Abstract

The environmental stability of 2D monolayers is critical for their applications across various technology-related fields. These monolayers can degrade when exposed to gaseous components in the environment, so minimizing these degrading effects is essential. In this paper, chlorine exposure to the 2D monolayers, specifically graphene, silicene, phosphorene, and h-BN monolayer, is investigated using van der Waals corrected density functional theory. The results find that atomic chlorine chemisorbs on graphene, h-BN, silicene, and phosphorene with adsorption energies of −1.09, −0.65, −3.10, and −1.74 eV/atom, and bond distances of 3.0, 2.6, 2.2, and 2.1 Å, respectively. In contrast, molecular Cl2 exhibits physisorption with adsorption energies around −0.22 eV and bond distances ranging from 3.3 to 3.6 Å. NEB calculations show that Cl2 dissociative chemisorption is exothermic on buckled monolayers (silicene and phosphorene) and endothermic on planar monolayers (graphene and h-BN). On buckled surfaces, Cl2 dissociates after overcoming energy barriers of 2.0 eV for silicene and 3.2 eV for phosphorene, forming a stable chemisorbed state that is 0.9 eV lower than the physisorbed state. However, on planar monolayers, Cl2 remains in the physisorbed state because the dissociated chemisorbed state is ≈ 1.5 eV higher in energy. These differences are due to the weaker π-bonds in buckled monolayers, which make dissociation easier, while planar monolayers stabilize the molecular form.
1. Introduction
Synthesis and characterization of two-dimensional (2D) materials, comprising single or few layers of atoms bonded by covalent bonds within each layer and weak van der Waals forces between layers, has opened new avenues in technology at the nanoscale.1−4 For example, graphene shows exceptional mechanical strength, electrical conductivity, and thermal properties, while the h-BN monolayer offers a substrate surface ideal for electronic devices. On the other hand, buckled silicene and phosphorene offer alternatives to graphene for applications in electronic devices. Before these novel 2D materials can be reliably used in devices, their environmental stability needs to be investigated5−8 in terms of how well they respond when exposed to various environmental gaseous elements such as oxygen and ozone.9,10
Chlorine is a highly reactive that can cause significant degradation by bonding with surface atoms, leading to chemical changes and compromising material property. Molecular chlorine (Cl2) is a noncombustible gas moderately soluble at ambient temperature and atmospheric pressure.11,12 Due to its higher density than air, it tends to settle in low-lying areas, providing microenvironmental exposure to technological devices. Some of the previous theoretical and experimental studies have studied the chlorination of graphene and phosphorene to study their feasibility for various applications.13,14 Chlorine also plays a significant role in the semiconductor industry, making it an essential component in various fabrication and processing steps such as dry etching15,16 wafer cleaning,17 chemical vapor deposition18,19 passivation,20 etc. Hence, an atomic-level understanding of the interaction of atomic chlorine (Cl) or Cl2 with 2D monolayers is essential. However, most of the studies focus on the degradation of the 2D materials due to oxidizing agents such as oxygen, ozone and moisture.10,21−23 This gap is now addressed in this paper, in which we present a comprehensive analysis of the interaction between Cl and Cl2 with graphene, h-BN monolayer, silicene, and phosphorene using density functional theory (DFT).
Graphene and silicene, despite both being composed of elements from group IV of the periodic table, differ significantly in their chemical bonding; the former has a strong sp2 hybridization leading to the planar configuration, whereas the latter has a weaker sp2 hybridization, resulting in a buckled (nonplanar) configuration. On the other hand, the bonding in the h-BN monolayer exhibits some polar covalent character, while phosphorene has a lone pair of electrons, resulting in sp3 hybridization in its puckered structure. Thus, this difference in chemical bonding is expected to influence the nature of Cl or Cl2 adsorption on the planar or buckled (puckered) monolayers. Nudged elastic band (NEB) calculations will also be performed to determine if an energy barrier exists during the Cl or Cl2 adsorption process on these monolayers.
2. Computational Method
Spin-polarized DFT Calculations were performed using the Vienna Ab initio Simulation Package (VASP).24,25 The Perdew Burke Ernzerhof (PBE) form of the generalized gradient approximations26 was employed to describe the exchange–correlation interaction, while the interactions between valence electrons and ion-core were accounted for using the projected augmented wave (PAW) method.27 The van der Waals (vdW) correction D3 term was utilized to capture nonlocal dispersive interactions in the chlorine-conjugated monolayers.28
A (4 × 4 × 1) supercell containing a Cl or Cl2 interacting with a pristine monolayer was used for the periodic DFT calculations. The supercell was fully relaxed until convergence criteria were met. A plane wave energy cutoff of 520 eV was set, with convergence criteria of 1 × 10–5 eV for energy and 0.01 eV/Å for force. A 16 Å vacuum along the z-axis direction was used to minimize interactions between periodic images. The Monkhorst–Pack scheme29 was employed for k–point sampling in the Brillouin zone, with a (3 × 3 × 1) k–point mesh used for the structure optimization. A larger (12 × 12 × 1) k–point mesh was used for density of states (DOS) calculations. The atomic charges of the interacting complexes were determined using Bader’s atoms-in-molecules approach.30
To gain a comprehensive understanding of the chlorination process, the NEB method31 was used to determine the energy pathway, representing the interaction between Cl or Cl2 and the monolayers. Five intermediate configurations between predefined initial and final states were considered. In the initial configuration, Cl2 is placed above a noninteracting distance of 4.5 Å from the monolayer, while the final configuration consists of atomic Cl adsorbed onto neighboring host atoms of the monolayer.32 In this way, the final configuration represents the dissociated chemisorption process of the molecular Cl2.
3. Results and Discussions
3.1. Chlorine
Atomic chlorine has an electronic configuration of [Ne] 3s2, 3p5 with a 2P3/2 electronic state. It is not stable due to its high electronegativity. Molecular chlorine in the ground state is linear, having a1Σg+ state with a bond distance of 1.988 Å, obtained by spectroscopic experiments.33,34 In the present case, the calculated bond distance is 2.0 Å, and the dissociation energy is 2.82 eV at the DFT (PBE) level of theory.
3.2. Pristine Monolayers
Graphene, h-BN monolayer, silicene and phosphorene are considered representative 2D monolayers in the present study. Graphene consists of a single layer of carbon atoms arranged in a planar sp2-hybridized honeycomb lattice. It has a lattice parameter of 2.46 Å and a (C–C) bond length of 1.42 Å.35,36 The h-BN monolayer, despite having a hexagonal lattice-like graphene, consists of alternating boron and nitrogen atoms, resulting in semi-ionic B–N bonds. Its lattice constant is 2.50 Å, and the (B–N) bond length of 1.42 Å.37,38 The silicon analog of graphene, silicene, has a honeycomb lattice, but due to sp3 hybridization, it forms a slightly buckled structure. This results in a sublattice separation of about 0.44 Å, a lattice parameter of 3.86 Å, and the (Si–Si) bond length of 2.25 Å.39 Black phosphorus, in its single-layer form, known as phosphorene, has a puckered hexagonal anisotropic configuration with lattice constants of 3.30 and 4.60 Å, and the (P–P) bond distance is 2.21 Å.40 The calculated lattice constants and the band gap for the monolayers are listed in Table 1, and the equilibrium configurations are shown in Figure S1 (Supporting Information). The calculated results are consistent with the previously reported DFT calculations.36,39,40 For example, the calculated lattice parameters and bandgap value for pristine silicene (a = 3.86 Å, Egap ≈0 eV) and phosphorene (a = 4.60 Å, b = 3.30 Å, Egap = 0.81 eV) are in good agreement with the previously reported values.39,40
Table 1. Calculated Structural Parameters Lattice Constants (a and b), the Nearest-Neighbor Distance (RX-X) and Bandgaps (Egap) for the Monolayers.
| monolayers | a = b (Å) | RX-X (Å) | Egap (eV) |
|---|---|---|---|
| graphene (D6h) | 2.46 | 1.42 | ∼0 |
| h-BN monolayer (D3h) | 2.51 | 1.44 | 4.32 |
| silicene (D3d) | 3.86 | 2.27 | ∼0 |
| phosphorene (D2h) | a = 4.60, b = 3.30 | 2.21 | 0.81 |
3.3. Cl/Cl2 Adsorbed Monolayers
Three different sites, top, hollow, and bridge (Figure S2, Table S1 Supporting Information), on a given monolayer, were considered to determine the energetically preferred adsorption site for Cl and Cl2. Physisorption involves weak van der Waals interactions with low adsorption energy (less than −0.5 eV), larger molecule–substrate distances (greater than 3.0 Å), and minimal charge transfer. In contrast, chemisorption is characterized by strong chemical bonding, higher adsorption energy (greater than −0.5 eV), shorter bond distances (less than 3.0 Å), and significant charge transfer.41,42
The results, displayed in Figure 1, indicate that Cl prefers to be adsorbed on the top site of the monolayers; the calculated bond distance is 3.0, 2.6, 2.1, and 2.2 Å for graphene, h-BN monolayer, silicene and phosphorene, respectively.
Figure 1.
Top and side views of the equilibrium configurations of (a) Cl and (b) Cl2 adsorbed on graphene, h-BN monolayer, phosphorene, and silicene. Color codes: red: Cl, yellow: C, blue: B, gray: N, black: P, and brown: Si.
It should be noted that on the h-BN monolayer, chlorine preferentially adsorbs on the top site of the nitrogen atom. However, this is not the case for Cl2, which indicates different preferences; Cl2, oriented parallel to the surface, prefers the top site for graphene and h-BN monolayer, while it prefers bridge site for silicene and phosphorene. The calculated bond distance is 3.6, 3.6, 3.4, and 3.3 Å for graphene, h-BN monolayer, silicene and phosphorene, respectively. The structural and electronic properties of the adsorbed complex configurations are listed in Table 2. Table 2 lists the adsorption energy (ΔEads), molecule–surface distance (RX-Cl), and charge transfer from the surface to the molecule (ΔQ) for the adsorbed monolayer complexes. The adsorption energy, ΔEad, was calculated using the equation
| 1 |
where ECl/Cl2+monolayer is the total energy of the adsorbed configuration, Emonolayer is the total energy of the pristine monolayer, and ECl/Cl2 is the total energy of the Cl or Cl2.
Table 2. Calculated Adsorption Energy (ΔEad), Molecule–Surface Distance (RX-Cl), and Charge Transfer from the Surface to Chlorine (ΔQ) of Graphene, h-BN Monolayer, Silicene and Phosphorene.
| monolayer | graphene | h-BN monolayer | silicene | phosphorene | ||
|---|---|---|---|---|---|---|
| Cl | chemisorption | ΔEad (eV/atom) | –1.09 | –0.65 | –3.10 | –1.74 |
| RX–Cl(Å) | 3.00 | 2.60 | 2.10 | 2.20 | ||
| ΔQ (e) | –0.41 | –0.24 | –0.59 | –0.29 | ||
| Cl2 | physisorption | ΔEad (eV/atom) | –0.22 | –0.22 | –0.23 | –0.23 |
| RX–Cl (Å) | 3.60 | 3.60 | 3.40 | 3.30 | ||
| ΔQ (e) | ∼0.0 | 0.02 | –0.06 | ∼0.0 |
For the case of atomic Cl, the calculated ΔEad is −1.09, −0.65, −3.1, and −1.74 eV/atom for graphene, h-BN monolayer, silicene, and phosphorene, respectively. The binding site is at the top of the primary element atom (i.e., C, N, Si, or P) with equilibrium bond distances, RX-Cl of 3.0, 2.6, 2.2, and 2.1 Å, respectively. On the other hand, ΔEad is approximately −0.22 (−0.23) eV for Cl2, with bond distances in the range of 3.3–3.6 Å. Based on the relative differences in ΔEad and the bond distances calculated for Cl and Cl2, it can be concluded that Cl undergoes chemisorption, while Cl2 undergoes physisorption on both planar and buckled monolayers.
Figure 2 shows the charge density difference plots for chemisorbed Cl and physisorbed Cl2 monolayer complexes, displaying that the latter case represents a weakly adsorbed state with a relatively small degree of charge density accumulation/depletion. This is further reaffirmed by the calculated charge transfer for both cases, as listed in Table 2; a noticeable charge transfer is seen in the Cl-adsorbed complexes, and negligible charge transfer occurs in the Cl2-adsorbed complexes. Since the alignment of the Fermi levels of the monolayers with the lowest unoccupied molecular orbitals (LUMO) of interacting molecules determines the direction and feasibility of electron transfer, we plot the Fermi levels and LUMOs in Figure S3 (Supporting Information). The results consistently predict a higher degree of charge transfer for atomic Cl as compared to molecular Cl2 in the adsorbed monolayer complexes.
Figure 2.
Top and side views of the charge density difference plots: (a) physisorbed Cl2 molecule and (b) dissociated chemisorbed Cl. The cyan and yellow colors indicate the charge depletion and accumulation, respectively. The isosurface value is 1/10th of the maximum isosurface value.
Next, we calculate the minimum energy path followed by Cl2 using the nudged elastic band (NEB) method. This method samples the configuration space, encompassing various intermediate states and their associated energies associated with the energy pathway. For the NEB calculations, the initial (noninteracting) configuration consisted of Cl2 placed at ∼4.5 Å above the monolayer, and the final configuration consisted of the dissociated molecule (Cl2 → Cl + Cl) adsorbed on the surface (Figure S4 Supporting Information). The molecular dissociation energy (ΔEdissociation) is calculated as the difference between the total energy of the noninteracting configuration (i.e., Cl2 about 4.5 Å above the surface) and that of the dissociated configuration, consisting of Cl adsorbed on the neighboring host atom of the surface (e.g., Figure 2b). The calculated molecular dissociation energy is 0.69 and 0.89 eV for silicene and phosphorene, respectively. Note that the chemisorption of atomic chlorine occurs without any energy barrier as it approaches the surface of either planar or buckled monolayers, except in the case of phosphorene (Figure S5 Supporting Information).
Figure 3a,b show the two energy pathways for Cl2 as it approaches the surface and the degradation pathway of silicene. When interacting with the planar structures, graphene, and h-BN monolayer, Cl2 energetically prefers to remain in a molecular physisorbed state. The dissociated chemisorbed state is ≈ 1.5 eV higher than the molecular physisorbed state, suggesting the dissociation to be endothermic. In contrast, for the buckled monolayers, silicene, and phosphorene, Cl2 favors a dissociated chemisorbed state after crossing relatively high energy barriers of 2.0 and 3.2 eV, respectively. The dissociated chemisorbed state is ≈ 0.9 eV lower than the molecular physisorbed state, suggesting the dissociation to be exothermic.
Figure 3.
A schematic diagram showing (a) the minimum energy path of Cl2 as it approaches the monolayers. The reaction coordinate is a set of coordinates along the minimum energy path. Zero of the energy is taken to be the noninteracting system consisting of Cl2 at a distance of 4.5 Å above the monolayer surface and (b) degradation pathway of silicene.
The dissimilar energy pathways shown in Figure 3 by the planar and buckled monolayers are primarily due to the nature of their chemical bonding, e.g., sp2 in graphene vs sp2-like in silicene (or sp3 in phosphorene),43 resulting in the presence of a relatively weaker π bonds in the buckled monolayers. This facilitates molecular dissociation with a high energy barrier on the buckled monolayers due to their enhanced chemical reactivity. Note that the molecular dissociation energy of Cl2 is calculated to be 2.82 eV.
Analysis of the total and partial density of states, as shown in Figure 4, further highlights the modifications in the electronic structure of the adsorbed monolayer configurations. For physisorption, weak van der Waals interactions dominate, leading to low adsorption energies, larger molecule–substrate distances, and negligible charge transfer. In contrast, chemisorption involves stronger bonding due to significant orbital hybridization and charge transfer, resulting in higher adsorption energies and shorter bond distances. As shown in Figure 4, analysis of the total and partial density of states (DOS) highlights these differences. In physisorbed cases, such as graphene and h-BN, molecular states appear near the Fermi level with minimal interaction with the substrate, indicating weak binding. However, in dissociated chemisorbed cases, such as silicene and phosphorene, Cl p-states strongly hybridize with the host-atom states, leading to stronger adsorption and promoting decomposition. The differences among the 2D materials are attributed to their electronic properties. Planar materials like graphene and h-BN, with delocalized π-electrons, show weak interactions with chlorine, while buckled materials like silicene and phosphorene, with reactive surface states, exhibit stronger chemisorption. Additionally, the energy barriers for dissociative chemisorption, as shown by NEB calculations, further explain their behavior.
Figure 4.
Total and atom-projected density of states of (a–d) pristine and (e–h) adsorbed monolayer configurations.
We further note that graphene and h-BN monolayer exhibit stability and resistance toward oxidation at ambient conditions. Enhanced sensitivity of phosphorene to oxygen is confirmed by experiments, which have demonstrated rapid degradation of this material.44 The DFT calculations have predicted the dissociated chemisorption of molecular oxygen only for the buckled monolayers, silicene and phosphorene in contrast to the case of the planar monolayers, graphene and h-BN monolayer.45
Our results are consistent with the previously published results on Cl-adsorbed graphene and phosphorene. For example, the DFT calculations on Cl-absorbed graphene reported ΔEad ranging from −0.8 to −1.05 eV.46−48 Compared to these results, we find ΔEad to be −1.09 eV in the present study. Likewise, Cl interacting with phosphorene was predicted to prefer the top site for adsorption at a distance of 2.28 Å with ΔEad of −1.61 eV.49,50 This is consistent with the values obtained in the present study: ΔEad of −1.74 eV at a distance of 2.20 Å. On the other hand, in a combined experimental and theoretical study of Br2 interacting with substrate-supported graphene or h-BN nanomesh, it was reported that Br2 gets trapped in the valley with a binding energy of −1.2 and −2.0 eV, respectively.51
We note that photochemical techniques have synthesized the chlorinated graphene.46,47 It was reported that edge chlorination under ambient conditions can occur due to the migration of Cl atoms across the surface with minimal energy barriers.48 Our DFT results are consistent with the previous results, as the adsorption of the ionic Cl– (i.e., the chloride ion) on both planar and buckled monolayers was predicted to be thermodynamically unfavorable, with the calculated ΔEad being positive. This is expected since ionic Cl– is highly stable due to its electronic configuration, similar to that of a noble gas argon.
4. Summary
In summary, a systematic and detailed investigation was performed on the interaction of the atomic Cl and molecular Cl2 molecule with the planar (i.e., graphene and h-BN monolayer) and the buckled (i.e., silicene and phosphorene), using the van der Waals corrected density functional theory together with the NEB method. The results find that atomic Cl prefers chemisorbed on top of a host atom with binding energies ranging from −0.65 to −3.10 eV. Further, the chemisorption is accompanied by a noticeable charge transfer from the surface to the atomic Cl. On the other hand, molecular Cl2 interacting with both planar and buckled monolayers undergo physisorption with a binding energy of approximately −0.22 eV. This process is associated with a negligible charge transfer to the molecule. Moreover, the NEB results show that a high energy barrier exists for the (exothermic) dissociated chemisorption of the molecular Cl2 on the buckled monolayers, silicene and phosphorene. This is not the case with the planar monolayers, graphene and h-BN monolayers, where the dissociated chemisorption is found to be endothermic. Overall, both planar and buckled monolayers are predicted to be relatively robust against exposure to either atomic or molecular chlorine, though the degree of robustness is lower for the buckled monolayers, silicene and phosphorene. These results are expected to help assess the environmentally induced degradation of the 2D monolayer-based electronic devices for technological applications.
Acknowledgments
We greatly appreciate the valuable discussions with Dr. S. Gowtham. We acknowledge using the SUPERIOR High-performance computational facility of Michigan Technological University, USA.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.5c01346.
The Supporting Information includes the equilibrium configurations of graphene, h-BN monolayer, silicene, and phosphorene; The adsorption energies of atomic and molecular chlorine on different monolayer binding sites; A schematic diagram illustrates the monolayer’s Fermi level and the lowest unoccupied molecular orbital of Cl/Cl– and Cl2; Additionally, the final configuration of the system shows the dissociated chlorine molecule adsorbed on the surface and the minimum energy path of Cl as it approaches the monolayers (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- Quesnel E.; Roux F.; Emieux F.; Faucherand P.; Kymakis E.; Volonakis G.; Giustino F.; Martín-García B.; Moreels I.; Gürsel S. A.; et al. Graphene-based technologies for energy applications, challenges and perspectives. 2D Mater. 2015, 2 (3), 030204. 10.1088/2053-1583/2/3/030204. [DOI] [Google Scholar]
- Cai X.; Luo Y.; Liu B.; Cheng H.-M. Preparation of 2D material dispersions and their applications. Chem. Soc. Rev. 2018, 47 (16), 6224–6266. 10.1039/C8CS00254A. [DOI] [PubMed] [Google Scholar]
- Glavin N. R.; Rao R.; Varshney V.; Bianco E.; Apte A.; Roy A.; Ringe E.; Ajayan P. M. Emerging applications of elemental 2D materials. Adv. Mater. 2020, 32 (7), 1904302. 10.1002/adma.201904302. [DOI] [PubMed] [Google Scholar]
- Liu C.; Wang L.; Qi J.; Liu K. Designed growth of large-size 2D single crystals. Adv. Mater. 2020, 32 (19), 2000046. 10.1002/adma.202000046. [DOI] [PubMed] [Google Scholar]
- Zhang P.; Wang F.; Yu M.; Zhuang X.; Feng X. Two-dimensional materials for miniaturized energy storage devices: from individual devices to smart integrated systems. Chem. Soc. Rev. 2018, 47 (19), 7426–7451. 10.1039/C8CS00561C. [DOI] [PubMed] [Google Scholar]
- Khan K.; Tareen A. K.; Aslam M.; Wang R.; Zhang Y.; Mahmood A.; Ouyang Z.; Zhang H.; Guo Z. Recent developments in emerging two-dimensional materials and their applications. J. Mater. Chem. C 2020, 8 (2), 387–440. 10.1039/C9TC04187G. [DOI] [Google Scholar]
- Liang S. J.; Cheng B.; Cui X.; Miao F. Van der Waals heterostructures for high-performance device applications: challenges and opportunities. Adv. Mater. 2020, 32 (27), 1903800. 10.1002/adma.201903800. [DOI] [PubMed] [Google Scholar]
- Fan F. R.; Wang R.; Zhang H.; Wu W. Emerging beyond-graphene elemental 2D materials for energy and catalysis applications. Chem. Soc. Rev. 2021, 50 (19), 10983–11031. 10.1039/C9CS00821G. [DOI] [PubMed] [Google Scholar]
- Wang G.; Slough W. J.; Pandey R.; Karna S. P. Degradation of phosphorene in air: understanding at atomic level. 2D Mater. 2016, 3 (2), 025011. 10.1088/2053-1583/3/2/025011. [DOI] [Google Scholar]
- Patra L.; Sachdeva G.; Pandey R.; Karna S. P. Ozonation of group-IV elemental monolayers: A first-principles study. ACS Omega 2021, 6 (30), 19546–19552. 10.1021/acsomega.1c01862. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Evans R. B. Chlorine: state of the art. Lung 2005, 183 (3), 151–167. 10.1007/s00408-004-2530-3. [DOI] [PubMed] [Google Scholar]
- Yadav A. K.; Bracher A.; Doran S. F.; Leustik M.; Squadrito G. L.; Postlethwait E. M.; Matalon S. Mechanisms and modification of chlorine-induced lung injury in animals. Proc. Am. Thorac. Soc. 2010, 7 (4), 278–283. 10.1513/pats.201001-009SM. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ji X.; Wang S.; Zhang E.; Zhang Y.; Ma Z.; Qiu Y. Chlorinated graphene via the photodecomposition of metal chlorides. ACS Sustainable Chem. Eng. 2019, 7 (12), 11024–11034. 10.1021/acssuschemeng.9b02193. [DOI] [Google Scholar]
- Taylor P. D.; Tawfik S. A.; Spencer M. J. Interplay of mechanical and chemical tunability of phosphorene for flexible nanoelectronic applications. J. Phys. Chem. C 2020, 124 (44), 24391–24399. 10.1021/acs.jpcc.0c05775. [DOI] [Google Scholar]
- He T.; Wang Z.; Zhong F.; Fang H.; Wang P.; Hu W. Etching techniques in 2D materials. Adv. Mater. Technol. 2019, 4 (8), 1900064. 10.1002/admt.201900064. [DOI] [Google Scholar]
- Sun H.; Dong J.; Liu F.; Ding F. Etching of two-dimensional materials. Mater. Today 2021, 42, 192–213. 10.1016/j.mattod.2020.09.031. [DOI] [Google Scholar]
- Watanabe S.; Sugino R.; Yamazaki T.; Nara Y.; Ito T. Wafer Cleaning with Photoexcited Chlorine and Thermal Treatment for High-Quality Silicon Epitaxy. Jpn. J. Appl. Phys. 1989, 28 (10R), 2167. 10.1143/JJAP.28.2167. [DOI] [Google Scholar]
- Modtland B. J.; Navarro-Moratalla E.; Ji X.; Baldo M.; Kong J. Monolayer tungsten disulfide (WS2) via chlorine-driven chemical vapor transport. Small 2017, 13 (33), 1701232. 10.1002/smll.201701232. [DOI] [PubMed] [Google Scholar]
- Wilson P. M.; Chin M. L.; Ekuma C. E.; Najmaei S.; Price K. M.; Schiros T.; Dubey M.; Hone J. Chlorine-mediated atomic layer deposition of HfO 2 on graphene. J. Mater. Chem. C 2021, 9 (48), 17437–17443. 10.1039/D1TC03502A. [DOI] [Google Scholar]
- Wu J.; Li M.-H.; Fan J.-T.; Li Z.; Fan X.-H.; Xue D.-J.; Hu J.-S. Regioselective multisite atomic-chlorine passivation enables efficient and stable perovskite solar cells. J. Am. Chem. Soc. 2023, 145 (10), 5872–5879. 10.1021/jacs.2c13307. [DOI] [PubMed] [Google Scholar]
- Zhang H.; Lee J. Y.; Liu H. Ozone decomposition on defective graphene: insights from modeling. J. Phys. Chem. C 2021, 125 (20), 10948–10954. 10.1021/acs.jpcc.1c02062. [DOI] [Google Scholar]
- Savazzi F.; Risplendi F.; Mallia G.; Harrison N. M.; Cicero G. Unravelling some of the structure–property relationships in graphene oxide at low degree of oxidation. J. Phys. Chem. Lett. 2018, 9 (7), 1746–1749. 10.1021/acs.jpclett.8b00421. [DOI] [PubMed] [Google Scholar]
- Peng Q.; Han L.; Lian J.; Wen X.; Liu S.; Chen Z.; Koratkar N.; De S. Mechanical degradation of graphene by epoxidation: insights from first-principles calculations. Phys. Chem. Chem. Phys. 2015, 17 (29), 19484–19490. 10.1039/C5CP02986D. [DOI] [PubMed] [Google Scholar]
- Kresse G.; Furthmüller J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6 (1), 15–50. 10.1016/0927-0256(96)00008-0. [DOI] [Google Scholar]
- Kresse G.; Furthmüller J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54 (16), 11169. 10.1103/PhysRevB.54.11169. [DOI] [PubMed] [Google Scholar]
- Perdew J. P.; Burke K.; Ernzerhof M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77 (18), 3865. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- Blöchl P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50 (24), 17953. 10.1103/PhysRevB.50.17953. [DOI] [PubMed] [Google Scholar]
- Grimme S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27 (15), 1787–1799. 10.1002/jcc.20495. [DOI] [PubMed] [Google Scholar]
- Monkhorst H. J.; Pack J. D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13 (12), 5188. 10.1103/PhysRevB.13.5188. [DOI] [Google Scholar]
- Bader R. F.; Molecules A. I.. A quantum theory; Clarendon: Oxford, UK, 1990. [Google Scholar]
- Henkelman G.; Uberuaga B. P.; Jónsson H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 2000, 113 (22), 9901–9904. 10.1063/1.1329672. [DOI] [Google Scholar]
- Patra L.; Mallick G.; Pandey R.; Karna S. P. Surface stability of WN ultrathin films under O2 and H2O exposure: A first-principles study. Appl. Surf. Sci. 2022, 588, 152940. 10.1016/j.apsusc.2022.152940. [DOI] [Google Scholar]
- Angeli C.; Cimiraglia R.; Malrieu J.-P. On a mixed Møller–Plesset Epstein–Nesbet partition of the Hamiltonian to be used in multireference perturbation configuration interaction. Chem. Phys. Lett. 2000, 317 (3–5), 472–480. 10.1016/S0009-2614(99)01458-X. [DOI] [Google Scholar]
- Herzberg G.Molecular Spectra and molecular structure-Vol I; Read Books Ltd, 2013; Vol. 1. [Google Scholar]
- Castro Neto A. H.; Guinea F.; Peres N. M.; Novoselov K. S.; Geim A. K. The electronic properties of graphene. Rev. Mod. Phys. 2009, 81 (1), 109–162. 10.1103/RevModPhys.81.109. [DOI] [Google Scholar]
- Avouris P. Graphene: electronic and photonic properties and devices. Nano Lett. 2010, 10 (11), 4285–4294. 10.1021/nl102824h. [DOI] [PubMed] [Google Scholar]
- Watanabe K.; Taniguchi T.; Kanda H. Direct-bandgap properties and evidence for ultraviolet lasing of hexagonal boron nitride single crystal. Nat. Mater. 2004, 3 (6), 404–409. 10.1038/nmat1134. [DOI] [PubMed] [Google Scholar]
- Zhong X.; Yap Y. K.; Pandey R.; Karna S. P. First-principles study of strain-induced modulation of energy gaps of graphene/BN and BN bilayers. Phys. Rev. B:Condens. Matter Mater. Phys. 2011, 83 (19), 193403. 10.1103/physrevb.83.193403. [DOI] [Google Scholar]
- Houssa M.; Dimoulas A.; Molle A. Silicene: a review of recent experimental and theoretical investigations. J. Phys.: Condens. Matter 2015, 27 (25), 253002. 10.1088/0953-8984/27/25/253002. [DOI] [PubMed] [Google Scholar]
- Jing Y.; Zhang X.; Zhou Z. Phosphorene: what can we know from computations?. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2016, 6 (1), 5–19. 10.1002/wcms.1234. [DOI] [Google Scholar]
- Tao J.; Tang H.; Patra A.; Bhattarai P.; Perdew J. P. Modeling the physisorption of graphene on metals. Phys. Rev. B 2018, 97 (16), 165403. 10.1103/PhysRevB.97.165403. [DOI] [Google Scholar]
- Lüth H.; Lüth H.. Adsorption on solid surfaces. Solid Surfaces. In Interfaces and Thin Films; Springer, 2015; pp 527–571. 10.1007/978-3-319-10756-1_10. [DOI] [Google Scholar]
- Cahangirov S.; Topsakal M.; Aktürk E.; Şahin H.; Ciraci S. Two-and one-dimensional honeycomb structures of silicon and germanium. Phys. Rev. Lett. 2009, 102 (23), 236804. 10.1103/PhysRevLett.102.236804. [DOI] [PubMed] [Google Scholar]
- Island J. O.; Steele G. A.; Zant H. S. J. v. d.; Castellanos-Gomez A. Environmental instability of few-layer black phosphorus. 2D Mater. 2015, 2 (1), 011002. 10.1088/2053-1583/2/1/011002. [DOI] [Google Scholar]
- Wang G.; Pandey R.; Karna S. P. Physics and chemistry of oxidation of two-dimensional nanomaterials by molecular oxygen. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2017, 7 (1), e1280 10.1002/wcms.1280. [DOI] [Google Scholar]
- Li B.; Zhou L.; Wu D.; Peng H.; Yan K.; Zhou Y.; Liu Z. Photochemical chlorination of graphene. ACS Nano 2011, 5 (7), 5957–5961. 10.1021/nn201731t. [DOI] [PubMed] [Google Scholar]
- Wu J.; Xie L.; Li Y.; Wang H.; Ouyang Y.; Guo J.; Dai H. Controlled chlorine plasma reaction for noninvasive graphene doping. J. Am. Chem. Soc. 2011, 133 (49), 19668–19671. 10.1021/ja2091068. [DOI] [PubMed] [Google Scholar]
- Sahin H.; Ciraci S. Chlorine adsorption on graphene: Chlorographene. J. Phys. Chem. C 2012, 116 (45), 24075–24083. 10.1021/jp307006c. [DOI] [Google Scholar]
- Hassani N.; Yagmurcukardes M.; Peeters F. M.; Neek-Amal M. Chlorinated phosphorene for energy application. Comput. Mater. Sci. 2024, 231, 112625. 10.1016/j.commatsci.2023.112625. [DOI] [Google Scholar]
- Musle V.; Choudhary S. Tuning the optical properties of phosphorene by adsorption of alkali metals and halogens. Opt. Quantum Electron. 2018, 50, 285. 10.1007/s11082-018-1548-3. [DOI] [Google Scholar]
- Freiberger E. M.; Steffen J.; Waleska-Wellnhofer N. J.; Hemauer F.; Schwaab V.; Görling A.; Steinrück H.-P.; Papp C. Bromination of 2D materials. Nanotechnology 2024, 35 (14), 145703. 10.1088/1361-6528/ad1201. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




