661 (S-C)1.90
Table 3. Adsorption energy (Eads) of various adsorption systems, Hirshfeld charge transferred from gases to doped PG (QT), and band gap of the system before adsorption (Eg) and after adsorption (Eg’). The positive (negative) QT means that gases lose (gain) charge.
Table 3. Adsorption energy (Eads) of various adsorption systems, Hirshfeld charge transferred from gases to doped PG (QT), and band gap of the system before adsorption (Eg) and after adsorption (Eg’). The positive (negative) QT means that gases lose (gain) charge.
StructuresNH3H2SSO2
EadsQTEg (Eg’)EadsQTEg (Eg’)EadsQTEg (Eg’)
B-doped PG−2.460.4061.81/1.23−1.630.3601.81/1.31−1.13−0.1051.81/1.82
N-doped PG−1.060.4651.88/2.27−0.200.0091.88/1.65−0.42−0.1331.88/0.00
P-doped PG−0.570.1470.00/0.00 −0.400.0670.00/0.00−0.29−0.0430.00/1.33
Al-doped PG−2.170.3120.54/0.76 −2.87−0.0430.54/0.42−0.97−0.2030.54/0.46
Si-doped PG−1.110.3541.32/0.84−0.410.0731.32/1.20−0.35−0.1681.32/1.10
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, G.; Gan, L.; Xiong, H.; Zhang, H. Density Functional Theory Study of B, N, and Si Doped Penta-Graphene as the Potential Gas Sensors for NH3 Detection. Membranes 2022, 12, 77. https://doi.org/10.3390/membranes12010077

AMA Style

Chen G, Gan L, Xiong H, Zhang H. Density Functional Theory Study of B, N, and Si Doped Penta-Graphene as the Potential Gas Sensors for NH3 Detection. Membranes. 2022; 12(1):77. https://doi.org/10.3390/membranes12010077

Chicago/Turabian Style

Chen, Guangjun, Lei Gan, Huihui Xiong, and Haihui Zhang. 2022. "Density Functional Theory Study of B, N, and Si Doped Penta-Graphene as the Potential Gas Sensors for NH3 Detection" Membranes 12, no. 1: 77. https://doi.org/10.3390/membranes12010077

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