Next Article in Journal
Supersaturation-Based Drug Delivery Systems: Strategy for Bioavailability Enhancement of Poorly Water-Soluble Drugs
Previous Article in Journal
A Calixarene Assembly Strategy of Combined Anti-Neuroinflammation and Drug Delivery Functions for Traumatic Brain Injury Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Harnessing Greenhouse Gases Absorption by Doped Fullerenes with Externally Oriented Electric Field

by
Rodrigo A. Lemos Silva
1,*,
Daniel F. Scalabrini Machado
2,
Núbia Maria Nunes Rodrigues
3,
Heibbe C. B. de Oliveira
4,
Luciano Ribeiro
3 and
Demétrio A. da Silva Filho
1,*
1
Institute of Physics, University of Brasília, Brasília 70919-970, Brazil
2
Laboratório de Modelagem de Sistemas Complexos (LMSC), Instituto de Química, Universidade de Brasília, Brasília 70919-970, Brazil
3
Grupo de Química Teórica e Estrutural de Anápolis, Campus de Ciências Exatas de Anápolis, Universidade Estadual de Goiás, Anápolis 75132-903, Brazil
4
Laboratório de Estrutura Eletrônica e Dinâmica Molecular (LEEDMOL), Instituto de Química, Universidade Federal de Goiás, Goiânia 74001-970, Brazil
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(9), 2968; https://doi.org/10.3390/molecules27092968
Submission received: 1 March 2022 / Revised: 13 April 2022 / Accepted: 29 April 2022 / Published: 6 May 2022
(This article belongs to the Topic Nanomaterials for Sustainable Energy Applications)

Abstract

:
In this work, a theoretical investigation of the effects caused by the doping of C20 with silicon (Si) atom as well as the adsorption of CO, CO2 and N2 gases to C20 and C19Si fullerenes was carried out. In concordance with previous studies, it was found that the choice of the doping site can control the structural, electronic, and energetic characteristics of the C19Si system. The ability of C20 and C19Si to adsorb CO, CO2 and N2 gas molecules was evaluated. In order to modulate the process of adsorption of these chemical species to C19Si, an externally oriented electric field was included in the theoretical calculations. It was observed that C19Si is highly selective with respect to CO adsorption. Upon the increase of the electric field intensity the adsorption energy was magnified correspondingly and that the interaction between CO and C19Si changes in nature from a physical adsorption to a partial covalent character interaction.

1. Introduction

Carbon nanostructures have been continuously studied for the most diverse applications [1,2,3,4,5,6]. Among these nanostructures, fullerenes [7,8] stand out because of their good physicochemical reactivity and their symmetrical and relatively simple structure [9,10]. These characteristics render fullerenes, whether in their pristine or doped form, an attractive molecular system to be object of several theoretical studies aiming at various sorts of applications [11,12,13,14,15,16].
Among the fullerenes, C20 is the smallest known example containing only twelve pentagonal faces. Since the existence of this allotrope of carbon was already predicted [17], even before its experimental discovery [18], C20 and its derivatives already had an extensive history of studies [19,20,21,22].
Considered the most reactive of the fullerenes [18], interest in C20 has continued over time [23,24,25,26,27]. This molecule has been the object of several studies that aim to carry out its doping to investigate aspects related to its electronic and structural characteristics, in addition to proposing possible applications in different areas.
In a recent theoretical investigation, Metin and co-authors [15] used calculations with the theoretical chemistry model B3LYP/6-31G(d) to investigate both the hydrogen storage capacity and the electronic properties of C20, C15M5 and H2@C15M5, with M = Al, Si, Ga, Ge. Among the conclusions of the research, the authors highlighted that, fullerenes doped with Si and Ge, more specifically in the form C15Si5 and C15Ge5, were highly sensitive to the presence of H2. In addition, the ability of C20-doping to manipulate the physicochemical and structural parameters of fullerenes was also confirmed.
The analysis of C20, in its pure form and doped with aluminum (C20−nAln; n = 1–5), was recently performed by Hassanpour and coauthors [23]. In a theoretical effort, employing various levels of Density Functional Theory (DFT) calculations, they investigated the Al substitution in the C20 cage. Through analysis of the infrared spectrum (Infrared-IR), the authors observed that both the number and the position of Al atoms can change the IR spectrum. Furthermore, they observed that by controlling the number of dopants, the concentration and distribution of atomic charges (APT charges) can be altered to favor adsorption of chemical species to the C20−nAln cage.
Amongst the various atoms possible for the doping of C20, Si has drawn attention in recent years. The Si atom belongs to the same family as carbon and may have interesting properties when used as an impurity in the cage of carbon nanostructures. Doping of C20 with Si was theoretically studied by Koohi, Amiri and SHariatI [28], using DFT calculations, the authors analyzed the effects of doping C20−nSin with n = 1–10. This strategy indicates that, in the same vein as in the case of Al doping, controlling the position and number of Si heteroatoms makes it possible to modulate structural changes of the molecule as well as the energies of the molecular orbitals. Similar results were also observed theoretically by Ajeel and co-authors [29].
Beside the aforementioned examples, there are interest in applying C20 and its derivatives for the most varied purposes, whether for hydrogen storage [15], as a drug delivery system [30,31] or for the detection of biomolecules [32]. A very promising application of these carbonaceous systems is in detection and capture of many gases [14,33] especially those associated with the greenhouse effect and/or harmful to health.
In this spirit, the present work is dedicated to the study of C20 and C19Si for the adsorption of carbon monoxide (CO) and carbon dioxide (CO2). Both CO and CO2 are products of burning fossil fuels. While CO2 is one of the main gases responsible for the greenhouse effect, along with methane gas, CO contamination is considered by the World Health Organization as one of the main causes of accidental poisoning around the world each year [34].
Theoretical DFT-based calculations were carried out to investigate the effects of the interaction between C19Si and CO and CO2 molecules. Recently, there has been much attention to the use of externally oriented electric fields (EOEF) to harness physical and chemical properties of molecular systems [35,36,37]. Intermolecular interactions between weakly interacting systems can be strongly aided by EOEFs, since it has the advantage of being easy to achieve and control, and is environmentally friendly [38]. On the realm of molecular sensors, previous works [16,39,40,41,42,43], have demonstrated that the adsorption of molecules and the storage of H2 can be facilitated in the presence of an EOEF. Additionally, EOEF might render selectivity to the sensor towards a given species in complex gaseous mixtures [16,39].
Based in our interest in theoretical investigations aiming at the development of new sensors [44], we disclose herein the influence of an externally oriented electric field to rationally control the adsorption process of gas molecules on doped C19Si fullerenes relying on DFT calculations. Since approximately 78% of the Earth’s atmosphere is made up of nitrogen gas (N2) [45], the selectivity of C19Si for the detection of CO and CO2 was evaluated by comparing the adsorption energy between the C19Si cage and the CO, CO2 and N2 molecules.

2. Materials and Methods

To confirm the structural and electronic trends observed in the fullerene cage after doping with a silicon atom, an investigation involving several theoretical levels was carried out. In all calculations, unrestricted optimizations were performed. The absence of negative frequencies confirmed that the molecules are at their energetic and structural minimum at each theoretical level considered. Several combinations of Exchange Correlation DFT functionals and basis sets (ωB97XD/6-31G(d), ωB97XD/6-311+G(d,p), ωB97XD/def2TZVP, M062X/6-31G(d), M062X/6-311+G(d,p), M062X/def2TZVP, M06L/6-31G(d), M06L/6-311G(d,p), M06L/def2TZVP, B3LYP/6-31G(d), B3LYP/6-311+G(d,p) and B3LYP/def2TZVP) were tested to screen electronic and structural properties of pristine and doped fullerenes. The screening of the theoretical model on the electronic and structural properties revealed the following combination of DFT Exchange-correlation function and basis set, ωB97XD/6-31G(d) and ωB97XD/6-311+G(d,p), as the optimal choice for the investigation of the dimers interaction considering the compromise of accuracy and computation cost. While the 6-31G(d) basis set has been reported to be a good choice for obtaining fairly reliable results in fullerenes [11,46], using the more extended 6-311+G(d,p) basis set improves the description of the non-covalent interactions, for further analysis of Quantum Theory of Atoms in Molecules (QTAIM) parameters and Reduced Density Gradient (RDG) properties as reported in recent studies [47,48].
Adsorption energies under electric field F , were calculated using the supramolecular approach considering the optimized structures by the expression [49],
E a d s ( F ) = E g a s C 19 X ( F ) [ E g a s ( F ) + E C 19 X ( F ) ] ;   X = S i   o r   C .
In Equation (1), E a d s ( F ) is the adsorption energy. E g a s C 19 X ( F ) refers to the energy of the dimer formatted by the C19X and the gas molecules, E g a s ( F ) is the energy of the gas molecule and E C 19 X ( F ) is the energy of the fullerene C20, for X = C or C19Si for X = Si.
The adsorption energies were corrected for the Basis Set Superposition Error (BSSE) using the Counterpoise method [50]. The calculations included EOEF, F , in atomic units (a.u.) (1 a.u. = 51 V/Å) in the range: F = 0.000 a.u., 0.001 a.u., 0.005 a.u., 0.010 a.u., 0.020 a.u. and 0.025 a.u. along the Si heteroatom (see Figure S1 for clarity).
The description of intermolecular interactions was performed by means of the QTAIM [51,52,53] and RDG analysis [54,55]. Molecular Electrostatic Potential (MEP) map, electronic and structural properties were calculate with the Gaussian 16 [56] software. QTAIM and RDG properties were calculated using Multiwfn [57] wave function analysis program. VMD software version 1.9.3 [58] was employed to render isosurfaces and molecular representations.

3. Results and Discussion

3.1. Doping Effects on the Electronic and Structural Properties of Fullerenes

The structure of doped fullerene was constructed by replacing a carbon atom for a silicon atom. Initially, the molecules had their structures optimized at the ωB97XD/6-31G(d) level of theory followed by another round of calculation at the ωB97XD/6-311+G(d,p) model chemistry. Both pure and doped fullerene are shown in Figure 1 and the respective coordinates of these molecules are shown in the Supplementary Materials (SM) Table S1.
From Figure 1 we clearly note that Si-doping promoted remarkable structural deformations of the fullerene cage. Due to its larger atomic radius, it was observed that the Si-C bond distances with the adjacent C-atoms was lengthened with respect to the C-C bond distances prior to the substitution. From a polarization perspective, Merz-Kollman (MK) charge analysis (using the 6-31G(d) basis set and in parenthesis with 6-311+G(d,p)) indicates that C20 has a modest charge distribution ranging from −0.018 (−0.103) to 0.018 (0.103) e (Figure 1D). For the C19Si, MK charges range from −0.341 (−0.366) to 0.341 (0.366) e (Figure 1E). Still on Figure 1D, it was also noted charge accumulation of −0.222 (−0.162) e on the Si-atom. MEPs shown in Figure 1D,E shows that, Si-doping imparts a polarization of the fullerene, with a dipole moment of C19Si of μ = 1.950 D (1.460 D) similar to what is observed in reported studies in literature [25,29]. As an immediate result of the polarization of the molecule, the effective intermolecular interaction can be strongly altered [11,59,60,61], so that C19Si display the ability to adsorb different molecular species aided by electrostatic forces (dipole-dipole, for instance) that was once absent in its pristine form.
Doping the C20 fullerene had direct impact on the electronic structure. Looking at the frontier eigenstates (HOMO and LUMO), shown in Table S3, addition of an Si-atom induced an increase in the HOMO-LUMO energy gap, E H L ~0.3 eV: for C20 fullerene, E H L = 5.46   eV   (6-31G(d)) and, E H L = 5.40   eV   (6-311+G (d,p)) whereas for C19Si DFT calculations delivered E H L = 5.81   eV   (6-31G(d)) and E H L = 5.72   eV   (6-311+G(d,p)). Thus, it is noted that the increase in E H L is a trend observed for both bases used. It is important to highlight that, in previous works, the doping of nanocarbon structures, including fullerenes C60 [62,63] and fullerenes C20 [64,65], led to a decrease in the value of E H L , which differs from the results observed in this work. Additionally, regarding the variation of the energy gap between the frontier orbitals, it is observed that the negative MK charges on the Si atom (see the MEP in Figure 1E), are not in line with the atomic charge on the heteroatom observed in the literature for C19Si [14,30].
To investigate whether such contradictions is an artifact of the level of theory employed, we performed additional unrestricted optimizations with different Exchange-Correlation DFT functionals (XCF) and basis set (see Materials and Methods section) keeping track to the absence of negative frequencies. The results of such extended theoretical survey on the frontier orbitals, as obtained at the B3LYP/6-31G(d), B3LYP/6-311+G(d,p), B3LYP/def2TZVP, ωB97XD/6-31G(d), ωB97XD/6-311+G(d,p), ωB97XD/def2TZVP, M06L/6-31G(d), M06L/6-311+G(d,p), M06L/def2TZVP, M062X/6-31G(d), M062X/6-311+G(d,p) and M062X/def2TZVP are reserved in Tables S3 and S4 found in the SM file to avoid proliferation of tables in the main text. Glancing at Tables S3 and S4 we report the percentual increase of HOMO-LUMO energy gap upon Si-doping, Δ E H L , and straightforwardly testify that regardless of the XCF/basis set combination, the addition of a Si-atom increases the chemical stability of C19Si over C20. Similarly, the apparent opposite findings of this work and those reported in literature [14,30] concerning the charge accumulation on Si-atom was retained for each of the screened XCF/basis set (see Table S4). From Table S4, partial MK charge on Si-atom, evidently is affected by the choice of XCF/basis set, nevertheless, all theoretical calculations consistently delivered a negative charge on Si. We, therefore, believe there is no dispute concerning this issue. Still on Table S4 we note that the different levels of calculations provoke slightly changes on bond lengths supporting our choice for ωB97XD/6-31G(d) and ωB97XD/6-311+G(d,p) as a reliable theoretical model for further analysis. Moreover, despite these contradictions with the results reported in some previous works, in which C20 was doped with Si [14,27,30], our calculations is in agreement with other similar works [24,25,29].
To further investigate the possible causes of the difference in results observed in these calculations with the results presented in the literature [14,27,30], and inspired by studies with the doping of C20 with Si and Al [23,28,29] we studied the impact of the position of Si-atoms in the fullerene cage. For this purpose, 20 configurations of C19Si were produced, named C19Si (X), where X = 1–20 indicate the label of the carbon atom replaced by a Si-atom, as depicted in Figure 2. All C19Si (X) were optimized without restrictions at the ωB97XD/6-311+G(d,p) level of theory.
Our results confirm that the doping of C20 with a Si atom generates a structural and electronic variation that depends on the position in which the impurity is inserted into the fullerene cage. However, only two types of C19Si geometrical conformations were retrieved. To simplify the discussion, these sets of substitutions will be referred to as the C19Si(A) set and the C19Si(B) set. Both sets are represented by Figure 2C,D, respectively.
The C19Si (A) set contains the following doped fullerenes: C19Si (1), C19Si (2), C19Si (4), C19Si (6), C19Si (7), C19Si (8), C19Si (11), C19Si (12), C19Si (13), C19Si (14), C19Si (15), C19Si (16), C19Si (17), C19Si (18) and C19Si (19). The C19Si (B) set is formed by the doped fullerenes C19Si (3), C19Si (5), C19Si (9), C19Si (10), C19Si (17) and C19Si (20). Looking at Figure 2 and Table 1, the C19Si (A) and C19Si (B) sets have similar members with close structural similarity, and the same tendency of charge accumulation on the Si atom, the same values of E H L in addition to the same total energy.
Thus, it can be seen in Table 1, that the E H L energy also presents different values depending on the position of the heteroatom. If the heteroatom is positioned to generate geometrical conformations of the C19Si (B) set (Figure 2D), E H L tends to increase. On the other hand, if the heteroatom replaces one of the carbons as indicated in C19Si (A) (Figure 2C), E H L tends to decrease. Considering that the variation of E H L is an indication of chemical stability in which higher values of E H L indicate a reduction in the chemical reactivity of a given molecule [66], we note from Table 1, that the geometrical conformers pertaining to C19Si (B) set are more chemically stable than those of the C19Si (A) set. Relying on B3LYP/3-21G calculations, Ajeel and coworker [29] did not observe significant changes in the E H L energy and in the structural properties of C19Si. Since only three out of ten C19Si geometrical conformers belong to the C19Si (B) set, a non-rational choice of doping site increases the odds of producing a C19Si (A) geometrical conformer.
Again, in Table 1, total electronic energy E t of the molecules was also considered to understand their relative stability. For the geometrical conformers of the C19Si (A) set E t = 1012.718   a . u . , whereas for the C19Si (B) set E t = 1012.777   a . u . Thus, based on this total energy criteria, the geometrical conformers of the C19Si (B) set are more stable than those of the C19Si (A) set corroborating the conclusions on the chemical stability made by comparing the values of E H L .
Table 1 also brings the relative population, η , of geometrical conformers based on a Boltzmann distribution, at 298.15 K, revealing that the molecules of the C19Si (A) group are close to zero while those of C19Si (B) are vastly dominant. We expect therefore, that geometrical conformers belonging to the C19Si (B) will be more likely produced in a synthesis of C19Si. Based on these calculations, it is observed that, in theory, the doping site can be chosen to generate the desired structural and electronic properties for each application. Consequently, the C19Si (10) was selected to be studied as a promising molecule to interact with the gas molecules, because C19Si (10) showed to be the most stable geometrical conformer belonging to the C19Si (B).

3.2. Adsorption Energies between C20 Fullerene and CO, CO2 and N2 Molecules

Now let’s turn our attention to the interacting energies between the investigated chemical species. Table S2 presents the respective coordinates of the dimers formatted between the C20 and C19Si fullerenes and the CO, CO2 and N2 molecules. In its pristine form, CO, CO2 and N2 molecules are adsorbed in a parallel orientation as shown in the molecular representations of the dimers in Table S7. The BSSE-corrected adsorption energy, E a d s , obtained through Equation (1), portrayed in Figure 3 and in Table S5.
From Figure 3 we note that both theoretical models point to the C20-CO2 as the most interacting structure with highest value of E a d s among the analyzed dimers. Interaction energies of C20-CO and C20-N2 are dissimilar, so there is no clear selectivity towards these two molecules. Our theoretical results indicate that the complexes studied in the present work are about twice as stable as the complexes investigated by Vessally et al. [33], which can be attributed to both the basis set and the functional chosen for the calculation.
To verify the sensitivity of C20 in relation to the adsorption of CO, CO2 and N2 molecules, the energy gap, E H L , of the dimers formed after the adsorption of the gases, was compared with its value prior to the adsorption of the gases. The results of this comparison are presented as a percentual change, represented by Δ E H L . The results of Δ E H L can be seen in Table S7. In general, it is observed that the HOMO-LUMO gap is nearly insensitive the presence of the interacting gases with Δ E H L ~ 0 in the absence of any EOEF. Thus, C20 fullerene is not sensitive for the detection of CO, CO2 and N2 molecules.

3.3. Electric Field Effect on the Adsorption Energies

To investigate the influence of the electric field, in both structural and energetic characteristics, adsorption calculations for the doped systems were performed. Following we report our findings following the increase of F .
F = 0.000   a . u .
The results indicate that the adsorption energies, when compared to the E a d s values of the dimers formed with the gases and the pristine fullerenes, increased for the adsorption of carbon monoxide and carbon dioxide. For C19Si-N2, a decrease in the energy module E a d s is observed when the calculation is performed with the level ωB97XD/6-31G(d).
For ease of interpretation, the energy values for the second theoretical level are shown in parentheses. The results show that E a d s increases to −0.703 eV (−0.738 eV) for C19Si-CO and 0.068   eV   ( 0.077   eV )   for C19Si-CO2. E a d s obtained with ωB97XD/6-311+G(d,p), resulted in a slight increase in the interaction between nitrogen gas and doped fullerene ( E a d s values observed for the C19Si-N2 dimer were 0.017   eV   ( 0.045   eV ) ). As can be seen in the Figures in Table S10 for C19Si-CO, the CO reorients to interact in a perpendicular configuration. For C19Si-CO2 and C19Si-N2, the molecules remain interacting in a parallel orientation.
Since the CO molecule has a permanent dipole moment, in which the carbon atom has a positive charge and the oxygen atom has a negative charge, the interaction between the CO molecule and C19Si is expected to be the most favorable among the three chemicals analyzed. This is most likely due to the nucleophilic behavior observed on the Si atom in C19Si fullerene. This charge accumulation on the heteroatom favors the adsorption of CO in the doped cage through a directional dipolar interaction. The results for the doped system are detailed in Table S6 and in the Figures in Table S10.
The percentage variation of the energy gap between the frontier orbitals, Δ E H L , indicates that C19Si is highly sensitive to the detection of carbon monoxide, and, as can be seen in Table S7, Δ E H L ~ 20 % for the two theoretical levels. The negative sign indicates a reduction in the energy gap. For CO2 and N2, Δ E H L values show no significant changes, i.e., C19Si is not sensitive to the presence of CO2 and N2.
F 0.000   a . u .
As can be seen in Tables S6 and S10 and in Figure 4, for both calculation levels, the E a d s values for F = 0.001   a . u . are very similar to the E a d s values in the absence of the electric field. From the Figures in Table S10, it is noted that the effects of the electric field, F = 0.001   a . u . , is to reorient the CO2 and N2 molecules, so that they interact with a pentagonal face of the C19Si cage, which has a more electrophilic character (see Figure 1E). Table S7 shows that as the electric field increases in intensity, the energy gap reduces correspondingly for all dimers, and this effect is more pronounced when C19Si absorbs CO2 and N2. The ωB97XD/6-311+G(d,p) derived Δ E H L values, with the strongest electric field F = 0.025   a . u . , of the C19Si-CO2 and C19Si-N2 dimers is very close to the variation observed for the C19Si-CO dimer. However, even for this strong F , C19Si presents greater sensitivity for the detection of the CO molecule.
Now let’s discuss the EOEF effect on the adsorption energies for the doped fullerene with F the range of 0.005   a . u . and 0.025   a . u . , presented in Figure 4. We note from Table S6 and Figure 4 that F and E a d s show direct proportionality, and for the C19Si-CO system, this increase appears almost linear. On the other hand, for the C19Si-CO2 and C19Si-N2 systems, F affects E a d s , with a prominent parabolic behavior as corroborated by fitting correlation polynomial expressions on the F   vs .   E a d s portrayed in Figure 4. Clearly, such correlation equations are valid only within the range of F considered in this work, so that even further increasing in the electric field tends to decrease the separation of the molecules and eventually steric repulsions start to dominate.

3.4. Intermolecular Interaction Characterization under EOEF Influence

To investigate the character of the intermolecular interactions, we employed QTAIM and RDG analyses. QTAIM analysis allows for investigation of the nature of intra/intermolecular interactions relying on topological properties of the electron density. The use of QTAIM analysis has been successful in characterizing and describing interactions in various chemical systems [60,61,67,68,69,70].
The results obtained for the pure and doped dimers are shown in Tables S8–S10. It is observed that, for all dimers in pure form, with F = 0 , the values of the electron density, ρ B C P , are in the order of 10 3 e / a 0 3 . The Laplacian values of electron density, ρ B C P 2 , always have positive values. Furthermore, the ratio of kinetic energy density to potential energy density, | G B C P / V B C P | have values greater than one unit. These results of ρ B C P , ρ B C P 2 and | G B C P / V B C P | indicate that C20-CO, C20-CO2 and C20-N2 dimers are stabilized primarily through van der Waals (vdW) interactions. The RDG scatter plot and the isosurfaces presented in the Figures in Table S10 corroborate these observations.
For doped fullerenes, as can be seen in Tables S8–S10, the dimers C19Si-CO2 and C19Si-N2 continue to present values of ρ B C P , ρ B C P 2 and | G B C P / V B C P | consistent with observed values for vdW interactions [71,72,73,74,75,76,77,78]. These results remain for the two theoretical levels and for F < 0.025   a . u . The RDG scatter plots, for both calculation levels, confirm the non-covalent character of vdW-type observed by the QTAIM analyses. When F = 0.025   a . u , the QTAIM parameters and the RDG plot indicate that the interaction changes from vdW to a dipolar character for C19Si-CO2. When C19Si-N2 is considered, the interaction between the N2 atom and the doped is still of vdW-type.
For the C19Si-CO dimer, the results for the two theoretical levels indicate higher values of ρ B C P , (about 10 2 e / a 0 3 ), for all values electric field. The values of ρ B C P 2 remain positive and | G B C P / V B C P | < 1 . According to the characterization of the QTAIM parameters [71,72,73,74,75,76,77,78], C19Si-CO shows a dipolar interaction character, at the ωB97XD/6-31G(d) level and a polar-covalent character for the ωB97XD/6-311+G(d,p) level when F < 0.0150   a . u . . Beyond F = 0.015   a . u . , C19Si starts to show a tendency to interact with CO through an interaction with partial polar-covalent character for both theoretical levels. This observation agrees with the characteristics observed in the MEP of C19Si (Figure 1), with the E a d s values obtained for C19Si-CO (Table S6) and with the directionality observed for the interaction between CO and C19Si shown in Table S10.
An interesting fact was observed in the analysis of QTAIM in the C19Si-CO system with the theoretical level ωB97XD/6-31G(d). For this complex, the presence of critical degenerate points was observed in the intermolecular region between the Si and O atoms, as can be seen in Table S10. This type of degeneracy is common in stability studies, especially in structures under the influence of an external factor such as temperature or in solvents [79,80,81,82]. Degenerate critical points are usually associated with unstable interactions [83]. The closer the RCP and the BCP are, the more unstable the interaction tends to be. In unstable interactions, a small energy disturbance can cause migration from an RCP to a BCP, which leads to the disappearance of RCP [83]. However, when the basis set was extended, this degeneracy disappears.

4. Conclusions

With the interest of investigating the adsorption of carbon monoxide, CO, carbon dioxide, CO2, and nitrogen gas, N2, the C20 fullerene was Si-doped. To investigate the impact of heteroatom positioning in the C20 cage, twenty C19Si were produced. In the present investigation, the Si substitution position influenced the structural and electronic characteristics of the doped system. This led to the formation of two groups of geometrical conformers, namely C19Si (A) and C19Si (B), classified according to their deformations. After an energetic analysis, it was observed that only the geometrical conformer of the C19Si(B) group are energetically favorable at the temperature of 298.15K.
The adsorption of CO, CO2 and N2 was not favorable to the C20 fullerene cage from the energetic point of view. However, it can be observed that the doping with a Si atom enabled the adsorption of the chemical species mentioned to the C19Si cage. This adsorption was intensified for all molecules aided by an externally oriented electric field. However, the formation of the C19Si-CO dimer was the most favored. Thus, C19Si has good selectivity in the adsorption of CO over CO2 and N2. Based on energetic and topological parameters, when the intensity of the EOEF is smaller than F < 0.015   a . u . , physisorption take place between CO and C19Si. Further increasing F , the adsorption shifts to a polar-covalent character as revealed by QTAIM parameters.
Adsorption energies shows a quadratic relationship with the EOEF with C19Si-CO2 and C19Si-N2 dimers, and a linear dependence in the case of C19S-CO. Obtaining the correlations for the interaction of these molecules with C19Si enable the ration use of EOEF to capture gaseous chemical species. DFT calculations revealed that C19Si fullerene is good prototype to selectively detection of carbon monoxide. Modulating the intensity of the EOEF C19Si displays is suitable for a CO uptake and detection.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27092968/s1, Figure S1: Orientation of the electric field vector with respect to the C19Si cage; Table S1: Optimized coordinates for C20 and C19Si fullerenes at theoretical levels ωB97XD/6-31G(d) and ωB97XD/6-311+G(d,p). Electic field F in atomic units and Total energies in hartree; Table S2: Optimized coordinates for the dimers formed by pure and doped fullerenes and CO, CO2 and N2 gases in each of the studied field values. Electic field F in atomic units and BSSE-corrected adsorption energies in electron-volts; Table S3: Frontier orbital energies (in eV) for C20 and C19Si fullerenes. Percentual increase of the HOMO-LUMO gap, ∆EHL is also shown; Table S4: Si-C bond lengths and Merz-Kollmann (MK) charges on the Si-atom and adjacent carbons.; Table S5: Adsorption energies (in eV), Eads, and adsorption energy with and without BSSE corrections, Eads (BSSE), for the dimers formed between C20 fullerenes with CO, CO2 and N2 molecules calculated with the theoretical levels ωB97XD/6-31G(d) and ωB97XD/6-311+G(d,p); Table S6: Externally oriented electric field effect on the BSSE correction, adsorption energies, Eads, and BSSE corrected, Eads (BSSE), for the interaction of C20 and C19Si fullerenes with the CO, CO2 and N2 molecules as obtained at the ωB97XD/6-31G(d) and ωB97XD/6-311+G(d,p) levels of theory. Adjusted adsorption energy, Eads (fit) and the R-squared coefficient of determination, R2, are also presented. All energies in eV; Table S7: Topological parameters calculated using the Quantum Theory of Atoms in Molecules (QTAIM) with the theoretical level ωB97XD/6-31G(d). Results are presented for all dimers formed by C20 and C19Si fullerenes with adsorbed gas molecules. The results were obtained for all electric field values and are all presented in atomic units; Table S8: Topological parameters calculated using the Quantum Theory of Atoms in Molecules (QTAIM) with the theoretical level ωB97XD /6-311+G(d,p). Results are presented for all dimers formed by C20 and C19Si fullerenes with adsorbed gas molecules. The results were obtained for all electric field values and are all presented in atomic units; Table S9: Critical bond points, BCP, isosurfaces and scatter plots for the dimers formed by C20 and C19Si fullerenes with the adsorbed gas molecules. The results were obtained for all electric field values using both the theoretical level ωB97XD/6-31G(d) and ωB97XD/6-311+G(d,p); Table S10: Energies (in eV) of the HOMO, EH, LUMO, EL and the gap between them, EHL, calculated with the theoretical levels ωB97XD/6-31G(d) and ωB97XD/6-311+G(d,p), for the dimers. The percentual change in the fullerene energy gap after adsorbing the gas, in relation to the fullerene energy gap before adsorption, is represented by ΔEHL (in %). Positive values of ΔEHL indicate an increase in the gap, negative values indicate a reduction.

Author Contributions

R.A.L.S. and D.F.S.M. designed the project, R.A.L.S., D.F.S.M. and H.C.B.d.O. performed the data analysis. R.A.L.S. wrote original draft preparation, R.A.L.S., N.M.N.R., D.A.d.S.F., H.C.B.d.O. and L.R writing—review and editing. R.A.L.S., D.A.d.S.F. and L.R. carried out the calculations. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior—Brasil (CAPES)—Finance Code 001 (Convênio nº 817164/2015 CAPES/PROAP), and by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq-grants 164561/2017-0) and Coordenação de Aperfeiçoamento Pessoal de Nível Superior (CAPES), and Fundação de Apoio à Pesquisa do Distrito Federal (FAP-DF), and Fundação de Amparo à Pesquisa do Estado de Goiás (FAPEG). D.A.S.F. acknowledges the financial support from the Edital DPI-UnB No. 02/2021, from CNPq (grants 305975/2019-6, and 420836/2018-7) and FAP-DF (grants 193.001.596/2017 and 193-00001220/2021-37).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article and in the Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Yang, C.M.; Chen, T.C.; Yang, Y.C.; Hsiao, M.C.; Meyyappan, M.; Lai, C.S. Ultraviolet illumination effect on monolayer graphene-based resistive sensor for acetone detection. Vacuum 2017, 140, 89–95. [Google Scholar] [CrossRef]
  2. Nomani, M.W.K.; Shishir, R.; Qazi, M.; Diwan, D.; Shields, V.B.; Spencer, M.G.; Tompa, G.S.; Sbrockey, N.M.; Koley, G. Highly sensitive and selective detection of NO2 using epitaxial graphene on 6H-SiC. Sens. Actuators B Chem. 2010, 150, 301–307. [Google Scholar] [CrossRef]
  3. Romero, H.E.; Joshi, P.; Gupta, A.K.; Gutierrez, H.R.; Cole, M.W.; Tadigadapa, S.A.; Eklund, P.C. Adsorption of ammonia on graphene. Nanotechnology 2009, 20, 245501. [Google Scholar] [CrossRef] [PubMed]
  4. Justino, C.I.L.; Gomes, A.R.; Freitas, A.C.; Duarte, A.C.; Rocha-Santos, T.A.P. Graphene based sensors and biosensors. TrAC Trends Anal. Chem. 2017, 91, 53–66. [Google Scholar] [CrossRef]
  5. Feng, X.; Zhang, Y.; Zhou, J.; Li, Y.; Chen, S.; Zhang, L.; Ma, Y.; Wang, L.; Yan, X. Three-dimensional nitrogen-doped graphene as an ultrasensitive electrochemical sensor for the detection of dopamine. Nanoscale 2015, 7, 2427–2432. [Google Scholar] [CrossRef]
  6. Li, X.; Zhao, H.; Shi, L.; Zhu, X.; Lan, M.; Zhang, Q.; Hugh Fan, Z. Electrochemical sensing of nicotine using screen-printed carbon electrodes modified with nitrogen-doped graphene sheets. J. Electroanal. Chem. 2017, 784, 77–84. [Google Scholar] [CrossRef]
  7. Kroto, H.W.; Walton, D.R.M. Stable derivatives of small fullerenes. Chem. Phys. Lett. 1993, 214, 353–356. [Google Scholar] [CrossRef]
  8. Osawa, E.; Kroto, H.W.; Fowler, P.W.; Wasserman, E. The evolution of the football structure for the C60 molecule: A retrospective. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 1993, 343, 1–8. [Google Scholar] [CrossRef]
  9. Kroto, H.W. The stability of the fulerenes Cn, with n = 24, 28, 32, 36, 50, 60 and 70. Nature 1987, 329, 529–531. [Google Scholar] [CrossRef]
  10. Mojica, M.; Alonso, J.A.; Méndez, F. Synthesis of fullerenes. J. Phys. Org. Chem. 2013, 26, 526–539. [Google Scholar] [CrossRef]
  11. Silva, R.A.L.; de Brito, S.F.; Machado, D.F.S.; Carvalho-Silva, V.H.; de Oliveira, H.C.B.; Ribeiro, L. The influence of the configuration of the (C70)2 dimer on its rovibrational spectroscopic properties: A theoretical survey. J. Mol. Model. 2018, 24, 235. [Google Scholar] [CrossRef] [PubMed]
  12. Mahdavifar, Z.; Nomresaz, Z.; Shakerzadeh, E. Hetero-fullerenes C59M (M = B, Al, Ga, Ge, N, P, As) for sulfur dioxide gas sensing: Computational approach. Chem. Phys. 2020, 530, 110606. [Google Scholar] [CrossRef]
  13. Heidari, M.; Janjanpour, N.; Vakili, M.; Daneshmehr, S.; Jalalierad, K.; Alipour, F. Study of the Ionization Potential, Electron Affinity and HOMO-LUMO Gaps in the Smal Fullerene Nanostructures. Chem. Rev. Lett. 2018, 1, 45–48. [Google Scholar]
  14. Motlagh, N.M.; Rouhani, M.; Mirjafary, Z. Aminated C20 fullerene as a promising nanosensor for detection of A-234 nerve agent. Comput. Theor. Chem. 2020, 1186, 112907. [Google Scholar] [CrossRef]
  15. Metin, T.; Parlak, C.; Alver, Ö.; Tepe, M. Hydrogen storage and electronic properties of C20, C15M5 and H2@C15M5 (M = Al, Si, Ga, Ge) nanoclusters. J. Mol. Struct. 2021, 1247, 131272. [Google Scholar] [CrossRef]
  16. Khan, A.A.; Ahmad, I.; Ahmad, R. Influence of electric field on CO2 removal by P-doped C60-fullerene: A DFT study. Chem. Phys. Lett. 2020, 742, 137155. [Google Scholar] [CrossRef]
  17. Stankevich, I.V.; Nikerov, M.V.; Bochvar, D.A. The Structural Chemistry of Crystalline Carbon: Geometry, Stability, and Electronic Spectrum. Russ. Chem. Rev. 1984, 53, 640–655. [Google Scholar] [CrossRef]
  18. Prinzbach, H.; Weller, A.; Landenberger, P.; Wahl, F.; Wörth, J.; Scott, L.T.; Gelmont, M.; Olevano, D.; Issendorff, B.V. Gas-phase production and photoelectron spectroscopy of the smallest fullerene, C20. Nature 2000, 407, 60–63. [Google Scholar] [CrossRef]
  19. Guo, B.C.; Kerns, K.P.; Castleman, A.W. Ti8C12+-metallo-carbohedrenes: A new class of molecular clusters? Science 1992, 255, 1411–1413. [Google Scholar] [CrossRef]
  20. Slanina, Z.; Adamowicz, L. MNDO study of charged complexes of dodecahedron-shaped C20 with Li. J. Mol. Struct. THEOCHEM 1993, 281, 33–37. [Google Scholar] [CrossRef]
  21. Magoulas, K.; Tassios, D. Thermophysical properties of n-Alkanes from C1 to C20 and their prediction for higher ones. Fluid Phase Equilib. 1990, 56, 119–140. [Google Scholar] [CrossRef]
  22. Deng, H.T.; Guo, B.C.; Kerns, K.P.; Castleman, A.W. Formation and stability of metallocarbohedrenes: TixMyC12(x + y = 8, M = Nb, Ta, Y, and Si). Int. J. Mass Spectrom. Ion Process. 1994, 138, 275–281. [Google Scholar] [CrossRef]
  23. Hassanpour, A.; Nezhad, P.D.K.; Hosseinian, A.; Ebadi, A.; Ahmadi, S.; Ebrahimiasl, S. Characterization of IR spectroscopy, APT charge, ESP maps, and AIM analysis of C20 and its C20-nAln heterofullerene analogous (n = 1–5) using DFT. J. Phys. Org. Chem. 2021, 34, e4198. [Google Scholar] [CrossRef]
  24. Huda, M.N.; Ray, A.K. Evolution of SiC nanocluster from carbon fullerene: A density functional theoretic study. Chem. Phys. Lett. 2008, 457, 124–129. [Google Scholar] [CrossRef]
  25. Salem, M.A.; Katin, K.P.; Kaya, S.; Kochaev, A.I.; Maslov, M.M. Interaction of dopants and functional groups adsorbed on the carbon fullerenes: Computational study. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 124, 114319. [Google Scholar] [CrossRef]
  26. Rad, A.S.; Aghaei, S.M.; Aali, E.; Peyravi, M. Study on the electronic structure of Cr- and Ni-doped fullerenes upon adsorption of adenine: A comprehensive DFT calculation. Diam. Relat. Mater. 2017, 77, 116–121. [Google Scholar] [CrossRef]
  27. Bai, H.; Wu, Y.; Qiao, W.; Ji, Y. Theoretical investigations on the geometrical structures, energies, and electronic properties of the heterofullerenes made of the smallest fullerene. Fuller. Nanotub. Carbon Nanostruct. 2015, 23, 399–405. [Google Scholar] [CrossRef]
  28. Koohi, M.; Soleimani Amiri, S.; Shariati, M. Silicon impacts on structure, stability and aromaticity of C20-nSin heterofullerenes (n = 1–10): A density functional perspective. J. Mol. Struct. 2017, 1127, 522–531. [Google Scholar] [CrossRef]
  29. Ajeel, F.N.; Mohammed, M.H.; Khudhair, A.M. Tuning the electronic properties of the fullerene C20 cage via silicon impurities. Russ. J. Phys. Chem. B 2017, 11, 850–858. [Google Scholar] [CrossRef]
  30. Alver, Ö.; Parlak, C.; Umar, Y.; Ramasami, P. DFT/QTAIM analysis of favipiravir adsorption on pristine and silicon doped C20 fullerenes. Main Group Met. Chem. 2019, 42, 143–149. [Google Scholar] [CrossRef]
  31. Rouhani, M. Theoretical investigation about the adsorption of the Sarin nerve agent on C20 fullerene and its boron-doped derivative. Dispos. Am. 2018, 15, 39–46. [Google Scholar] [CrossRef]
  32. Ahmadi, R.; Jalali Sarvestani, M.R. Adsorption of proline amino acid on the surface of fullerene (C20) and boron nitride cage (B12N12): A comprehensive DFT study. Q. J. Iran. Chem. Commun. 2019, 7, 344–351. [Google Scholar] [CrossRef]
  33. Vessally, E.; Siadati, S.A.; Hosseinian, A.; Edjlali, L. Selective sensing of ozone and the chemically active gaseous species of the troposphere by using the C20 fullerene and graphene segment. Talanta 2017, 162, 505–510. [Google Scholar] [CrossRef] [PubMed]
  34. Carratu, M.; Chen, Q.; Cotti, G.; Hazucha, M.; Jantunen, M.; Lahmann, E.; Lauriola, P.; Mathieu-Nolf, M.; Pankow, D.; Penney, D. Environmental Health Criteria 213 Carbon Monoxide; World Health Organisation: Geneva, Switzerland, 1999; pp. 1–101. [Google Scholar]
  35. Léonard, N.G.; Dhaoui, R.; Chantarojsiri, T.; Yang, J.Y. Electric Fields in Catalysis: From Enzymes to Molecular Catalysts. ACS Catal. 2021, 11, 10923–10932. [Google Scholar] [CrossRef]
  36. Yu, S.; Vermeeren, P.; Hamlin, T.A.; Bickelhaupt, F.M. How Oriented External Electric Fields Modulate Reactivity. Chem.— Eur. J. 2021, 27, 5683–5693. [Google Scholar] [CrossRef]
  37. Sun, K.-B.; Zhang, S.-H.; Ren, F.-D.; Hao, Y.-P.; Ba, S.H. Theoretical prediction of the trigger linkage, cage strain, and explosive sensitivity of CL-20 in the external electric fields. J. Mol. Model. 2021, 27, 85. [Google Scholar] [CrossRef]
  38. Li, Y.; Chen, J.; Zhao, C. Influence of external electric field on the electronic structure and optical properties of pyrite. RSC Adv. 2017, 7, 56676–56681. [Google Scholar] [CrossRef] [Green Version]
  39. Khan, A.A.; Ahmad, R.; Ahmad, I.; Su, X. Selective adsorption of CO2 from gas mixture by P-decorated C24N24 fullerene assisted by an electric field: A DFT approach. J. Mol. Graph. Model. 2021, 103, 107806. [Google Scholar] [CrossRef]
  40. Esrafili, M.D. Electric field assisted activation of CO2 over P-doped graphene: A DFT study. J. Mol. Graph. Model. 2019, 90, 192–198. [Google Scholar] [CrossRef]
  41. Song, N.; Wang, Y.; Gao, H.; Jiang, W.; Zhang, J.; Xu, B.; Sun, Q.; Jia, Y. Electric field improved hydrogen storage of Ca-decorated monolayer MoS2. Phys. Lett. Sect. A Gen. At. Solid State Phys. 2015, 379, 815–819. [Google Scholar] [CrossRef]
  42. Liu, H.; Lee, J.Y. Electric field effects on the adsorption of CO on a graphene nanodot and the healing mechanism of a vacancy in a graphene nanodot. J. Phys. Chem. C 2012, 116, 3034–3041. [Google Scholar] [CrossRef]
  43. Sun, Q.; Zhou, J.; Wang, Q.; Jena, P. Electric field enhanced hydrogen storage on polarizable materials substrates. ACS Natl. Meet. B. Abstr. 2010, 107, 2801–2806. [Google Scholar]
  44. Baggio, A.R.; Machado, D.F.S.; Carvalho-Silva, V.H.; Paterno, L.G.; de Oliveira, H.C.B.; Kummel, A.C.; Trogler, W.C.; Sanyal, B.; Eriksson, O.; Puglia, C.; et al. Rovibrational spectroscopic constants of the interaction between ammonia and metallo-phthalocyanines: A theoretical protocol for ammonia sensor design. Phys. Chem. Chem. Phys. 2017, 19, 10843–10853. [Google Scholar] [CrossRef]
  45. Ader, M.; Thomazo, C.; Sansjofre, P.; Busigny, V.; Papineau, D.; Laffont, R.; Cartigny, P.; Halverson, G.P. Interpretation of the nitrogen isotopic composition of Precambrian sedimentary rocks: Assumptions and perspectives. Chem. Geol. 2016, 429, 93–110. [Google Scholar] [CrossRef] [Green Version]
  46. Nibe, M.; Svanborg, M.; Lemos-silva, R.; Da, D.A.; Filho, S.; Jørgensen, B.; Albrektsen, O.; Kjelstrup-hansen, J. Work function difference of naphthyl end-capped oligothiophene in different crystal alignments studied by Kelvin probe force microscopy. Org. Electron. 2021, 89, 106060. [Google Scholar] [CrossRef]
  47. Nemati-Kande, E.; Abbasi, M.; Doust Mohammadi, M. DFT, QTAIM and NBO Investigation of the Interaction of Rare Gases with Pristine and Decorated Boron Nitride Nanotube. ChemistrySelect 2018, 3, 9833–9840. [Google Scholar] [CrossRef]
  48. Agheli, Z.; Pordel, M.; Beyramabadi, S.A. Synthesis, characterization, optical properties, computational characterizations, QTAIM analysis and cyclic voltammetry of new organic dyes for dye-sensitized solar cells. J. Mol. Struct. 2020, 1202, 127228. [Google Scholar] [CrossRef]
  49. Acharya, C.K.; Turner, C.H. Effect of an electric field on the adsorption of metal clusters on boron-doped carbon surfaces. J. Phys. Chem. C 2007, 111, 14804–14812. [Google Scholar] [CrossRef]
  50. Boys, S.; Bernardi, F. The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  51. Bader, R.F.W. A quantum theory of molecular structure and its applications. Chem. Rev. 1991, 91, 893–928. [Google Scholar] [CrossRef]
  52. Oliveira, B.G.; Araújo, R.C.M.U.; Ramos, M.N. A topologia molecular QTAIM e a descrição mecânico-quântica de ligações de hidrogênio e ligações de di-hidrogênio. Quim. Nova 2010, 33, 1155–1162. [Google Scholar] [CrossRef]
  53. Bader, R.F.W.; Essén, H. The characterization of atomic interactions. J. Chem. Phys. 1984, 80, 1943. [Google Scholar] [CrossRef]
  54. Contreras-García, J.; Johnson, E.R.; Keinan, S.; Chaudret, R.; Piquemal, J.; Beratan, D.N.; Yang, W. NCIPLOT: A Program for Plotting Noncovalent Interaction Regions. J. Chem. Theory Comput. 2011, 7, 625–632. [Google Scholar] [CrossRef] [PubMed]
  55. Johnson, E.R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A.J.; Yang, W. Revealing noncovalent interactions. J. Am. Chem. Soc. 2010, 132, 6498–6506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  57. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  58. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  59. Javan, M.B.; Ganji, M.D. Theoretical investigation on the encapsulation of atomic hydrogen into heterofullerene nanocages. Curr. Appl. Phys. 2013, 13, 1525–1531. [Google Scholar] [CrossRef]
  60. Shariatinia, Z.; Shahidi, S. A DFT study on the physical adsorption of cyclophosphamide derivatives on the surface of fullerene C60 nanocage. J. Mol. Graph. Model. 2014, 52, 71–81. [Google Scholar] [CrossRef]
  61. Silva, R.A.L.; da Silva Filho, D.A.; Moberg, M.E.; Pappenfus, T.M.; Janzen, D.E. Halogen Interactions in Halogenated Oxindoles: Crystallographic and Computational Investigations of Intermolecular Interactions. Molecules 2021, 26, 5487. [Google Scholar] [CrossRef]
  62. Bai, H.; Ji, W.; Liu, X.; Wang, L.; Yuan, N.; Ji, Y. Doping the buckminsterfullerene by substitution: Density functional theory studies of C59X (X = B, N, Al, Si, P, Ga, Ge, and As). J. Chem. 2012, 2013, 571709. [Google Scholar] [CrossRef] [Green Version]
  63. Simeon, T.M.; Yanov, I.; Leszczynski, J. Ab initio quantum chemical studies of fullerene molecules with substitutes C59X [X = Si, Ge, Sn], C59X[X = B, Al, Ga, In], and C59X+ [X = N, P, As, Sb]. Int. J. Quantum Chem. 2005, 105, 429–436. [Google Scholar] [CrossRef]
  64. Paul, D.; Deb, J.; Bhattacharya, B.; Sarkar, U. Density functional theory study of pristine and transition metal doped fullerene. AIP Conf. Proc. 2017, 1832, 1–4. [Google Scholar] [CrossRef]
  65. Dheivamalar, S.; Sugi, L. Density functional theory (DFT) investigations on doped fullerene with heteroatom substitution. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 151, 687–695. [Google Scholar] [CrossRef]
  66. Aihara, J.I. Reduced HOMO-LUMO Gap as an Index of Kinetic Stability for Polycyclic Aromatic Hydrocarbons. J. Phys. Chem. A 1999, 103, 7487–7495. [Google Scholar] [CrossRef]
  67. Etminan, N.; Yoosefian, M.; Raissi, H.; Hakimi, M. Solvent effects on the stability and the electronic properties of histidine/Pd-doped single-walled carbon nanotube biosensor. J. Mol. Liq. 2016, 214, 313–318. [Google Scholar] [CrossRef]
  68. Behjatmanesh-Ardakani, R.; Pourroustaei-Ardakani, F.; Taghdiri, M.; Kotena, Z.M. DFT-B3LYP study of interactions between host biphenyl-1-aza-18-crown-6 ether derivatives and guest Cd2+: NBO, NEDA, and QTAIM analyses. J. Mol. Model. 2016, 22, 149. [Google Scholar] [CrossRef] [PubMed]
  69. Saadat, K.; Tavakol, H. An exceptional functionalization of doped fullerene observed via theoretical studies on the interactions of sulfur-doped fullerenes with halogens and halides. RSC Adv. 2015, 5, 55227–55237. [Google Scholar] [CrossRef]
  70. Tariq, S.; Khalid, M.; Raza, A.R.; Rubab, S.L.; de Alcântara Morais, S.F.; Khan, M.U.; Tahir, M.N.; Braga, A.A.C. Experimental and computational investigations of new indole derivatives: A combined spectroscopic, SC-XRD, DFT/TD-DFT and QTAIM analysis. J. Mol. Struct. 2020, 1207, 127803. [Google Scholar] [CrossRef]
  71. Grabowski, S.J. What Is the Covalency of Hydrogen Bonding? Chem. Rev. 2011, 111, 2597–2625. [Google Scholar] [CrossRef]
  72. Lipkowski, P.; Grabowski, S.J.; Robinson, T.L.; Leszczynski, J. Properties of the C−H···H Dihydrogen Bond: An ab Initio and Topological Analysis. J. Phys. Chem. A 2004, 108, 10865–10872. [Google Scholar] [CrossRef]
  73. Bianchi, R.; Gervasio, G.; Marabello, D. Experimental electron density analysis of Mn2(CO)10: Metal-metal and metal-ligand bond characterization. Inorg. Chem. 2000, 39, 2360–2366. [Google Scholar] [CrossRef] [PubMed]
  74. Kumar, P.S.V.; Raghavendra, V.; Subramanian, V. Bader’s Theory of Atoms in Molecules (AIM) and its Applications to Chemical Bonding. J. Chem. Sci. 2016, 128, 1527–1536. [Google Scholar] [CrossRef]
  75. Matta, C.F. Hydrogen-hydrogen bonding: The non-electrostatic limit of closed-shell interaction between two hydro. In Hydrogen Bonding—New Insights; Grabowski, S., Ed.; Springer: Łódz, Poland, 2006; pp. 337–375. ISBN 1402048521. [Google Scholar]
  76. Berryman, V.E.J.; Whalley, Z.J.; Shephard, J.J.; Ochiai, T.; Price, A.N.; Arnold, P.L.; Parsons, S.; Kaltsoyannis, N. Computational analysis of M-O covalency in M(OC6H5)4 (M = Ti, Zr, Hf, Ce, Th, U). Dalt. Trans. 2019, 48, 2939–2947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Ziółkowski, M.; Grabowski, S.J.; Leszczynski, J. Cooperativity in hydrogen-bonded interactions: Ab initio and “atoms in molecules” analyses. J. Phys. Chem. A 2006, 110, 6514–6521. [Google Scholar] [CrossRef] [PubMed]
  78. Berryman, V.E.J.; Shephard, J.J.; Ochiai, T.; Price, A.N.; Arnold, P.L.; Parsons, S.; Kaltsoyannis, N. Quantum chemical topology and natural bond orbital analysis of M-O covalency in M(OC6H5)4 (M = Ti, Zr, Hf, Ce, Th, Pa, U, Np). Phys. Chem. Chem. Phys. 2020, 22, 16804–16812. [Google Scholar] [CrossRef] [PubMed]
  79. Astakhov, A.V.; Chernyshev, V.M. Molecular structure of 3-amino[1,2,4]triazolo-[4,3-a]pyrimidin-5-one in various tautomeric forms: Investigation by DFT and Qtaim methods. Chem. Heterocycl. Compd. 2014, 50, 319–326. [Google Scholar] [CrossRef]
  80. Baryshnikov, G.V.; Minaev, B.F.; Minaeva, V.A.; Baryshnikova, A.T. Structure and intramolecular stabilization of geometric isomers of Bi- and trithiazolidine-4-ones and their methyl derivatives: A DFT and QTAIM study. J. Struct. Chem. 2012, 53, 428–435. [Google Scholar] [CrossRef]
  81. Baryshnikov, G.V.; Minaev, B.F.; Korop, A.A.; Minaeva, V.A.; Gusev, A.N. Structure of zinc complexes with 3-(pyridin-2-yl)-5-(arylideneiminophenyl)-1H-1,2,4-triazoles in different tautomeric forms: DFT and QTAIM study. Russ. J. Inorg. Chem. 2013, 58, 928–934. [Google Scholar] [CrossRef]
  82. Ayarde-Henríquez, L.; Guerra, C.; Duque-Noreña, M.; Rincón, E.; Pérez, P.; Chamorro, E. Are There Only Fold Catastrophes in the Diels-Alder Reaction between Ethylene and 1,3-Butadiene? J. Phys. Chem. A 2021, 125, 5152–5165. [Google Scholar] [CrossRef]
  83. Bader, R.F.W. Atoms in Molecules: A Quantum Theory; Clarendon Press: Oxford, UK, 1990; ISBN 9780198558651. [Google Scholar]
Figure 1. Optimized geometries of the (A) C20 and (B) C19Si fullerenes as obtained at the ωB97XD/6-311+G(d,p) level of theory. (C) brings an upper view perspective of the C19Si structure. Molecular Electrostatic Potential (MEP) map of (D) C20 and (E) C19Si.
Figure 1. Optimized geometries of the (A) C20 and (B) C19Si fullerenes as obtained at the ωB97XD/6-311+G(d,p) level of theory. (C) brings an upper view perspective of the C19Si structure. Molecular Electrostatic Potential (MEP) map of (D) C20 and (E) C19Si.
Molecules 27 02968 g001
Figure 2. (A) C20 fullerene structure and (B) its corresponding Schlegel diagram. The atom indexes are presented to elucidate the Si doping positions. (C) represents the substitutions which produced the set C19Si (A) formed by: C19Si (1), C19Si (2), C19Si (4), C19Si (6), C19Si (7), C19Si (8), C19Si (11), C19Si (12), C19Si (13), C19Si (14), C19Si (15), C19Si (16), C19Si (18) and C19Si (19). (D) represents the substitutions which produced the set C19Si (B) formed by: C19Si (3), C19Si (5), C19Si (9), C19Si (10), C19Si (17) and C19Si (20). (E) In the Schlegel diagram, the blue dots represent the position of the heteroatoms that resulted in geometrical conformers with the characteristics shown in C19Si (A). The red dots indicate the positions of the heteroatoms that resulted in geometrical conformers with the characteristics shown in C19Si (B).
Figure 2. (A) C20 fullerene structure and (B) its corresponding Schlegel diagram. The atom indexes are presented to elucidate the Si doping positions. (C) represents the substitutions which produced the set C19Si (A) formed by: C19Si (1), C19Si (2), C19Si (4), C19Si (6), C19Si (7), C19Si (8), C19Si (11), C19Si (12), C19Si (13), C19Si (14), C19Si (15), C19Si (16), C19Si (18) and C19Si (19). (D) represents the substitutions which produced the set C19Si (B) formed by: C19Si (3), C19Si (5), C19Si (9), C19Si (10), C19Si (17) and C19Si (20). (E) In the Schlegel diagram, the blue dots represent the position of the heteroatoms that resulted in geometrical conformers with the characteristics shown in C19Si (A). The red dots indicate the positions of the heteroatoms that resulted in geometrical conformers with the characteristics shown in C19Si (B).
Molecules 27 02968 g002
Figure 3. Adsorption energies (in eV), E a d s , for the dimers formed between C20 fullerenes with CO, CO2 and N2 molecules calculated with the theoretical levels ωB97XD/6-31G(d) and ωB97XD/6-311+G(d,p).
Figure 3. Adsorption energies (in eV), E a d s , for the dimers formed between C20 fullerenes with CO, CO2 and N2 molecules calculated with the theoretical levels ωB97XD/6-31G(d) and ωB97XD/6-311+G(d,p).
Molecules 27 02968 g003
Figure 4. Theoretical, E a d s , and fitted, E a d s ( f i t ) , adsorption energies as a function of the electric field, F .
Figure 4. Theoretical, E a d s , and fitted, E a d s ( f i t ) , adsorption energies as a function of the electric field, F .
Molecules 27 02968 g004
Table 1. MK charges on the Si-atom, total energy, E t , (in a.u.) HOMO energy, E H , LUMO energy, E L , and HOMO-LUMO energy gap E H L , (in eV) also shown for discussion in the text. The last column presents the relative population, η in %, of each of the geometrical conformers calculated using the Boltzmann distribution at 298.15 K.
Table 1. MK charges on the Si-atom, total energy, E t , (in a.u.) HOMO energy, E H , LUMO energy, E L , and HOMO-LUMO energy gap E H L , (in eV) also shown for discussion in the text. The last column presents the relative population, η in %, of each of the geometrical conformers calculated using the Boltzmann distribution at 298.15 K.
MK E t E H E L E H L η
C19Si (1)1.033−1012.718−7.237−2.4394.798~0
C19Si (2)1.044−1012.718−7.237−2.4394.798~0
C19Si (3)−0.174−1012.777−7.765−2.0475.71816.784
C19Si (4)1.033−1012.718−7.237−2.4394.798~0
C19Si (5)−0.173−1012.777−7.763−2.0475.71616.590
C19Si (6)1.044−1012.718−7.237−2.4384.799~0
C19Si (7)1.034−1012.718−7.237−2.4364.801~0
C19Si (8)1.034−1012.718−7.237−2.4364.801~0
C19Si (9)−0.185−1012.777−7.763−2.0485.71616.555
C19Si (10)−0.185−1012.777−7.763−2.0475.71616.590
C19Si (11)1.033−1012.718−7.237−2.4394.798~0
C19Si (12)1.033−1012.718−7.237−2.4474.790~0
C19Si (13)1.045−1012.718−7.237−2.4364.801~0
C19Si (14)1.044−1012.718−7.237−2.4394.798~0
C19Si (15)1.045−1012.718−7.237−2.4364.801~0
C19Si (16)1.033−1012.718−7.237−2.4474.790~0
C19Si (17)−0.178−1012.777−7.765−2.0475.71816.784
C19Si (18)1.044−1012.718−7.237−2.4394.798~0
C19Si (19)1.044−1012.718−7.237−2.4394.798~0
C19Si (20)−0.162−1012.777−7.765−2.0445.72216.696
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lemos Silva, R.A.; Scalabrini Machado, D.F.; Nunes Rodrigues, N.M.; de Oliveira, H.C.B.; Ribeiro, L.; da Silva Filho, D.A. Harnessing Greenhouse Gases Absorption by Doped Fullerenes with Externally Oriented Electric Field. Molecules 2022, 27, 2968. https://doi.org/10.3390/molecules27092968

AMA Style

Lemos Silva RA, Scalabrini Machado DF, Nunes Rodrigues NM, de Oliveira HCB, Ribeiro L, da Silva Filho DA. Harnessing Greenhouse Gases Absorption by Doped Fullerenes with Externally Oriented Electric Field. Molecules. 2022; 27(9):2968. https://doi.org/10.3390/molecules27092968

Chicago/Turabian Style

Lemos Silva, Rodrigo A., Daniel F. Scalabrini Machado, Núbia Maria Nunes Rodrigues, Heibbe C. B. de Oliveira, Luciano Ribeiro, and Demétrio A. da Silva Filho. 2022. "Harnessing Greenhouse Gases Absorption by Doped Fullerenes with Externally Oriented Electric Field" Molecules 27, no. 9: 2968. https://doi.org/10.3390/molecules27092968

Article Metrics

Back to TopTop