Next Article in Journal
Indole Derivatives: Unveiling New Frontiers in Medicinal and Synthetic Organic Chemistry
Previous Article in Journal
Highly Sensitive Voltammetric Method for Quinoline Yellow Determination on Renewable Amalgam Film Electrode
Previous Article in Special Issue
Structures and Stabilities of Carbon Chain Clusters Influenced by Atomic Antimony
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption Features of Tetrahalomethanes (CX4; X = F, Cl, and Br) on β12 Borophene and Pristine Graphene Nanosheets: A Comparative DFT Study

by
Mahmoud A. A. Ibrahim
1,2,*,
Amna H. M. Mahmoud
1,
Nayra A. M. Moussa
1,
Gamal A. H. Mekhemer
1,
Shaban R. M. Sayed
3,
Muhammad Naeem Ahmed
4,
Mohamed K. Abd El-Rahman
5,
Eslam Dabbish
6 and
Tamer Shoeib
6,*
1
Computational Chemistry Laboratory, Chemistry Department, Faculty of Science, Minia University, Minia 61519, Egypt
2
School of Health Sciences, University of KwaZulu-Natal, Westville Campus, Durban 4000, South Africa
3
Department of Botany and Microbiology, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
4
Department of Chemistry, The University of Azad Jammu and Kashmir, Muzaffarabad 13100, Pakistan
5
Department of Chemistry and Chemical Biology, Harvard University, 12 Oxford Street, Cambridge, MA 02138, USA
6
Department of Chemistry, The American University in Cairo, New Cairo 11835, Egypt
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(14), 5476; https://doi.org/10.3390/molecules28145476
Submission received: 21 June 2023 / Revised: 14 July 2023 / Accepted: 14 July 2023 / Published: 18 July 2023
(This article belongs to the Special Issue Computational Chemistry for Material Research)

Abstract

:
The potentiality of the β12 borophene (β12) and pristine graphene (GN) nanosheets to adsorb tetrahalomethanes (CX4; X = F, Cl, and Br) were investigated using density functional theory (DFT) methods. To provide a thorough understanding of the adsorption process, tetrel (XC-X3∙∙∙β12/GN)- and halogen (X3C-X∙∙∙β12/GN)-oriented configurations were characterized at various adsorption sites. According to the energetic manifestations, the adsorption process of the CX4∙∙∙β12/GN complexes within the tetrel-oriented configuration led to more desirable negative adsorption energy (Eads) values than that within the halogen-oriented analogs. Numerically, Eads values of the CBr4∙∙∙Br1@β12 and T@GN complexes within tetrel-/halogen-oriented configurations were −12.33/−8.91 and −10.03/−6.00 kcal/mol, respectively. Frontier molecular orbital (FMO) results exhibited changes in the EHOMO, ELUMO, and Egap values of the pure β12 and GN nanosheets following the adsorption of CX4 molecules. Bader charge transfer findings outlined the electron-donating property for the CX4 molecules after adsorbing on the β12 and GN nanosheets within the two modeled configurations, except the adsorbed CBr4 molecule on the GN sheet within the tetrel-oriented configuration. Following the adsorption process, new bands and peaks were observed in the band structure and density of state (DOS) plots, respectively, with a larger number in the case of the tetrel-oriented configuration than in the halogen-oriented one. According to the solvent effect affirmations, adsorption energies of the CX4∙∙∙β12/GN complexes increased in the presence of a water medium. The results of this study will serve as a focal point for experimentalists to better comprehend the adsorption behavior of β12 and GN nanosheets toward small toxic molecules.

Graphical Abstract

1. Introduction

Two-dimensional (2D) nanomaterials have recently been of universal interest owing to their outstanding chemical and physical properties [1,2,3,4]. As a premier developed 2D material, pristine graphene (GN) was regarded as the most intriguing star in the realm of materials science [5,6,7,8,9,10,11]. GN-based materials were denoted with unique features, including a high specific surface area [12], quantum Hall effect [13], high thermal conductivity [14], and ambipolar electric field effect [15]. Such properties shed light on their vast-ranging applications, like energy storage [16,17], drug delivery [18,19,20,21], spintronics [22], and catalysis [23]. Because of their low electronic noise, GN-based materials were also announced as an appealing candidate for adsorbing chemical systems [6,24].
Following the astonishing discovery of GN sheets, considerable research has been directed to develop various 2D materials, including antimonene [25], hexagonal boron nitride (h-BN) [26], bismuthine [27], silicene [28,29], and borophene [30,31]. In the parallel area, borophene, a 2D boron sheet, was announced with extraordinary properties, like electron mobility, anisotropic properties, superconductivity, and its phonon-mediated form [32,33,34]. Borophene was successfully fabricated on a single surface of Ag(111) under ultrahigh-vacuum conditions [30,31]. Different borophene phases were observed at various deposition temperatures using a high-resolution scanning tunneling microscope (STM), such as the striped, β12, and χ3 phases [31]. The puckered shape and metallic characteristics of the striped phase led to its utility in various potential applications for metal ion storage and electric conduction [35]. Compared with the striped phase, the preferable stability of the β12 and the χ3 phases, with planar shapes having hexagonal and triangular vacancies, was demonstrated [36,37]. Because of its structure with a hexagonal vacancy, borophene was utilized in adsorbing gas molecules [10,38,39,40].
An upsurge in interest has recently been oriented toward investigating the utility of borophene and GN in the detection of gas molecules, like NO, CO, NO2, CO2, CS2, and NH3 molecules [40,41,42,43]. Halomethanes are known for being toxic molecules [44,45,46]; however, scant attention has been directed toward exploring novel nanomaterials for adsorbing them. Using density functional theory (DFT), the adsorption of tetrahalomethanes CX4 (X = F, Cl, and Br) was studied on carbon nanotubes [47] and GN nanosheets [48]. Nevertheless, no comparative study provided a full insight into the adsorption process of the tetrahalomethanes via all their possible oriented configurations on the surface of the borophene and GN nanosheets.
Herein, the adsorption features of tetrahalomethanes (CX4, where X = F, Cl, and Br) on the β12 borophene (β12) nanosheet were unveiled and compared with those with the utilization of the GN nanosheet as the starting 2D nanomaterial. The CX4∙∙∙β12/GN complexes were selectively studied within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations (Figure 1) using various density functional theory (DFT) method. Initially, geometry relaxation of the potential binding modes of the two suggested configurations and their corresponding adsorption energies were first carried out. Additionally, to assess the change in the electronic characteristics of the studied 2D nanosheets following the adsorption of the CX4 molecules, Bader charge, electronic band structure, and density of state (DOS) calculations were conducted. Further, the solvent effect on the adsorption energy of the studied complexes was evaluated. The obtained results would be an informative base for the utilization of the β12 and GN in adsorbing small molecules, such as tetrahalomethanes.

2. Results and Discussion

2.1. Geometric Structures

β12 and GN structures were modeled and relaxed before the adsorption process of the tetrahalomethanes. The optimized β12 and GN structures are presented in Figure 2. The obtained equilibrium lattice constants for the primitive cells of the β12 nanosheet were a = 5.06 Å and b = 2.93 Å. For the GN nanosheet, the equilibrium lattice constants were a = b = 2.47 Å. The current findings are consistent with earlier research [30,49,50]. On the β12 optimized structure, six adsorption sites were detected, comprising three top (T1, T2, and T3), two bridge (Br1 and Br2), and one hollow (H) sites (Figure 2). Looking at the GN surface, three adsorption sites, namely the top (T), bridge (Br), and hollow (H) sites, were noticed (Figure 2).

2.2. Adsorption Energy Calculations

The adsorption of tetrahalomethanes CX4 (where X = F, Cl, and Br) on the surfaces of β12 and GN was investigated at different adsorption sites within the tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. The adsorption energies and the corresponding equilibrium distances of all relaxed CX4∙∙∙β12/GN complexes were calculated and are summarized in Table 1. Figure S1 illustrates all relaxed complexes. The relaxed CX4∙∙∙β12/GN complexes at the most energetically preferable adsorption sites are displayed in Figure 3.
For the adsorption process of the CX4 on the β12 nanosheet within the tetrel-oriented configuration, the BrC-Br3∙∙∙β12 complexes had the most significant Eads values, followed by the ClC-Cl3∙∙∙β12, then the FC-F3∙∙∙β12 complexes (Table 1). Numerically, the Eads of the BrC-Br3∙∙∙, ClC-Cl3∙∙∙, and FC-F3∙∙∙Br1@β12 complexes were −12.33, −7.74, and −4.46 kcal/mol, respectively. These findings were in accord with a prior study, indicating that the adsorption energies increased with the increasing atomic size of the halogen atom (decreasing the electronegativity of the halogen atom) [51]. It is worth noting that the most preferred complex was the BrC-Br3∙∙∙Br1@β12 complex, with an Eads value of −12.33 kcal/mol and an equilibrium distance of 4.11 Å. In line with the tetrel-oriented configuration, energetic manifestations of the halogen-oriented complexes (i.e., X3C-X∙∙∙β12) showed the existence of a direct correlation between the adsorption energy and the atomic size of the halogen atom. Apparently, the H@β12 site was the most appropriate adsorption site for adsorbing the X3C-X molecules in the halogen-oriented configuration. Moreover, the Br3C-Br∙∙∙H@β12 complex had the most prominent Eads with a value of −9.00 kcal/mol at an equilibrium distance of 2.98 Å. The efficiency of the β12 nanosheet to adsorb the CX4 molecules was more significant in the tetrel-oriented configuration than in the halogen-oriented one (Table 1). For instance, the Eads values of the adsorption of the CBr4 at the Br1@β12 site were −12.33 and −8.91 kcal/mol in the tetrel- and halogen-oriented configurations, respectively.
For the adsorption of CX4 on the GN nanosheet, all complexes showed negative Eads values, confirming the occurrence of the adsorption process. Similar to the CX4∙∙∙β12 complexes, the preferentiality of the adsorption process of the CX4 molecules on the GN nanosheet increased upon the following order X = F < Cl < Br. Obviously, the T@GN site had the highest tendency for adsorbing the studied tetrahalomethanes on the GN sheet, exhibiting significant Eads values. Numerically, the ClC-Cl3∙∙∙ and Cl3C-Cl∙∙∙T@GN complexes had Eads values of −7.32 and −4.22 kcal/mol, respectively (Table 1).
For all CX4∙∙∙β12/GN complexes, the obtained Eads values ranged from −2.46 to −12.33 kcal/mol, demonstrating the occurrence of physisorption processes. The latter observation was in line with the literature, which reported the adsorption energy of CH4∙∙∙, CF4∙∙∙, and CCl4∙∙∙GN complexes with values of −1.61, −3.46, and −8.99 kcal/mol, respectively [52]. While the adsorption of the CH4 molecule on the borophene nanosheet exhibited a small Eads value of −2.54 kcal/mol and was accordingly documented as a physisorption process [53]. For a given type of halogen, the β12 nanosheet showed more affinity to adsorb the CX4 molecules than the GN nanosheet, which can be attributed to the lower electronegativity of boron relative to carbon and, hence, a lower electronegativity difference compared with that of the halogen.
Besides, the adsorption of the tetrahalomethanes became more favorable by decreasing the electronegativity of the halogens in the following order CF4∙∙∙ > CCl4∙∙∙ > CBr4∙∙∙β12/GN, which was accompanied by a lower electronegativity difference in the case of boron compared with carbon atoms. The favorability of the adsorption process within the tetrel-oriented configuration might be attributed to the contribution of the three halogen atoms of XC-X3 molecules to the overall interaction.

2.3. Frontier Molecular Orbital (FMO) Calculations

In order to comprehend the effect of the adsorption process on the electronic characteristics of the examined systems, the energies of the highest occupied molecular orbitals (EHOMO), the lowest unoccupied molecular orbitals (ELUMO), and the energy gap (Egap) values were assessed. Table 2 shows data of the EHOMO, ELUMO, and Egap values of the investigated systems before and following the adsorption process.
According to the data in Table 2, notable differences in the EHOMO, ELUMO, and Egap values were observed for the studied systems before and following the adsorption process. For instance, in the tetrel-oriented configuration, the EHOMO value of the BrC-Br3∙∙∙Br1@β12 complex was −2.544 eV, whereas the pure β12 nanosheet had an EHOMO value of −2.875 eV (Table 2). Moreover, the Egap values of all CX4 molecules, β12, and GN nanosheets were altered, confirming the occurrence of adsorption processes. For example, the pure β12 nanosheet had an Egap value of −0.626 eV that was changed to −0.602 eV after the adsorption process within the BrC-Br3∙∙∙Br1@β12 complex (Table 2).

2.4. Charge Transfer Calculations

The Bader charge method is a reliable appliance for determining the charge transfer over the adsorption process [54,55]. The transferred charge between the CX4 molecules and the 2D nanosheets within all the studied complexes was evaluated in terms of the charge transfer difference (Qt) values (Table 1). The Qt values with negative signs remarked that the charge was shifted from the CX4 molecules towards the β12 and GN nanosheets, and vice versa was true for the positive Qt values.
Table 1 shows Qt values with a negative sign for the CX4∙∙∙β12 complexes, demonstrating the ability of the inspected tetrahalomethanes to donate electrons to the β12 nanosheets within the tetrel- and halogen-oriented configurations. Notably, the Qt values of the CX4∙∙∙β12 complexes within the halogen-oriented configuration generally decreased as the adsorption energies decreased (i.e., in the order CBr4∙∙∙ > CC14∙∙∙ > CF4∙∙∙β12). For instance, the Qt values of the Br3C-Br∙∙∙, Cl3C-Cl∙∙∙, and F3C-F∙∙∙T1@β12 complexes were −0.0654, −0.0410, and −0.0164 e, respectively. The reversed trend was noticed for the complexes within the tetrel-oriented configuration, outlining the noticeable contributions of the three coplanar halogen atoms to the adsorption process. For example, the Qt values for the CF4∙∙∙, CCl4∙∙∙, and CBr4∙∙∙Br1@β12 complexes within the tetrel-oriented configuration were −0.0313, −0.0283, and −0.0263 e, respectively.
The Qt values of the CX4∙∙∙GN complexes within the tetrel- and halogen-oriented configurations showed similar trends to the CX4∙∙∙β12 complexes, except for the CBr4∙∙∙GN complexes within the former configuration that had positive Qt values. For the latter complexes, the electron-accepting property increased in the following order, H@GN < Br@GN < T@GN adsorption sites, and was confirmed with positive Qt values of 0.0070, 0.0023, and 0.0036 e, respectively.
Charge density difference (∆ρ) maps were generated in order to investigate the distribution of the charge within the relaxed CX4∙∙∙β12/GN complexes at the most preferable adsorption sites, and the maps are provided in Figure 4. As demonstrated in Figure 4, the electron depletion and accumulation regions (i.e., cyan- and yellow-colored regions, respectively) revealed the distribution of the charge between the tetrahalomethanes and the investigated 2D nanosheets. Apparently, the most remarkable electron-accumulated region was observed for the CBr4∙∙∙β12/GN complexes, demonstrating the further ability of CBr4 molecules to be adsorbed on the studied 2D nanosheets among tetrahalomethane analogs (Figure 4).
Overall, the negative Qt values confirmed that all the CX4 molecules had an electron-donating character except for the CBr4∙∙∙GN complexes within the tetrel-oriented configuration. According to the Qt values, the amount of charge transferred from the tetrahalomethanes to the β12 nanosheet was more significant than that of the GN nanosheet, which was in line with the adsorption energy values. Based on ∆ρ maps, it was observed that the amount of the distribution charge area (colored area) increased as the electronegativity of the halogen atom decreased. For instance, the size of the distribution charge area increased as the atomic size of the halogen atom increased in the order FC-F3∙∙∙ < ClC-Cl3∙∙∙ < BrC-Br3∙∙∙Br1@β12 complexes.
Given that the Qt values of the CX4∙∙∙β12/GN complexes within the halogen-oriented configuration generally decreased as the adsorption energies decreased, while the reversed trend was noticed for the complexes within the tetrel-oriented configuration. This implies that the halogen orientation relies on a more localized charge transfer and electronegativity difference during the binding mechanism, while the tetrel orientation is accompanied by a more distributed charge transfer binding that is more significant with large-sized bromine atoms.

2.5. Band Structure Calculations

To ascertain the impact of the adsorption of the CX4 molecules on the electronic properties of the β12 and GN nanosheets, band structure analysis was carried out for the pure and combined 2D nanosheets. Using PBE functional along the high-symmetry paths of the Brillouin zone, band structures were extracted. The Γ-Y-S-X-Γ path was chosen for the β12 nanosheet, and the Y-S-X-Γ-Y path was selected for the GN nanosheet. The band structures of the pure 2D nanosheets are demonstrated in Figure S2.
Looking at Figure S2, a metallic character of the pure β12 surface was noted by several bands, which crossed the Fermi level along the high-symmetry path. For the pure GN surface, the existence of the Dirac point at the Fermi level announced its semiconducting property.
Band structures of the relaxed CX4∙∙∙β12/GN complexes at the most preferable adsorption sites are plotted in Figure 5. After the adsorption process, slight differences were noticed in the electronic band structures of the pure nanosheets, outlining the physisorption process of tetrahalomethanes on the pure nanosheets (Figure 5).
For the adsorption of CF4 molecules, insignificant changes were denoted in the electronic band structures of the β12 nanosheet. Upon adsorbing CCl4 and CBr4 molecules, further new bands appeared in the band structures of the combined nanosheets compared with the pure analogs. Such new bands remarked the adsorption of the CCl4 and CBr4 molecules on the β12 nanosheet. Illustratively, the CBr4∙∙∙β12 complexes displayed a new conduction band at 1.35 eV and new valence bands at −0.60 and −2.00 eV. It was also observed that the bands shifted towards the Fermi level in the case of the complexes within the tetrel-oriented configuration more than the halogen-oriented analog. For instance, the adsorption of CCl4 at the Br1@β12 and H@β12 within the tetrel- and halogen-oriented configurations led to the appearance of a conduction band at around 2.70 and 2.15 eV, respectively. This observation demonstrated the higher favorability of the adsorption process within the former configuration than the latter one.
Similar to the β12 nanosheet, the CF4 molecules had a neglected effect on the band structure of the pure GN surface (Figure 5). Besides, the band structures of the CCl4∙∙∙ and CBr4∙∙∙GN complexes showed many new valence and conduction bands, confirming the higher potentiality of the GN nanosheet to adsorb these molecules compared with CF4 molecules. For instance, in the CBr4∙∙∙GN complexes, a new conduction band appeared at 0.60 and 0.67 eV, respectively, while in the valence region, many valence bands appeared at −2.40 eV and then ranged from −2.62 to −2.65 eV (Figure 5). It can be seen that the valence and conduction bands in the CX4∙∙∙GN complexes shifted towards the Fermi level as the atomic size of the halogen atom increased, demonstrating a favorable adsorption process. For example, the valence band at around −2.55 eV in the FC-F3∙∙∙T@GN complex shifted to −2.60 eV in the ClC-Cl3∙∙∙T@GN complex, and then to −2.65 eV in the BrC-Br3∙∙∙T@GN complex (Figure 5).
Summing up, the band structures of the β12 nanosheet demonstrated more new bands after adsorbing the CX4 molecules than those of the GN nanosheet. The latter affirmation indicated the further desirability of the adsorption process on the β12 nanosheet than the GN nanosheet. The obtained findings were in line with the adsorption energy affirmations. The appearance of the new bands after the adsorption process indicated the overlap of the bands of the adsorbent and substrate, confirming the interaction between the CX4 molecule and the studied 2D nanosheet. Further, the number of the new bands increased as the electronegativity of the halogen atom decreased. Illustratively, the CBr4∙∙∙β12/GN complexes, which exhibited the highest negative adsorption energy, showed the largest number of new bands among the other complexes (Figure 5).

2.6. Density of State Calculations

The total density of state (TDOS), together with the projected density of state (PDOS), were extracted for pure and combined 2D nanosheets to truly comprehend the impact of the adsorption process on the electronic characteristics of the 2D nanosheets (Figure S3). TDOS and PDOS plots of the most favorable complexes are shown in Figure 6.
The PDOS plots with the contribution of the p-orbital of B, C, and X atoms within the studied complexes were plotted in the energy range from −7.00 to 7.00 eV for β12 and from −8.00 to 8.00 eV for GN.
As shown in Figure 6, intense and feeble peaks were observed for the contributions of the PDOS of the X P C X 4 and C P C X 4 , respectively, to the TDOS of all the studied complexes. Accordingly, the halogens and carbon atoms of the CX4 molecules exhibited major and minor roles within the adsorption process on the 2D nanosheets, respectively.
For example, the contribution of Clp to the CCl4∙∙∙β12 and ∙∙∙GN complexes within the tetrel-oriented configuration were found in the valence region ranging from −2.50 to −4.70 eV and −3.00 to −5.10 eV, respectively. At the same time, the contribution of Clp also appeared in the conduction regions from 1.70 to 2.30 eV and; 3.00 to 3.60 eV for the adsorption over the β12 nanosheet and between 2.50 and 3.00 eV for the GN analog. Within the halogen-oriented configuration, the Clp peaks of the adsorbed CCl4 molecule on the β12 and GN nanosheets were noticed in the valence region between −2.40 and −4.50 eV and −3.00 and −5.00 eV, respectively. In the conduction region, the contribution of the Clp of the adsorbed CCl4 molecule on the β12 and GN nanosheets were found in the energy ranges of 1.80–2.50 and 3.00–4.00 eV, and 1.10–1.60 and 2.50–3.10 eV, respectively.
Notably, hybridizations between the p-orbital of the 2D nanosheets and the p-orbital of the CX4 molecules were also observed, revealing the occurrence of the adsorption process (Figure 6). For instance, an overlap between the PDOS(Bp) and the PDOS(Clp) appeared in the range from −3.50 to −3.90 eV in the CCl4∙∙∙Br1@β12 complex within the tetrel-oriented configuration, affirming the ability of the β12 nanosheet to adsorb the CCl4 molecule. The latter observation was consistent with the Eads of the CCl4∙∙∙Br1@β12 complex with a value of −7.74 kcal/mol (Table 1). While in the CCl4∙∙∙H@β12 complex within the halogen-oriented configuration, a small overlap between the PDOS(Bp) of the β12 nanosheet and the PDOS(Clp) of the CCl4 molecule was noticed in the conduction region from 1.80 to 2.20 eV. This finding was in agreement with the small Eads value of −5.58 kcal/mol.
From the DOS outlines, halogens had the dominant role in the adsorption of the CX4 molecules on the β12 and GN nanosheets within the modeled configurations. In line with the adsorption-energy and band structure findings, the DOS plots revealed the favorability of the β12 nanosheet over the GN analog to adsorb the tetrahalomethanes.

2.7. Solvent Effect Calculations

To speculate the effect of the solvent on the adsorption process within the CX4∙∙∙β12/GN complexes, the adsorption energy was evaluated in the presence of a water solvent. Afterwards, the solvent effect ( E a d s s o l v e n t e f f e c t ) energy was computed for the most preferable complexes as the difference between the adsorption energies of the water solvent and vacuum (see the Computational Methodology section for details). The obtained E a d s w a t e r and E a d s s o l v e n t e f f e c t values are listed in Table 3.
According to the data presented in Table 3, the adsorption energies of the CX4∙∙∙β12/GN complexes in the water medium showed higher negative values compared with those in a vacuum. For instance, the E a d s w a t e r and E a d s v a c u u m values of the CBr4∙∙∙Br1@β12 complex within the tetrel-oriented configuration were −15.99 and −12.33 kcal/mol, respectively (Table 1 and Table 3, respectively). Subsequently, E a d s solvent effect exhibited negative values, confirming the occurrence of the adsorption process in the water medium. As an illustration, the E a d s solvent effect value of the CBr4∙∙∙Br1@β12 complex within the tetrel-oriented configuration was −3.66 kcal/mol. Similar to the energetic manifestation obtained in a vacuum, the more prevalent effect of the water solvent on the favorability of the adsorption process was ascribed to the complexes within the tetrel-oriented configuration compared with the halogen-oriented configuration. Numerically, as an example, the E a d s solvent effect values of the CBr4∙∙∙Br1@β12 and ∙∙∙H@β12 complexes within the tetrel- and halogen-oriented configurations were −3.66 and −1.60 kcal/mol, respectively.

3. Computational Methods

The density functional theory (DFT) method was applied for all calculations [56,57] via the Quantum ESPRESSO 6.4.1 package [58,59]. Based on the Perdew–Burke–Ernzerhof (PBE) scheme, the electron exchange-correlation function was conducted utilizing the generalized gradient approximation (GGA) [60]. To represent the electron–core interaction, the ultrasoft pseudopotential (USPP) was employed [61]. The van der Waals interactions for all the executed computations were taken into account using the Grimme-D2 method [62]. The utilized energy cutoff and charge density cutoff values were 50 and 500 Ry, respectively. The total energy and the atomic force convergence criteria were 1 × 10−5 eV and 1 × 10−4 eV/Å, respectively. Based on the Monkhorst-Pack mesh, the 6 × 6 × 1 and 12 × 12 × 1 k-points grids were adopted for the first Brillouin zone sampling within the geometry relaxation and density of state calculations, respectively. The convergence was enhanced using the Marzari–Vanderbilt smearing method [63]. For preventing image–image interaction, a vacuum thickness of 20 Å was added along the z-direction of the β12 and GN nanosheets.
To model the adsorption of the tetrahalomethanes (CX4; X = F, Cl, and Br) on β12 and GN nanosheets, 3 × 4 × 1 and 6 × 5 × 1 supercells were constructed for β12 and GN nanosheets, respectively. Adsorption energies (Eads) of the CX4∙∙∙β12/GN complexes within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations were assessed as follows:
E a d s = E C X 4 · · · 2 D   n a n o s h e e t ( E C X 4 + E 2 D   n a n o s h e e t )
where E C X 4 · · · 2 D   n a n o s h e e t , E C X 4 , and E 2 D   n a n o s h e e t are the energies of complex, tetrahalomethane, and 2D nanosheet, respectively. Frontier molecular orbital (FMO) calculations were carried out to gain a better understanding of the adsorption process of CX4 molecules on the investigated 2D nanosheets. Within the FMO analyses, the energies of the highest occupied molecular orbitals (EHOMO) and lowest unoccupied molecular orbitals (ELUMO) for the most stable relaxed CX4∙∙∙β12/GN complexes were computed. The energy gap (Egap) was estimated according to the following equation:
E g a p = E L U M O E H O M O
The charge transfer of the adsorbed CX4 molecules was determined using the Bader charge method [55,64] based on the following equation:
Q t = Q c o m b i n e d   2 D   n a n o s h e e t s Q i s o l a t e d   2 D   n a n o s h e e t s
where Q c o m b i n e d   2 D   n a n o s h e e t s and Q i s o l a t e d   2 D   n a n o s h e e t s are the charges of the 2D nanosheets after adsorbing tetrahalomethanes and the charge of the isolated 2D nanosheets, respectively. The charge density difference (∆ρ) was plotted according to the following equation:
ρ = ρ C X 4 · · · 2 D   n a n o s h e e t ρ C X 4 ρ 2 D   n a n o s h e e t
where ρ C X 4 · · · 2 D   n a n o s h e e t , ρ C X 4 , and ρ 2 D   n a n o s h e e t are the charge densities of complex, tetrahalomethane, and 2D nanosheet, respectively. VESTA 3 visualization software was invoked for generating the charge density plots [65]. To comprehend the influence of the adsorption process of the tetrahalomethanes on the electronic characteristics of the β12 and GN nanosheets, band structure and density of state (DOS) calculations were executed. For implicit water solvent calculations, the Environ code [66] of Quantum ESPRESSO was utilized with a dielectric constant of 78.3. The solvent effect on the adsorption energy of the studied complexes ( E a d s solvent effect ) was computed according to the following equation:
E a d s solvent effect = E a d s w a t e r E a d s v a c u u m
where E a d s w a t e r and E a d s v a c u u m are the adsorption energies of the complex in water and vacuum media, respectively.

4. Conclusions

In the presented work, a DFT study was conducted to comparatively illustrate the adsorption features of tetrahalomethanes (CX4, where X = F, Cl, and Br) on β12 borophene (β12) and GN nanosheets. To attain a thorough investigation, geometry relaxation, adsorption energies, Bader charge, electronic band structures, and DOS computations were conducted for the adsorption of the CX4 molecules on the studied 2D nanosheets within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. From the energetic perspective, the adsorption of the CX4 model on the β12 and GN nanosheets within the tetrel-oriented configuration was more desirable than that within the halogen-oriented configuration. Further favorability of the Br1@β12 and T@GN adsorption sites were announced toward adsorbing the CX4 molecules within the tetrel-oriented configuration and showed the most significant Eads for the CBr4 molecule with values of −12.33 and −10.03 kcal/mol, respectively. According to the FMO results, the EHOMO, ELUMO, and Egap values of the β12 and GN nanosheets were changed following the adsorption process. Based on the Bader charge results, the electron-donating characters for all the CX4 molecules after adsorbing on the β12 and GN nanosheets within tetrel- and halogen-oriented configurations were illustrated, except the CBr4∙∙∙GN complexes within the former configuration. In the latter complexes, the adsorbed CBr4 molecule showed an electron-accepting property confirmed by the small positive Qt values. From the band structure and DOS plots, new bands and peaks were observed, respectively, after the adsorption of CX4 molecules on the 2D nanosheets, indicating the occurrence of the adsorption process. The energetic results are pertinent to the solvent effect demonstrated, that the presence of the water solvent led to more observable negative adsorption energies compared with the adsorption in a vacuum. The emerging findings would provide a foundation for any future consideration of β12 and GN nanosheets to adsorb small molecules.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28145476/s1, Figure S1: Side and top representations for the relaxed structures of the tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations of the CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at all the adsorption sites. Equilibrium distances (d) are given in Å; Figure S2: Electronic band structures of β12 and GN nanosheets along the high symmetry points of the Brillouin zone. The Fermi energy was set at zero energy, and the Dirac point is defined by the dotted circle; Figure S3: Total and projected density of state (TDOS/PDOS) plots for the pure surfaces of β12 and GN nanosheets, assuming Fermi level as the reference level. The dotted circle defines the Dirac point. The contributions of the p-orbital for boron (B) and carbon (C) atoms are represented by Bp and Cp, respectively.

Author Contributions

Conceptualization, M.A.A.I. and T.S.; Methodology, M.A.A.I., A.H.M.M. and T.S.; Software, M.A.A.I.; Formal analysis, A.H.M.M.; Investigation, A.H.M.M. and N.A.M.M.; Resources, M.A.A.I., S.R.M.S. and T.S.; Data curation, A.H.M.M.; Writing—original draft preparation, A.H.M.M.; Writing—review and editing, M.A.A.I., N.A.M.M., G.A.H.M., S.R.M.S., M.N.A., M.K.A.E.-R., E.D. and T.S.; Visualization, A.H.M.M.; Supervision, M.A.A.I. and G.A.H.M.; Project administration, M.A.A.I. and G.A.H.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Acknowledgments

The authors extend their appreciation to the Researchers Supporting Project number (RSPD2023R743), King Saud University, Riyadh, Saudi Arabia, for funding this work. The computational work was performed with resources provided by the Science and Technology Development Fund (STDF-Egypt, Grants Nos. 5480 and 7972), Bibliotheca Alexandrina (http://hpc.bibalex.org, accessed on 1 July 2023), and The American University in Cairo. Mahmoud A. A. Ibrahim extends his appreciation to the Academy of Scientific Research and Technology (ASRT, Egypt) for funding the Graduation Projects conducted at CompChem Lab, Egypt.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds (XYZ coordinates) are available from the authors.

References

  1. Cao, M.S.; Wang, X.X.; Zhang, M.; Shu, J.C.; Cao, W.Q.; Yang, H.J.; Fang, X.Y.; Yuan, J. Electromagnetic response and energy conversion for functions and devices in low-dimensional materials. Adv. Funct. Mater. 2019, 29, 1807398. [Google Scholar] [CrossRef]
  2. Tang, Q.; Zhou, Z. Graphene-analogous low-dimensional materials. Prog. Mater. Sci. 2013, 58, 1244–1315. [Google Scholar] [CrossRef]
  3. Xia, F.; Wang, H.; Jia, Y. Rediscovering black phosphorus as an anisotropic layered material for optoelectronics and electronics. Nat. Commun. 2014, 5, 4458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ibrahim, M.A.A.; Mahmoud, A.H.M.; Soliman, K.A.; Mekhemer, G.A.H.; Ahmed, M.N.; Shawky, A.M.; Abourehab, M.A.S.; Elkaeed, E.B.; Soliman, M.E.S.; Moussa, N.A.M. Borophene and pristine graphene 2D sheets as potential surfaces for the adsorption of electron-rich and electron-deficient pi-systems: A comparative DFT study. Nanomaterials 2022, 12, 1028. [Google Scholar] [CrossRef]
  5. Novoselov, K.S.; Fal’ko, V.I.; Colombo, L.; Gellert, P.R.; Schwab, M.G.; Kim, K. A roadmap for graphene. Nature 2012, 490, 192–200. [Google Scholar] [CrossRef]
  6. Schedin, F.; Geim, A.K.; Morozov, S.V.; Hill, E.W.; Blake, P.; Katsnelson, M.I.; Novoselov, K.S. Detection of individual gas molecules adsorbed on graphene. Nat. Mater. 2007, 6, 652–655. [Google Scholar] [CrossRef]
  7. Leenaerts, O.; Partoens, B.; Peeters, F.M. Adsorption of H2O, NH3, CO, NO2, and NO on graphene: A first-principles study. Phys. Rev. B 2008, 77, 125416–125421. [Google Scholar] [CrossRef] [Green Version]
  8. Ren, Z.; Zhang, H.; Liu, N.; Lei, D.; Zhang, Q.; Su, T.; Wang, L.; Su, J.; Gao, Y. Self-powered 2D nanofluidic graphene pressure sensor with Serosa-Mimetic structure. EcoMat 2022, 5, e12299. [Google Scholar] [CrossRef]
  9. Xiao, H.; Li, Y.; Chen, R.; Xie, T.; Xu, P.; Zhu, H.; He, J.; Zheng, W.; Huang, S. Integrative design of laser-induced graphene array with lithiophilic MnOx nanoparticles enables superior lithium metal batteries. eScience 2023, 100134. [Google Scholar] [CrossRef]
  10. Mujib, S.B.; Ren, Z.; Mukherjee, S.; Soares, D.M.; Singh, G. Design, characterization, and application of elemental 2D materials for electrochemical energy storage, sensing, and catalysis. Mater. Adv. 2020, 1, 2562–2591. [Google Scholar] [CrossRef]
  11. Liu, X.M.; Xu, T.; Li, Y.L.; Zang, Z.G.; Peng, X.S.; Wei, H.Y.; Zha, W.Y.; Wang, F. Enhanced X-ray photon response in solution-synthesized CsPbBr3 nanoparticles wrapped by reduced graphene oxide. Sol. Energy Mater. Sol. Cells 2018, 187, 249–254. [Google Scholar] [CrossRef]
  12. Rao, C.N.; Sood, A.K.; Subrahmanyam, K.S.; Govindaraj, A. Graphene: The new two-dimensional nanomaterial. Angew. Chem. Int. Ed. Engl. 2009, 48, 7752–7777. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, Y.; Tan, Y.W.; Stormer, H.L.; Kim, P. Experimental observation of the quantum Hall effect and Berry’s phase in graphene. Nature 2005, 438, 201–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Balandin, A.A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C.N. Superior thermal conductivity of single-layer graphene. Nano Lett. 2008, 8, 902–907. [Google Scholar] [CrossRef]
  15. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [Green Version]
  16. Miura, Y.; Kasai, H.; Diño, W.; Nakanishi, H.; Sugimoto, T. First principles studies for the dissociative adsorption of H2 on graphene. J. Appl. Phys. 2003, 93, 3395–3400. [Google Scholar] [CrossRef]
  17. Durgun, E.; Ciraci, S.; Yildirim, T. Functionalization of carbon-based nanostructures with light transition-metal atoms for hydrogen storage. Phys. Rev. B 2008, 77, 085405. [Google Scholar] [CrossRef] [Green Version]
  18. Liu, J.; Cui, L.; Losic, D. Graphene and graphene oxide as new nanocarriers for drug delivery applications. Acta Biomater. 2013, 9, 9243–9257. [Google Scholar] [CrossRef]
  19. Davis, M.E.; Chen, Z.G.; Shin, D.M. Nanoparticle therapeutics: An emerging treatment modality for cancer. Nat. Rev. Drug Discov. 2008, 7, 771–782. [Google Scholar] [CrossRef]
  20. Ibrahim, M.A.A.; Hamad, M.H.A.; Mahmoud, A.H.M.; Mekhemer, G.A.H.; Sayed, S.R.M.; El-Rahman, M.K.A.; Sidhom, P.A.; Dabbish, E.; Shoeib, T. On the use of graphene nanosheets for drug delivery: A case study of cisplatin and some of its analogs. Pharmaceutics 2023, 15, 1640. [Google Scholar] [CrossRef]
  21. Ibrahim, M.A.A.; Hamad, M.H.A.; Mahmoud, A.H.M.; Mekhemer, G.A.H.; Sidhom, P.A.; Sayed, S.R.M.; Moussa, N.A.M.; Rabee, A.I.M.; Dabbish, E.; Shoeib, T. Adsorption of Favipiravir on pristine graphene nanosheets as a drug delivery system: A DFT study. RSC Adv. 2023, 13, 17465–17475. [Google Scholar] [CrossRef] [PubMed]
  22. Palacios, J.J.; Fernandez-Rossier, J.; Brey, L. Vacancy-induced magnetism in graphene and graphene ribbons. Phys. Rev. B 2008, 77, 195428. [Google Scholar] [CrossRef] [Green Version]
  23. Hou, Y.; Wen, Z.H.; Cui, S.M.; Ci, S.Q.; Mao, S.; Chen, J.H. An advanced nitrogen-doped graphene/cobalt-embedded porous carbon polyhedron hybrid for efficient catalysis of oxygen reduction and water splitting. Adv. Funct. Mater. 2015, 25, 872–882. [Google Scholar] [CrossRef]
  24. Varghese, S.S.; Lonkar, S.; Singh, K.K.; Swaminathan, S.; Abdala, A. Recent advances in graphene based gas sensors. Sens. Actuators B-Chem. 2015, 218, 160–183. [Google Scholar] [CrossRef]
  25. Zhang, S.; Yan, Z.; Li, Y.; Chen, Z.; Zeng, H. Atomically thin arsenene and antimonene: Semimetal-semiconductor and indirect-direct band-gap transitions. Angew. Chem. Int. Ed. Engl. 2015, 54, 3112–3115. [Google Scholar] [CrossRef]
  26. Khan, A.F.; Randviir, E.P.; Brownson, D.A.C.; Ji, X.B.; Smith, G.C.; Banks, C.E. 2D hexagonal boron nitride (2D-hBN) explored as a potential electrocatalyst for the oxygen reduction reaction. Electroanalysis 2017, 29, 622–634. [Google Scholar] [CrossRef] [Green Version]
  27. Zhang, S.; Xie, M.; Li, F.; Yan, Z.; Li, Y.; Kan, E.; Liu, W.; Chen, Z.; Zeng, H. Semiconducting group 15 monolayers: A broad range of band gaps and high carrier mobilities. Angew. Chem. Int. Ed. Engl. 2016, 55, 1666–1669. [Google Scholar] [CrossRef]
  28. Aufray, B.; Kara, A.; Vizzini, S.; Oughaddou, H.; Leandri, C.; Ealet, B.; Le Lay, G. Graphene-like silicon nanoribbons on Ag(110): A possible formation of silicene. Appl. Phys. Lett. 2010, 96, 183102. [Google Scholar] [CrossRef] [Green Version]
  29. Du, Y.; Zhuang, J.; Liu, H.; Xu, X.; Eilers, S.; Wu, K.; Cheng, P.; Zhao, J.; Pi, X.; See, K.W.; et al. Tuning the band gap in silicene by oxidation. ACS Nano 2014, 8, 10019–10025. [Google Scholar] [CrossRef] [Green Version]
  30. Feng, B.; Zhang, J.; Zhong, Q.; Li, W.; Li, S.; Li, H.; Cheng, P.; Meng, S.; Chen, L.; Wu, K. Experimental realization of two-dimensional boron sheets. Nat. Chem. 2016, 8, 563–568. [Google Scholar] [CrossRef] [Green Version]
  31. Mannix, A.J.; Zhou, X.F.; Kiraly, B.; Wood, J.D.; Alducin, D.; Myers, B.D.; Liu, X.; Fisher, B.L.; Santiago, U.; Guest, J.R.; et al. Synthesis of borophenes: Anisotropic, two-dimensional boron polymorphs. Science 2015, 350, 1513–1516. [Google Scholar] [CrossRef] [Green Version]
  32. Ranjan, P.; Lee, J.M.; Kumar, P.; Vinu, A. Borophene: New sensation in flatland. Adv. Mater. 2020, 32, e2000531. [Google Scholar] [CrossRef]
  33. Peng, B.; Zhang, H.; Shao, H.Z.; Xu, Y.F.; Zhang, R.J.; Zhua, H.Y. The electronic, optical, and thermodynamic properties of borophene from first-principles calculations. J. Mater. Chem. C 2016, 4, 3592–3598. [Google Scholar] [CrossRef] [Green Version]
  34. Gao, M.; Li, Q.Z.; Yan, X.W.; Wang, J. Prediction of phonon-mediated superconductivity in borophene. Phys. Rev. B 2017, 95, 024505. [Google Scholar] [CrossRef] [Green Version]
  35. Jiang, H.R.; Lu, Z.H.; Wu, M.C.; Ciucci, F.; Zhao, T.S. Borophene: A promising anode material offering high specific capacity and high rate capability for lithium-ion batteries. Nano Energy 2016, 23, 97–104. [Google Scholar] [CrossRef]
  36. Tang, H.; Ismail-Beigi, S. Novel precursors for boron nanotubes: The competition of two-center and three-center bonding in boron sheets. Phys. Rev. Lett. 2007, 99, 115501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Yang, X.B.; Ding, Y.; Ni, J. Ab initio prediction of stable boron sheets and boron nanotubes: Structure, stability, and electronic properties. Phys. Rev. B 2008, 77, 041402. [Google Scholar] [CrossRef]
  38. Li, W.L.; Jiang, Q.G.; Li, D.D.; Ao, Z.M.; An, T.C. Density functional theory investigation on selective adsorption of VOCs on borophene. Chin. Chem. Lett. 2021, 32, 2803–2806. [Google Scholar] [CrossRef]
  39. Joshi, D.J.; Malek, N.I.; Kailasa, S.K. Borophene as a rising star in materials chemistry: Synthesis, properties and applications in analytical science and energy devices. New J. Chem. 2022, 46, 4514–4533. [Google Scholar] [CrossRef]
  40. Ta, L.T.; Hamada, I.; Morikawa, Y.; Dinh, V.A. Adsorption of toxic gases on borophene: Surface deformation links to chemisorptions. RSC Adv. 2021, 11, 18279–18287. [Google Scholar] [CrossRef]
  41. Liu, T.T.; Chen, Y.H.; Zhang, M.L.; Yuan, L.H.; Zhang, C.R.; Wang, J.; Fan, J.J. A first-principles study of gas molecule adsorption on borophene. AIP Adv. 2017, 7, 125007. [Google Scholar] [CrossRef] [Green Version]
  42. Huang, C.S.; Murat, A.; Babar, V.; Montes, E.; Schwingenschlogl, U. Adsorption of the gas molecules NH3, NO, NO2, and CO on borophene. J. Phys. Chem. C 2018, 122, 14665–14670. [Google Scholar] [CrossRef]
  43. Ibrahim, M.A.A.; Mahmoud, A.H.M.; Mekhemer, G.A.H.; Shawky, A.M.; Soliman, M.E.S.; Moussa, N.A.M. Adsorption behavior of toxic carbon dichalcogenides (CX2; X = O, S, or Se) on β12 borophene and pristine graphene sheets: A DFT study. Nanomaterials 2022, 12, 3411. [Google Scholar] [CrossRef] [PubMed]
  44. Radilov, A.S.; Shkayeva, I.E.; Solntseva, S.A.; Kondrashov, V.A.; Menshikov, N.M.; Nikulin, O.S. Experimental studies of the toxic properties of halogen-derivatives of saturated hydrocarbons (Chladons). Toxicol. Rev. 2017, 35–41. [Google Scholar] [CrossRef] [Green Version]
  45. Agarwal, A.K.; Berndt, W.O.; Mehendale, H.M. Possible nephrotoxic effect of carbon tetrabromide and its interaction with chlordecone. Toxicol. Lett. 1983, 17, 57–62. [Google Scholar] [CrossRef]
  46. Klingensmith, J.S.; Mehendale, H.M. Potentiation of CCl4 lethality by chlordecone. Toxicol. Lett. 1982, 11, 149–154. [Google Scholar] [CrossRef]
  47. Babaa, M.R.; Dupont-Pavlovsky, N.; McRae, E.; Masenelli-Varlot, K. Physical adsorption of carbon tetrachloride on as-produced and on mechanically opened single walled carbon nanotubes. Carbon 2004, 42, 1549–1554. [Google Scholar] [CrossRef]
  48. Bermudez, V.M.; Robinson, J.T. Effects of molecular adsorption on the electronic structure of single-layer graphene. Langmuir 2011, 27, 11026–11036. [Google Scholar] [CrossRef]
  49. Li, K.; Li, N.; Yan, N.N.; Wang, T.Y.; Zhang, Y.T.; Song, Q.; Li, H.J. Adsorption of small hydrocarbons on pristine, N-doped and vacancy graphene by DFT study. Appl. Surf. Sci. 2020, 515, 146028. [Google Scholar] [CrossRef]
  50. Chakarova-Kack, S.D.; Schroder, E.; Lundqvist, B.I.; Langreth, D.C. Application of van der Waals density functional to an extended system: Adsorption of benzene and naphthalene on graphite. Phys. Rev. Lett. 2006, 96, 146107. [Google Scholar] [CrossRef] [Green Version]
  51. Ha, M.; Kim, D.Y.; Li, N.; Madridejos, J.M.L.; Park, I.K.; Youn, I.S.; Lee, J.; Baig, C.; Filatov, M.; Min, S.K.; et al. Adsorption of carbon tetrahalides on coronene and graphene. J. Phys. Chem. C 2017, 121, 14968–14974. [Google Scholar] [CrossRef]
  52. Singh, D.; Kumar, A.; Kumar, D. Adsorption of small gas molecules on pure and Al-doped graphene sheet: A quantum mechanical study. Bull. Mater. Sci. 2017, 40, 1263–1271. [Google Scholar] [CrossRef] [Green Version]
  53. Qin, G.Q.; Cui, Q.Y.; Du, A.J.; Sun, Q. Borophene: A metal-free and metallic electrocatalyst for efficient converting CO2 into CH4. Chemcatchem 2020, 12, 1483–1490. [Google Scholar] [CrossRef]
  54. Bader, R.F.W.; Nguyen-Dang, T.T. Quantum theory of atoms in molecules–dalton revisited. In Advances in Quantum Chemistry; Löwdin, P.-O., Ed.; Academic Press: Cambridge, MA, USA, 1981; Volume 14, pp. 63–124. [Google Scholar]
  55. Henkelman, G.; Arnaldsson, A.; Jonsson, H. A fast and robust algorithm for Bader decomposition of charge density. Comp. Mater. Sci. 2006, 36, 354–360. [Google Scholar] [CrossRef]
  56. Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  57. Kresse, G.; Furthmuller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comp. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  58. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef] [PubMed]
  59. Giannozzi, P.; Andreussi, O.; Brumme, T.; Bunau, O.; Buongiorno Nardelli, M.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Cococcioni, M.; et al. Advanced capabilities for materials modelling with Quantum ESPRESSO. J. Phys. Condens. Matter 2017, 29, 465901. [Google Scholar] [CrossRef] [Green Version]
  60. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  61. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892–7895. [Google Scholar] [CrossRef]
  62. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  63. Marzari, N.; Vanderbilt, D.; De Vita, A.; Payne, M.C. Thermal contraction and disordering of the Al(110) surface. Phys. Rev. Lett. 1999, 82, 3296–3299. [Google Scholar] [CrossRef] [Green Version]
  64. Kutzelnigg, W. Book review: Atoms in molecules. A quantum theory. (International Series Monographs on Chemistry, Vol. 22). By R. F. W. Bader. Angew. Chem. Int. Ed. Engl. 1993, 32, 128–129. [Google Scholar] [CrossRef]
  65. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  66. Andreussi, O.; Dabo, I.; Marzari, N. Revised self-consistent continuum solvation in electronic-structure calculations. J. Chem. Phys. 2012, 136, 064102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Side and top representations of the CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) within (a) tetrel (XC-X3)- and (b) halogen (X3C-X)-oriented configurations.
Figure 1. Side and top representations of the CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) within (a) tetrel (XC-X3)- and (b) halogen (X3C-X)-oriented configurations.
Molecules 28 05476 g001
Figure 2. Side and top perspectives of the relaxed structures of 3 × 4 × 1 β12 borophene (β12) and 6 × 5 × 1 pristine graphene (GN) with the studied adsorption sites. The boron and carbon atoms are represented by pink and gray colors, respectively. Top, hollow, and bridge adsorption sites are referred to as T, H, and Br, respectively.
Figure 2. Side and top perspectives of the relaxed structures of 3 × 4 × 1 β12 borophene (β12) and 6 × 5 × 1 pristine graphene (GN) with the studied adsorption sites. The boron and carbon atoms are represented by pink and gray colors, respectively. Top, hollow, and bridge adsorption sites are referred to as T, H, and Br, respectively.
Molecules 28 05476 g002
Figure 3. Side and top perspectives of the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at the most preferable adsorption sites within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. Equilibrium distances (d) are in Å. The boron, carbon, fluorine, chlorine, and bromine atoms are defined by pink, gray, pale green, green, and red colors, respectively.
Figure 3. Side and top perspectives of the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at the most preferable adsorption sites within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. Equilibrium distances (d) are in Å. The boron, carbon, fluorine, chlorine, and bromine atoms are defined by pink, gray, pale green, green, and red colors, respectively.
Molecules 28 05476 g003
Figure 4. Charge density difference (∆ρ) maps of the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at the most preferable adsorption sites within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. Regions with cyan and yellow colors refer to depletion (negative) and accumulation (positive) charges, respectively. The isosurface values for the CX4∙∙∙β12 and ∙∙∙GN were set to be 3.08 × 10−5 and 5.0 × 10−5 e3, respectively.
Figure 4. Charge density difference (∆ρ) maps of the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at the most preferable adsorption sites within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. Regions with cyan and yellow colors refer to depletion (negative) and accumulation (positive) charges, respectively. The isosurface values for the CX4∙∙∙β12 and ∙∙∙GN were set to be 3.08 × 10−5 and 5.0 × 10−5 e3, respectively.
Molecules 28 05476 g004
Figure 5. Band structure plots for the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at the most preferable adsorption sites within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. The Fermi energy was positioned at zero energy.
Figure 5. Band structure plots for the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at the most preferable adsorption sites within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. The Fermi energy was positioned at zero energy.
Molecules 28 05476 g005
Figure 6. Total density of state (TDOS) plots for the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at the most preferable adsorption sites within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. The projected density of state (PDOS) with the contribution of Bp, Cp-GN, C P C X 4 , and X P C X 4 . The Fermi energy was set at zero energy.
Figure 6. Total density of state (TDOS) plots for the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at the most preferable adsorption sites within tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. The projected density of state (PDOS) with the contribution of Bp, Cp-GN, C P C X 4 , and X P C X 4 . The Fermi energy was set at zero energy.
Molecules 28 05476 g006
Table 1. Adsorption energies (Eads, kcal/mol) and equilibrium distances (d, Å) of the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at all possible sites within the tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. Charge transfer difference (Qt, e) for the 2D nanosheets before and after the adsorption process.
Table 1. Adsorption energies (Eads, kcal/mol) and equilibrium distances (d, Å) of the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at all possible sites within the tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations. Charge transfer difference (Qt, e) for the 2D nanosheets before and after the adsorption process.
2D NanosheetsAdsorption
Site a
X = FX = ClX = Br
Eads
(kcal/mol)
d
(Å)
Qtb
(e)
Eads
(kcal/mol)
d
(Å)
Qtb
(e)
Eads
(kcal/mol)
d
(Å)
Qtb
(e)
Tetrel-oriented configuration
β12T1−4.423.69−0.0309−7.474.09−0.0275−11.454.14−0.0231
T2−4.253.72−0.0289−7.694.06−0.0230−11.424.13−0.0080
T3−4.143.68−0.0309−7.224.11−0.0275−11.074.17−0.0215
H−4.393.65−0.0306−7.214.07−0.0219−11.154.13−0.0115
Br1−4.463.67−0.0313−7.744.05−0.0283−12.334.11−0.0263
Br2−4.143.74−0.0291−7.154.12−0.0269−11.034.18−0.0175
GNT−4.663.50−0.0175−7.323.93−0.0072−10.034.070.0036
Br−4.363.57−0.0177−6.824.02−0.0051−9.494.130.0023
H−4.123.63−0.0168−6.574.06−0.0025−9.434.130.0070
Halogen-oriented configuration
β12T1−2.543.10−0.0164−5.143.16−0.0410−8.653.10−0.0654
T2−2.623.12−0.0149−4.333.26−0.0254−6.723.21−0.0291
T3−2.633.06−0.0167−4.973.17−0.0366--- c--- c--- c
H−2.712.93−0.0185−5.583.02−0.0321−9.002.98−0.0424
Br1−2.463.11−0.0165−5.253.13−0.0413−8.912.98−0.0697
Br2−2.693.05−0.0163−4.263.26−0.0258−6.373.24−0.0317
GNT−2.463.00−0.0095−4.223.16−0.0149−6.003.17−0.0196
Br−2.472.99−0.0094−4.183.17−0.0152−5.933.18−0.0174
H−2.612.89−0.0093−3.993.18−0.0093−5.543.23−0.0086
a All adsorption sites on the investigated 2D nanosheets are depicted in Figure 2. b Qt was calculated based on Equation (3). c Desired configuration was not observed after geometry relaxation (see Figure S1).
Table 2. The energies of the highest occupied molecular orbitals (EHOMO, eV), the lowest unoccupied molecular orbitals (ELUMO, eV), and the energy gap (Egap, eV) of the CX4 molecules and the β12/GN nanosheets before and after the adsorption process.
Table 2. The energies of the highest occupied molecular orbitals (EHOMO, eV), the lowest unoccupied molecular orbitals (ELUMO, eV), and the energy gap (Egap, eV) of the CX4 molecules and the β12/GN nanosheets before and after the adsorption process.
Complex aEHOMO
(eV)
ELUMO
(eV)
Egap
(eV)
Isolated systems
GN Nanosheet−2.354−2.3430.011
β12 Nanosheet−2.875−3.501−0.626
CF4−10.333−0.4779.855
CCl4−7.416−2.6804.735
CBr4−6.644−3.3943.250
Tetrel-oriented Configuration
FC-F3∙∙∙Br1@β12−2.734−3.358−0.625
ClC-Cl3∙∙∙Br1@β12−2.602−3.217−0.615
BrC-Br3∙∙∙Br1@β12−2.544−3.146−0.602
FC-F3∙∙∙T@GN−2.202−2.1910.0104
ClC-Cl3∙∙∙T@GN−2.064−2.0540.0107
BrC-Br3∙∙∙T@GN−2.010−1.9990.0108
Halogen-oriented Configuration
F3C-F∙∙∙H@β12−2.737−3.364−0.627
Cl3C-Cl∙∙∙H@β12−2.612−3.237−0.626
Br3C-Br∙∙∙H@β12−2.564−3.190−0.626
F3C-F∙∙∙H@GN−2.200−2.1900.0104
Cl3C-Cl∙∙∙T@GN−2.066−2.0560.0104
Br3C-Br∙∙∙T@GN−2.021−2.0100.0105
a Structures of the most stable relaxed CX4∙∙∙β12/GN complexes within both configurations are presented in Figure 3.
Table 3. Adsorption energy in the vacuum medium ( E a d s v a c u u m , kcal/mol), water medium ( E a d s w a t e r , kcal/mol), and the energy of the solvent effect ( E a d s solvent effect , kcal/mol) for the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at the most preferable adsorption sites within the tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations.
Table 3. Adsorption energy in the vacuum medium ( E a d s v a c u u m , kcal/mol), water medium ( E a d s w a t e r , kcal/mol), and the energy of the solvent effect ( E a d s solvent effect , kcal/mol) for the relaxed CX4∙∙∙β12/GN complexes (where X = F, Cl, and Br) at the most preferable adsorption sites within the tetrel (XC-X3)- and halogen (X3C-X)-oriented configurations.
System a E a d s v a c u u m
(kcal/mol)
E a d s w a t e r
(kcal/mol)
E a d s solvent effect b
(kcal/mol)
Tetrel-oriented Configuration
FC-F3∙∙∙Br1@β12−4.46−6.91−2.45
ClC-Cl3∙∙∙Br1@β12−7.74−11.21−3.47
BrC-Br3∙∙∙Br1@β12−12.33−15.99−3.66
FC-F3∙∙∙T@GN−4.66−7.14−2.48
ClC-Cl3∙∙∙T@GN−7.32−10.99−3.67
BrC-Br3∙∙∙T@GN−10.03−14.09−4.06
Halogen-oriented Configuration
F3C-F∙∙∙H@β12−2.71−4.22−1.51
ClC-Cl3∙∙∙H@β12−5.58−7.26−1.68
BrC-Br3∙∙∙H@β12−9.00−10.60−1.60
F3C-F∙∙∙H@GN−2.61−4.17−1.56
Cl3C-Cl∙∙∙T@GN−4.22−5.98−1.76
Br3C-Br∙∙∙T@GN−6.00−7.72−1.72
a The structures of the relaxed complexes are depicted in Figure 3. b E a d s solvent effect = E a d s w a t e r E a d s v a c u u m .
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ibrahim, M.A.A.; Mahmoud, A.H.M.; Moussa, N.A.M.; Mekhemer, G.A.H.; Sayed, S.R.M.; Ahmed, M.N.; Abd El-Rahman, M.K.; Dabbish, E.; Shoeib, T. Adsorption Features of Tetrahalomethanes (CX4; X = F, Cl, and Br) on β12 Borophene and Pristine Graphene Nanosheets: A Comparative DFT Study. Molecules 2023, 28, 5476. https://doi.org/10.3390/molecules28145476

AMA Style

Ibrahim MAA, Mahmoud AHM, Moussa NAM, Mekhemer GAH, Sayed SRM, Ahmed MN, Abd El-Rahman MK, Dabbish E, Shoeib T. Adsorption Features of Tetrahalomethanes (CX4; X = F, Cl, and Br) on β12 Borophene and Pristine Graphene Nanosheets: A Comparative DFT Study. Molecules. 2023; 28(14):5476. https://doi.org/10.3390/molecules28145476

Chicago/Turabian Style

Ibrahim, Mahmoud A. A., Amna H. M. Mahmoud, Nayra A. M. Moussa, Gamal A. H. Mekhemer, Shaban R. M. Sayed, Muhammad Naeem Ahmed, Mohamed K. Abd El-Rahman, Eslam Dabbish, and Tamer Shoeib. 2023. "Adsorption Features of Tetrahalomethanes (CX4; X = F, Cl, and Br) on β12 Borophene and Pristine Graphene Nanosheets: A Comparative DFT Study" Molecules 28, no. 14: 5476. https://doi.org/10.3390/molecules28145476

Article Metrics

Back to TopTop