Abstract
Sulfur dioxide is a toxic gas with serious environmental and health implications, making the development of selective and efficient sensors an urgent need. In this work, we investigate, using density functional theory (DFT)/B3LYP-D3/6–31G(d,p) calculations, the potential of B12N12 nanocages functionalized with Cr in different configurations (doped, decorated, and encapsulated) for application in SO2 chemical sensing. The results show that the encapsulated configuration (Cr@B12N12) exhibits the best combination of properties, including high electronic sensitivity (ΔE gap = 79.3%), moderate adsorption energy (E ads = −0.96 eV), and an appropriate recovery time (τ = 167 s), key parameters for reusable sensors under atmospheric conditions. In addition, the system demonstrates high selectivity toward interfering gases such as CO, CO2, COCl2, CH4, H2S, N2, and H2O, corroborated by molecular dynamics simulations. The data analysis suggests that Cr functionalization represents a promising strategy for SO2 sensor design, although the system’s performance remains dependent on the adsorption energy range and the experimental feasibility of metal encapsulation.
1. Introduction
In the context of modern industries, air pollution has become a severe threat to the environment and human health. Among the polluting gases, we can mention SO2, a colorless, poisonous, and hazardous gas that poses a significant risk to human health. − Obeso et al. highlighted that developing technologies capable of precisely monitoring specific air pollutants in diverse settings is essential to control emissions and ensure safe exposure limits are not exceeded. Detection and control of gases in residential and industrial environments are necessary, and the study and development of sensors for detecting toxic gases can prevent potentially serious health problems.
In recent decades, theoretical and experimental studies have been carried out to promote the development of new gas sensors and nanoscale materials for removing toxic gases from the environment. − Nanomaterials have aroused interest among research groups, motivated by their potential for rapid detection, high sensitivity, and low recovery time. − In particular, the B12N12 nanocage modified with metal atoms has received special attention for improving the sensitivity, adsorption capacity, and selectivity. ,−
The interaction of SO2 gas with various surfaces has already been investigated in theoretical studies using density functional theory (DFT). − Soltani et al. tested the doping of Al12N12 with Ga and Mg and observed that although the Mg-doped Al12N12 cluster exhibits high sensitivity to the presence of SO2 and NO2 molecules, nickel-decorated B12N12 were shown to differentiate SO2 from O3, and the B24N24 nanocage, which differentiates SO2 from CS2. Badran et al. showed that Be12O12 and Mg12O12 nanocages can be used to detect and remove H2S and SO2 gases. In a study using the Zn24 cluster, Mohammadi et al. showed that the material can act as an adsorbent for SO2, NO, and NO2 gases. Albargi et al. showed that the Cu2Zn10O12 nanocage is a promising candidate for H2S and SO2 detecting applications. Hussain et al. demonstrated that the modification in B12P12 nanocage, particularly the inclusion of zinc, enhances the SO2 adsorptive and sensitivity capacities.
The aza-macrocycle and g-C3N4, both pure and modified, emerged as promising materials for SO2 and SO3 selective detection. The g-C3N4 showed a higher sensitivity to SO2, underscoring its potential for gas detection. Rad and collaborators tested Pt-decorated graphene and showed that the system detects SO2 and O3 gases, with better electronic sensitivity. Shamim and collaborators tested T-graphene (TG), T-boron nitride (TBN), and their heterostructure (TG-TBN) toward CO, SO2, NO, and NO2 gas molecules and observed that TBN can be used as a gas sensor for SO2 and NO2 gases. An et al. and Parkar et al. tested Fe-doped and Si-doped carbon nanotubes and showed that both systems could detect and remove SO2, although Si-doped carbon nanotubes exhibited greater sensitivity. Ahmed et al. reported that phosphorus-doped T-graphene nanocapsule is a promising candidate for SO2 and O3 detection.
In addition to the studies cited above, − recently, others have been developed using chromium as a surface-modifying element. − Hou et al. tested the Cr3-doped GaSe monolayer to detect and remove Cl2, NO, and SO2 gases and observed that Cr3–GaSe monolayer, as a substrate material of disposable resistive sensors and removers (with high recovery times), has enormous potential in the area of detection and removal of these toxic gases. Shao et al. showed that Cr-doped armchair graphene nanoribbons performed more efficiently than zigzag graphene nanoribbons for SO2 detection. Yan et al. suggested that Cr- and Ti-decorated graphene systems were able to detect SO2 and SO3. Wu et al. reported that Cr-NbS2 and Mo-NbS2 systems could eliminate sulfur-containing gases (SO2, H2S, and SO3) in the atmosphere. Tang et al. using V-, Cr-, and Mn-decorated MoTe2 systems or hazardous gases adsorption observed that Cr-MoTe2 and Mn-MoTe2 were effective in detecting CH4, HCHO, SO2, and NO2 gases. In this sense, these findings motivated us to design a system by modifying the B12N12 nanocage with Cr metal, aiming for efficient SO2 adsorption and detection in the environment.
Notably, studies on Cr-modified B12N12 nanocages for SO2 adsorption and sensing are absent in the current literature. In the meantime, the objective of this study is to develop a theoretical study at the DFT level to investigate how different B12N12 nanocage modifications (doped, decorated, and encapsulated) by chromium affect the interaction with SO2 and the potential application as a material for SO2 selective detection in the environment.
2. Computational Methodology
The ground state geometry calculation of isolated B12N12 and all systems formed by Cr-modified B12N12 and its interaction with SO2 were made using the 6–311G(d,p) basis set in the ORCA 5.0 program. The B3LYP functional, which is based on the normalized gradient approximation (GGA), was used because this functional is able to describe pure and modified B12N12 very well. , To improve the description of long-range interaction, ,,, the Grimme dispersion function (B3LYP-D3) was used. The selected basis set seeks to maintain consistency with previous works − and has been extensively used in the literature for studies involving B12N12 nanocages modified with transition metals, ,− showing consistent results and lower computational costs when compared to more elaborate basis sets. Frequency analysis was also performed at the same level of theory to confirm true global minima, and no imaginary frequencies were recorded. The RMS gradient, RMS displacement, maximum gradient, and maximum displacement were 5 × 10–6 Hartree, 1 × 10–4 Hartree/Bohr, 2 × 10–3 Bohr, 3 × 10–4 Hartree/Bohr, and 4 × 10–3 Bohr, respectively.
The inclusion of Cr metal in B12N12 was done in five different configurations: doped (CrB11N12 and B12N11Cr, with the replacement of one boron atom and one nitrogen atom by one Cr atom, respectively); decorated (Cr@b64 and Cr@b66, with a Cr atom positioned above the atomic bond between the tetragonal and hexagonal rings and above the atomic bond between two hexagonal rings, respectively) and encapsulated (Cr@B12N12, in which a Cr atom is positioned inside the B12N12 nanocage). The structures were optimized with zero charge and different spin multiplicities, and the lowest energy structures were used in the analyses.
The HOMO–LUMO gap (E gap) value was calculated for all investigated structures using the equation defined as the energy difference between the frontier molecular orbitals (FMOs).
1 |
where E LUMO and E HOMO are the energies of the lowest unoccupied molecular orbital and the highest occupied molecular orbital, respectively. The cohesive energy (E coh in eq ) was used to investigate the nanocage’s stability ,
2 |
where E nanocage is the nanocage total energy (pure or Cr-modified); E B, E N, and E Cr are the B, N, and Cr atom energies, respectively; x, y, and z are the B, N, and Cr quantities in the structure, respectively, and N indicates the total number of atoms.
The ionization potential (IP), electron affinity (eA), chemical hardness (η), and chemical potential (μ) were calculated as a FMOs function, as shown in eqs , , , and : −
3 |
4 |
5 |
6 |
A comparison between the stability and reactivity of isolated B12N12 and the nanocages after interaction with Cr was made since the stability and reactivity of molecular systems can be evaluated using DFT calculations. Thus, as proposed by Parr et al., electrophilicity (ω) can be calculated by the chemical potential and chemical hardness (eq )
7 |
After modification of the B12N12 nanocage with Cr, a SO2 molecule was adsorbed on the surface of the nanocages and the adsorption energy (E ads) was calculated using eq
8 |
where E (nanocage‑SO2) is the system energy with adsorbed SO2, E (nanocage) is the energy of the pure or modified nanocage, E (SO2) is the SO2 energy, and E BSSE is the basis set superposition error (BSSE).
The charge transfer (Q) between the nanocage and SO2, which is the charge difference between adsorbed SO2 (Q (nanocage‑SO2)) and free SO2 (Q (SO2)), was calculated with eq .
9 |
To aid the adsorption analyses, S–O stretching (v S–O), Gibbs free energy change (ΔG ads), distance, and bond order (B.O.) were obtained. Data regarding the density of states (DOS) and molecular electrostatic potential (MEP) were generated with the Multiwfn and ChimeraX.
To investigate the potential of nanocages as a material for a chemoresistive sensor for SO2 detection, the electronic sensitivity of the material to the gas (ΔE gap), which is related to the electrical conductivity (σ), sensor recovery time (τ), sensitivity (S) and selectivity coefficient (κ), and highest occupied molecular orbital-lowest-unoccupied molecular orbital gap (HOMO–LUMO) (E gap) were calculated. The electrical conductivity (σ) of the sensor is experimentally dependent on the gap energy and can be calculated according to eq : ,
10 |
where A (electron/m3 K3/2) is a constant, T is the thermodynamic temperature (K), E gap is the gap energy, and k B is Boltzmann’s constant (8.62 × 10–5 eV K–1).
After determining the most sensitive nanocage that best adsorbs SO2, the recovery time (τ) was calculated using eq . −
11 |
where v 0 are attempt frequencies (1.0 × 1012, 5.2 × 1014, and 1.0 × 1016 s–1). −
The system with the best result for detecting SO2 was also subjected to interaction with other gases (CO, COCl2, CH4, H2O, N2, CO2, H2S, and N2O), to compare the selectivity of the system for adsorption of SO2 with that of other gases. For this, the sensor response (S) and selectivity coefficient (κ SO2‑int) were calculated using eqs and : −
12 |
13 |
In eq , σgas represents the conductivity of the gas adsorbed on the nanocage surface, σpure is the conductance of the isolated nanocage, and R is the resistance, which is inversely proportional to the electrical conductivity (see eq ). In eq , S SO2 is the SO2 sensitivity and S int is the sensitivity of other gases.
The SO2 sensibility and structural stabilities of the adsorption system were assessed through molecular dynamics (MD) simulations conducted for 500 ps, with a 2 fs integration time step and a dump of 500 fs at room temperature. Interatomic force uses the GFN1 Hamiltonian as implemented in the xTB software package.
3. Results and Discussion
3.1. Spin Multiplicities
Different spin multiplicities of the Cr-modified B12N12 nanocages (i.e., singlet, triplet, and quintet) were evaluated to determine the most stable configuration of each structure formed. The energy values of the different spin states, referenced to the most stable spin configuration, are presented in Table . It is possible to observe that Cr is more stable with a singlet spin state in the encapsulated system (Cr@B12N12), triplet in the doped systems (CrB11N12 and B12N11Cr), and quintet when decorated in B12N12 (Cr@b64 and Cr@b66). The formation of Cr-decorated systems is more stable when the metal assumes a high-spin multiplicity; this observation is in agreement with the results published by Arshad et al.
1. Relative Energy Values (in Hartree) for Different Cr Spin Multiplicities, Referenced to the Most Stable Cr-Modified B12N12 System.
Cr spin | CrB11N12 | B12N11Cr | Cr@b64 | Cr@b66 | Cr@B12N12 |
---|---|---|---|---|---|
1 | 0.036 | 0.035 | 0.106 | 0.114 | 0.000 |
3 | 0.000 | 0.000 | 0.044 | 0.440 | 0.012 |
5 | 0.075 | 0.030 | 0.000 | 0.000 | 0.024 |
The spin contamination analysis for Cr-functionalized B12N12 nanocage revealed varying degrees of spin state purity depending on the metal’s doping or encapsulation position. Table summarizes the obtained values for the total spin operator expectation value, , their theoretical expectations, and the corresponding deviations. All systems exhibited deviations below 10%, indicating adequately described spin states with negligible contamination. This supports the stability of the spin states and the appropriate application of the chosen functional-basis set combination for the studied systems.
2. Spin Contamination Analysis of Cr-Modified B12N12 Nanocages.
nanocage | system spin | expected | deviation | |
---|---|---|---|---|
CrB11N12 | 4 | 3.811014 | 3.750000 | 0.061014 |
B12N11Cr | 4 | 4.097592 | 3.750000 | 0.347592 |
Cr@b64 | 5 | 6.094094 | 6.000000 | 0.094094 |
Cr@b66 | 5 | 6.154875 | 6.000000 | 0.154875 |
Cr@B12N12 | 1 | 0.000000 | 0.000000 | 0.000000 |
3.2. Structural Analysis
Initially, it was observed that the B12N12 nanocage ground state geometry is symmetrical and formed by eight hexagonal rings and six tetragonal rings (see Figure ), with bond lengths of b 64 = 1.484 Å and b 66 = 1.437 Å. This symmetry confers a zero electric dipole moment to B12N12, and all data are consistent with works from the literature. − Furthermore, MEP analysis confirmed that negative charge accumulation occurs around N atoms (red region) and decreases around B atoms (blue region) in the B12N12 nanocage. For the orbitals, the HOMO is concentrated exclusively on the N atoms, while the LUMO is on the B atoms regions. DOS analysis showed that the pristine B12N12 nanocage presented an E gap = 6.88 eV, which is in good agreement with the results published by Beheshtian et al. (E gap = 6.84 eV) and by Escobedo-Morales et al. (E gap = 6.67 eV).
1.
Ground state geometry, MEP, FMOs, and DOS for the B12N12 nanocage.
The ground-state geometry of the Cr-modified nanocages can be seen in Figure . In these systems, the doped structures exhibit local structural deformation caused by the displacement of the Cr atom outward from the nanocage surface; this can be attributed to the larger atomic radius of Cr compared to B and N atoms. For the decorated Cr@b66 and Cr@b64 nanocages, it was observed that the B–N bond lengths adjacent to the Cr atom are 0.208 and 0.889 Å longer than those of pure B12N12, respectively, indicating that decoration also promotes local structural changes. In the encapsulated nanocage (Cr@B12N12), the B–N bond distances (b 66 = 1.485 Å and b 64 = 1.520 Å) are close to the values found for the unmodified B12N12 nanocage, indicating that the Cr atom is well accommodated inside the B12N12 nanocage.
2.
Ground state geometry for B12N12 nanocages modified with Cr: CrB11N12, B12N11Cr, Cr@b64, Cr@b66, and Cr@B12N12.
3.3. Energy and Stability Properties
Properties of interest of this work for the pure and modified B12N12 are presented in Table , and from this, it is possible to notice a significant increase in the dipole moment after the modification with Cr (0.51–5.58 D); this occurs due to an asymmetry in the nanocage (see Section ). This effect was less pronounced for Cr@B12N12, and is associated with a little distortion of the nanocages that promotes a smaller separation of charges.
3. Dipole Moment (DM) Energies, Ionization Potential (IP), Electron Affinity (eA), Chemical Hardness (η), Chemical Potential (μ), Electrophilicity (ω), Mulliken Charge for Cr Atom (Q Cr), and Cohesion Energy (E coh) Values for the Isolated Nanocages.
systems | DM/Debye | IP/eV | eA/eV | H/eV | μ/eV | ω/eV | QCr/|e| | Ecoh/eV |
---|---|---|---|---|---|---|---|---|
B12N12 | 0.00 | 7.63 | 0.75 | 3.44 | –4.19 | 2.56 | - | –7.401 |
CrB11N12 | 3.94 | 7.14 | 2.20 | 2.47 | –4.67 | 4.41 | 0.81 | –7.428 |
B12N11Cr | 5.17 | 6.13 | 2.37 | 1.88 | –4.25 | 4.79 | 0.53 | –7.180 |
Cr@b64 | 5.58 | 5.69 | 2.10 | 1.80 | –3.89 | 4.22 | 0.55 | –7.319 |
Cr@b66 | 5.17 | 5.47 | 2.48 | 1.50 | –3.98 | 5.29 | 0.46 | –7.308 |
Cr@B12N12 | 0.51 | 4.98 | 1.46 | 1.76 | –3.22 | 2.94 | –0.53 | –7.086 |
The results showed that the Cr atom exhibits a positive net charge in the doped and decorated structures, suggesting a preferential interaction with the regions of highest electron density of the SO2 molecule, i.e., with the oxygen atoms. In contrast, in the encapsulated Cr@B12N12 nanocage, the Cr atom acquires a negative charge, resulting in a lower electron density (i.e., more positive partial charges) on the surrounding boron atoms. This behavior also explains the interaction of the encapsulated nanocage with the negative or partially negative regions of the SO2 molecule, as will be further discussed in the next section.
To gain deeper insight into the stability and reactivity of the studied systems, it is necessary to examine the quantum chemical descriptors: chemical hardness (η), chemical potential (μ), and electrophilicity index (ω). The results show that the Cr-modified nanocages present lower η and higher ω values than pristine B12N12 (see Table ), indicating that the unmodified nanocage is thermodynamically more stable. According to Parr et al. and Pearson, systems with lower η and higher ω values are generally more reactive, supporting the conclusion that Cr-functionalized nanocages exhibit enhanced reactivity compared to the pristine B12N12. The negative cohesive energy (E coh) values confirm the thermodynamic feasibility of all studied structures, with CrB11N12 being the most stable, consistent with previous results, and, among them, the Cr@B12N12 nanocage exhibiting notable kinetic stability.
From a practical perspective, experimental studies have demonstrated the feasibility of synthesizing B12N12 nanocages via top-down approaches, such as plasma discharge and plasma-assisted chemical vapor deposition. Additionally, the synthesis of BN nanocages modified with encapsulated transition metals such as Fe, Y, Ag, and La has already been achieved, as reported by Oku et al. − Therefore, based on the present theoretical results and existing experimental evidence, the synthesis of the Cr@B12N12 nanocage at both laboratory and industrial scales appears feasible, opening the path for experimental studies involving toxic gases and indicating promising prospects for real-world applications.
3.4. SO2 Adsorption
In this stage, the adsorption of a single SO2 molecule was evaluated, forming the following systems: B12N12–SO2, CrB11N12–SO2, B12N11Cr–SO2, Cr@b64–SO2, Cr@b66–SO2, and Cr@B12N12–SO2. The corresponding ground-state geometries are shown in Figure . It was observed that the most favorable interaction occurs through the oxygen atom of SO2, in agreement with previous studies. ,, The adsorption sites exhibiting the strongest interactions are the positively charged Cr atoms in the CrB11N12, B12N11Cr, Cr@b64, and Cr@b66 structures, and the positively charged B atoms in the encapsulated Cr@B12N12 structure.
3.
Ground state geometry for the systems: B12N12–SO2, CrB11N12–SO2, B12N11Cr–SO2, Cr@b64–SO2, Cr@b66–SO2, and Cr@B12N12–SO2.
For both the isolated and SO2-adsorbed systems, the HOMO energy (E H), LUMO energy (E L), HOMO–LUMO gap (E gap), and electronic sensitivity (ΔE gap) were calculated and are summarized in Table . A significant decrease in E gap was observed for the Cr-modified systems, with values ranging from 2.69 to 4.95 eV, indicating an increase in the electronic reactivity. This decrease is consistent with previous findings for: (i) B12N12 doped with Fe, Co, Ni, Cu, and Zn; (ii) encapsulated 3d, 4d, and 5d transition metals in B12N12 nanocages; and (iii) B12N12 nanocages decorated with first-row transition metals. Among the Cr-functionalized structures, the Cr@b66-decorated nanocage exhibited the most pronounced gap reduction, indicating its elevated reactivity. However, the electronic sensitivity did not correlate directly with this increased reactivity: the pristine B12N12 nanocage showed a higher sensitivity to SO2 adsorption than most Cr-modified systems. Notably, the Cr@B12N12 nanocage exhibited a sensitivity of 79.3%, highlighting its potential for SO2 detection and reinforcing the relevance of its future experimental synthesis, in line with previous transition-metal-based studies − and its demonstrated performance in N2O detection.
4. HOMO Energy (E H), LUMO Energy (E L), HOMO-LUMO Gap (E gap), and Electronic Sensitivity in Percentage (ΔE gap) Values, for Isolated and SO2 Adsorbed Systems for Pure and Modified Nanocages .
isolated |
with
SO2 adsorbed |
|||||||
---|---|---|---|---|---|---|---|---|
system | EH (eV) | EL (eV) | Egap (eV) | EH (eV) | EL (eV) | Egap (eV) | ΔE gap (%) | |
B12N12 | –7.63 | –0.75 | 6.88 | –7.50 | –4.02 | 3.48 | 49.33 | |
CrB11N12 | α | –6.83 | –2.79 | 4.04 | –6.59 | –3.68 | 2.91 | 28.0 |
β | –7.14 | –2.20 | 4.95 | –7.46 | –3.71 | 3.74 | 24.3 | |
B12N11Cr | α | –6.13 | –2.37 | 3.76 | –6.44 | –2.96 | 3.48 | 7.4 |
β | –5.56 | –2.45 | 3.12 | –6.00 | –3.70 | 2.29 | 26.5 | |
Cr@b64 | α | –5.65 | –2.38 | 3.27 | –5.93 | –2.88 | 3.05 | 6.7 |
β | –5.69 | –2.10 | 3.60 | –5.88 | –2.23 | 3.65 | 1.4 | |
Cr@b66 | α | –5.47 | –2.48 | 2.99 | –5.89 | –2.89 | 3.00 | 0.4 |
β | –4.86 | –2.17 | 2.69 | –5.89 | –2.22 | 3.68 | 36.8 | |
Cr@B12N12 | –4.98 | –1.46 | 3.53 | –4.73 | –4.00 | 0.73 | 79.3 |
α and β indicate spin-up and spin-down, respectively, for the systems that presented an open shell.
Table presents the geometric, spectroscopic, energetic, and electronic parameters used to elucidate the adsorption mechanism of SO2 gas on the surfaces of both pristine and chromium-modified B12N12 nanocages. The equilibrium distances between the SO2 molecule and the nanocage surfaces, obtained from ground-state geometries, reveal that SO2 remains farther from the pristine B12N12 nanocage (corroborated by a low bond order <0.1) and significantly closer to the Cr-modified structures, which exhibit markedly higher bond orders ranging from 0.38 to 0.79. These findings indicate that Cr modification enhances the interaction strength between the nanocage and the adsorbed molecule. This conclusion is further supported by the variations in the S–O bond lengths and S–O stretching frequencies observed in the SO2-adsorbed complexes (Nanocage–SO2) relative to the isolated SO2 molecule. For the B12N12–SO2 system, both parameters closely match those of the free SO2 molecule, confirming the physisorption character of the interaction. In contrast, the Cr-modified nanocages exhibit longer S–O bond distances and significant red shifts in the S–O stretching frequencies, indicating substantial orbital overlap and electron density redistribution (clear evidence of chemisorption). It is worth noting that this spectroscopic criterion has previously been employed by our research group to distinguish between physisorption and chemisorption in analogous systems involving modified B12N12 nanocages interacting with NO, CNCl, N2O, and CO. These precedents reinforce the validity of our current interpretation regarding the nature of the SO2 adsorption process.
5. Calculated Values of Adsorption Energy (E ads), Bond Length Cage-SO2 (d cage‑SO2 ), Bond Length SO2 (d S–O), Mulliken Charge SO2 and Cr (Q SO2 and Q Cr), Dipole Moment (DM), Stretching Frequencies (v S–O), Cage-SO2 Mayer Bond Order (B.O.), and Adsorption Gibbs Free Energy Changes (ΔG ads) for the Complexes between SO2 and Isolated Nanocages.
system | Eads/eV | dcage‑SO2 /Å | dS–O(1)/Å | dS–O(2)/Å | QSO2 /|e| | QCr/|e| | DM/Debye | vS–O/cm–1 | B.O. | ΔG ads/eV |
---|---|---|---|---|---|---|---|---|---|---|
B12N12–SO2 | –0.19 | 2.286 | 1.475 | 1.461 | 0.07 | - | 1.47 | 1129.48 | <0.1 | + 0.20 |
CrB11N12–SO2 | –1.17 | 1.869 | 1.618 | 1.618 | –0.31 | 0.82 | 1.48 | 805.10 | 0.79 | –0.91 |
B12N11Cr–SO2 | –1.79 | 2.024 | 1.572 | 1.570 | –0.36 | 0.59 | 1.98 | 922.53 | 0.55 | –1.62 |
Cr@b64–SO2 | –2.03 | 1.443 | 1.627 | 1.485 | –0.37 | 0.67 | 4.86 | 814.22 | 0.70 | –1.93 |
Cr@b66–SO2 | –2.27 | 1.907 | 1.612 | 1.489 | –0.37 | 0.66 | 5.11 | 818.68 | 0.71 | –2.18 |
Cr@B12N12–SO2 | –0.96 | 1.766 | 1.700 | 1.492 | –0.21 | –0.43 | 3.41 | 812.48 | 0.38 | –0.63 |
SO2 | - | - | 1.464 | 1.464 | 0.00 | - | 1.79 | 1337.49 | - | - |
The adsorption energy (E ads) analysis for SO2 on the nanocages indicates that the interaction between the SO2 molecule and pristine B12N12 is weak, characterized by a low E ads value of 0.21 eV, a bond order below 0.1, and a minimal charge transfer of 0.07|e|. ,, These results confirm the physisorption nature of the process, which is further supported by the positive ΔG ads, indicating a nonspontaneous adsorption under the simulated conditions. In contrast, the introduction of Cr atoms significantly enhanced the adsorption properties of the nanocages. As shown in Table , all Cr-modified systems exhibited spontaneous adsorption processes (ΔG ads < 0). However, the E ads values for the CrB11N12–SO2, B12N11Cr–SO2, Cr@b64–SO2, and Cr@b66–SO2 systems were higher than −1.0 eV, ranging between −1.17 and −2.27 eV, thus making their application as chemical sensors unfeasible, as it makes the adsorption process irreversible, with a prolonged adsorption–desorption cycle. ,, Notably, these high E ads values are confirmed by the high bond orders ranging from 0.55 to 0.79, and by the high charge transfers (greater than −0.31|e|). The E ads value of −0.96 eV for the Cr@B12N12–SO2 system, on the other hand, was shown to be adequate since the value is in the range −0.3 eV < E ads < −1.0 eV. This adequacy converges with the moderate values of charge transfer for the SO2 molecule (−0.21|e|) and of bond order between the nanocage and the SO2 gas (0.38). Furthermore, it is noteworthy that after the interaction with SO2, there was an increase in the dipole moment only for Cr@B12N12, among the modified nanocages, increasing the charge separation sensitivity, which is highly desired for electrochemical sensors.
The interaction between the modified nanocages and the SO2 molecule causes changes in E gap (see Table ), as well as in the energy and shape of the frontier molecular orbitals (HOMO and LUMO), as shown in Figure , from which it is possible to infer the chemical interaction between the nanocages with Cr and SO2. In general, the adsorption of the SO2 gas causes a slight stabilization of the HOMO and the LUMO orbitals for the CrB11N12, B12N11Cr, Cr@b64, and Cr@b66 nanocages. On the other hand, for the Cr@B12N12 nanocage, there was a weak destabilization of the HOMO (0.25 eV) and a significant stabilization of the LUMO (2.54 eV), which resulted in a greater electronic sensitivity for the Cr@B12N12 nanocage. Furthermore, from the DOS diagrams, it is possible to observe a preponderant role of the cage in the formation of frontier molecular orbitals after interaction with SO2. However, a greater participation of chromium in the formation of LUMO was observed for the Cr@B12N12–SO2 system, confirming that the Cr-encapsulated nanocage is more sensitive to the SO2 molecule. The MEPs confirm an increase in the electron density distribution in SO2 (accumulation of electronic charges), meaning that SO2 receives a certain amount of charge from the modified nanocages (negative Mulliken charges).
4.
DOS, MEP, and HOMO and LUMO data, for ground state geometry of the CrB11N12–SO2, B12N11Cr–SO2, Cr@b64–SO2, Cr@b66–SO2, and Cr@B12N12–SO2 systems.
3.5. Recovery Time
The interaction of the SO2 molecule on the Cr@B12N12 nanocage surface showed moderate adsorption energy (−0.96 eV) and high sensitivity to SO2 (79.3%); for this reason, the recovery time (τ) for the Cr@B12N12 nanocage was also evaluated. The value was approximately 167 s, a suitable recovery time for chemical sensor applications. Different recovery times depending on the temperature (T) and the attempt frequency (v 0) can be observed in Table , , and it is possible to observe that the recovery times calculated for Cr@B12N12–SO2 varied between 0.51 μs and 4.63 h for the attempt frequencies and temperatures tested. Confirming that, at 298.15 K and under visible light, the recovery time of the Cr@B12N12 nanocage presents a suitable result for applications such as the SO2 sensor, it was also possible to observe that the recovery time can be reduced with the use of ultraviolet light or with the increase in temperature.
6. Calculated Values of Recovery Time (τ) for SO2 Adsorption on Cr@B12N12 Nanocage Using Three Different Attempt Frequencies and Temperatures.
recovery
time |
||||
---|---|---|---|---|
light | v0 (s–1) | 298.15 K | 398.15 K | 498.15 K |
infrared | 1.0 × 1012 | 4.63 h | 1.41 s | 5.12 ms |
yellow | 5.2 × 1014 | 166.8 s | 14 ms | 0.05 ms |
ultraviolet | 1.0 × 1016 | 1.67 s | 0.14 ms | 0.51 μs |
A more comprehensive approach to the recovery time (τ) behavior for the Cr@B12N12–SO2 system as a function of a temperature variation from 250 to 450 K can be observed in Figure , considering the evaluated temperatures. It is possible to notice that τ tends to 10 s, for example, at 280 K (ultraviolet light), 325 K (yellow light), and 380 K (infrared light), indicating that it is possible to adjust the desorption process of a molecule on a sensitive surface. When room temperature is available, it is observed that for infrared light, the time for desorption of the SO2 molecule is greater than 4 h, which makes it impossible to apply the Cr@B12N12 nanocage as a sensor for intermittent use in the detection of SO2. In this sense, the Cr@B12N12 nanocage is suitable for application in disposable chemical sensors to rapidly detect the toxic gas SO2, taking advantage of the high electronic sensitivity and adequate recovery time using visible light and adsorption energy.
5.
Variation in recovery time of the Cr@B12N12–SO2 adsorption system as a function of the temperature and attempt frequency.
3.6. Effect of Interfering Gases
The selectivity of the Cr@B12N12 nanocage toward SO2 was further evaluated in the presence of potential interfering gases, namely, CO, COCl2, CH4, H2O, N2, CO2, H2S, and N2O. The optimized geometries of the adsorption complexes are shown in Figure .
6.
Optimized structures of the Cr@B12N12 nanocage and the adsorbed gases CO, COCl2, CH4, H2O, N2, CO2, H2S, and N2O.
Table summarizes the key parameters associated with the adsorption of each gas on the Cr@B12N12 nanocage, including adsorption energy (E ads), HOMO–LUMO energy gap (E gap), electronic sensitivity (ΔE gap), sensor response (S), selectivity coefficient (κ), and charge transfer to the gas (Q gas). The data clearly indicate that SO2 exhibits the strongest interaction with the nanocage, with the highest electronic sensitivity among all gases studied. Notably, only H2O was found to be chemisorbed, whereas CO, COCl2, CH4, N2, CO2, H2S, and N2O showed weak physisorption, with absolute E ads values below 0.3 eV, indicating limited interaction with the Cr@B12N12 surface.
7. Values of Adsorption Energy (E ads), HOMO-LUMO Gap (E gap), Electronic Sensitivity (ΔE gap), Sensitivity (S), Selectivity Coefficient (κ), and Mulliken Charge SO2 (Q SO2 ) Calculated for the Interaction of Gases CO, COCl2, CH4, H2O, N2, CO2, H2S, and N2O with Cr@B12N12 .
system | Eads/eV | Egap/eV | ΔE gap/% | S | κSO2‑int | QSO2 /|e| |
---|---|---|---|---|---|---|
Cr@B12N12–SO2 | –0.96 | 0.73 | 79.30 | 4.5 × 1023 | –0.207 | |
Cr@B12N12–CO | –0.07 | 1.97 | 44.14 | 1.4 × 1013 | 3.0 × 1010 | 0.070 |
Cr@B12N12–COCl2 | –0.19 | 2.30 | 34.96 | 2.7 × 1010 | 1.7 × 1013 | 0.088 |
Cr@B12N12–CH4 | –0.06 | 3.60 | 2.03 | 0.75 | 5.9 × 1023 | 0.015 |
Cr@B12N12–H2O | –0.67 | 3.47 | 1.64 | 2.08 | 2.1 × 1023 | 0.243 |
Cr@B12N12–N2 | –0.06 | 3.51 | 0.57 | 0.48 | 9.4 × 1023 | 0.019 |
Cr@B12N12–CO2 | –0.13 | 3.51 | 0.49 | 0.40 | 1.1 × 1024 | 0.022 |
Cr@B12N12–H2S | –0.26 | 3.52 | 0.24 | 0.18 | 2.5 × 1024 | 0.269 |
Cr@B12N12–N2O | –0.13 | 3.52 | 0.14 | 0.10 | 4.3 × 1024 | 0.040 |
Charge transfer analysis revealed that, for the Cr@B12N12–SO2 system, electron flow occurs from the nanocage to the SO2 molecule (−0.207|e|), suggesting the presence of back-donation mechanisms in the interaction, in agreement with previous studies. , Conversely, for all other interfering gases, the direction of charge transfer was predominantly from the gas molecules to the nanocage.
The sensitivity parameter (S) quantifies the change in electronic properties of the nanocage upon gas adsorption, while the selectivity coefficient (κ) reflects the nanocage’s capability to distinguish SO2 in a gas mixture. Higher values of κ indicate a better discrimination capability. The results confirm that the Cr@B12N12 nanocage exhibits excellent selectivity and sensitivity toward SO2 detection, suggesting its applicability in the development of high-performance chemical sensors for atmospheric monitoring.
To investigate the thermodynamic stability of the studied adsorption system and to reinforce the analysis of the nanocage’s selectivity toward SO2 in the presence of interfering gases, a more realistic system was subjected to a 500 ps molecular dynamics (MD) simulation. The system consisted of a Cr@B12N12 nanocage surrounded by SO2 gas and all other gases tested. The graph in Figure (left side) shows an abrupt drop in the system’s energy at around 100 ps of the MD simulation, which is related to the dispersion of CH4, N2, CO2, and N2O gases, none of which bind to the nanocage. CO, H2S, and water molecules bind to the nanocage, whereas COCl2 remains near the system but does not interact directly with the cage. In contrast, the SO2 molecule binds both to the metal center and to a neighboring boron site via its two oxygen atoms, establishing a stronger interaction with the nanocage throughout the trajectory, as illustrated in the system snapshot at the end of the MD run (Figure , right).
7.
Molecular dynamics analysis of the Cr@B12N12 nanocage with SO2, CO, COCl2, CH4, N2, CO2, H2S, and N2O gases. Trajectory graph (left) and final state of the trajectory (right).
The molecular dynamics analysis showed that interactions with higher sensitivity (such as CO ΔE gap = 44.14%) or with higher relative adsorption energy values (such as H2O E ads = −0.67 eV and H2S E ads = −0.26 eV) are possible in realistic systems. However, these do not interfere with the detection or adsorption of the SO2 gas. The MD results reaffirm the system’s stability and selectivity. The MD results, therefore, corroborate both the thermodynamic stability and selectivity of the Cr@B12N12 nanocage toward SO2, reinforcing its potential as a robust sensing material under practical operational conditions.
Finally, a comparative analysis between the Cr@12N12 nanocage and other materials reported in the literature for SO2 adsorption − is presented in Table . The Cr@B12N12 nanocage demonstrated one of the highest electronic sensitivities (ΔE gap = 79.3%) among the systems investigated, outperforming most candidates and closely followed by the MgB11N12 nanocage (ΔE gap = 75.37%) and the phosphorus-doped tetragonal graphene nanocapsule (P-doped TGC, ΔE gap = 74.19%). Despite its high sensitivity, the P-doped TGC presents a weak interaction with SO2 (E ads = −0.21 eV), which may impair selectivity in the presence of interfering gases. , Conversely, the MgB11N12 nanocage exhibits excessively strong binding (E ads = −3.39 eV), which compromises sensor reusability due to irreversible adsorption. In contrast, the Cr@B12N12 nanocage achieves an optimal balance between adsorption energy and sensitivity, combining selective SO2 detection with feasible desorption and regeneration properties. These findings highlight the Cr@B12N12 nanocage as a highly promising candidate for the design of efficient, selective, and reusable chemoresistive sensors for SO2 monitoring in atmospheric environments.
8. Comparison of Sensing Materials for the Detection of Sulfur Dioxide (SO2).
sensor | functional/basis set | Eads/eV | ΔE gap/% | refs |
---|---|---|---|---|
GaB11N12 | B3LYP/6–31G(d) | –2.80 | 67.45 | Soltani et al. |
MgB11N12 | B3LYP/6–31G(d) | –3.39 | 75.37 | Soltani et al. |
Ni-decorated B12N12 | B3LYP/6–31G(d,p) | –1.82 | 36.90 | Rad and Ayub |
B24N24 | B3LYP-D3/6–31G(d) | –0.49 | 45.87 | Ding and Gu |
Be12O12 | B3LYP-D3/6–31G(d,p) | –0.68 | 15.97 | Badran et al. |
Mg12O12 | B3LYP-D3/6–31G(d,p) | –2.11 | 2.60 | Badran et al. |
Zn24 | TPSSh/Lanl2DZ | –4.26 | 46.08 | Mohammadi et al. |
Zn12O12 | TPSSh/Lanl2DZ | –0.54 | 55.13 | Mohammadi et al. |
Cu2Zn10O12 | B3LYP/LanL2DZ | –2.64 | 4.98 | Albargi et al. |
B12P12 | B3LYP/6–31G(d,p) | –0.15 | 45.67 | Hussain et al. |
Zn–B12P12 | B3LYP/6–31G(d,p) | –1.08 | 55.97 | Hussain et al. |
aza-macrocycle | ωB97XD/6–31 + G(d,p) | –0.24 | 1.29 | Siddique et al. |
g-C3N4 | B3LYP/6–31G(d,p) | –0.28 | 7.45 | Ashiq et al. |
Pt-decorated graphene | B3LYP/6–31G(d,p) | –0.85 | 25.10 | Rad and Zareyee |
T-boron nitride (TBN) | PBE/DNP | –0.911 | 6.5 | Shamim et al. |
Fe-CNT | PBE/DNP | –1.298 | 28.40 | An et al. |
Si-CNT | PBE/DNP | –1.66 | 42.59 | Parkar et al. |
P-doped TGC | B3LYP-D3/6–31G(d) | –0.21 | 74.19 | Ahmed et al. |
Cr@B12N12 | B3LYP-D3/6–31G(d,p) | –0.96 | 79.30 | This work |
Without BSSE correction.
With BSSE correction.
Calculated value using E gap from the literature.
Literature data.
4. Conclusions
In this theoretical study, pristine and chromium-modified B12N12 nanocages (doped, decorated, and encapsulated) were evaluated as potential candidates for the selective sulfur dioxide (SO2) detection. Among the investigated systems, the Cr@B12N12 nanocage exhibited the most balanced performance, with moderate adsorption energy (E ads = −0.96 eV), high electronic sensitivity (ΔE gap = 79.3%), and an adequate recovery time at room temperature (τ = 167 s). These values indicate a favorable balance between effective detection and reversible adsorption, which are key features for reusable gas sensors.
Importantly, Cr@B12N12 showed remarkable selectivity toward SO2 in the presence of common atmospheric interfering gases, including CO, COCl2, CH4, N2O, N2, CO2, H2S, and water, confirmed by molecular dynamics tests. While H2O exhibited chemical adsorption, its electronic sensitivity (ΔE gap = 1.64%) was significantly lower than that of SO2. All other gases showed weak physisorption (E ads < 0.3 eV) and minimal gap variation, confirming SO2-specific detection.
Taken together, these findings position Cr@B12N12 as a promising platform for SO2 chemoresistive sensing applications. Nevertheless, the adsorption energy is near the upper limit for reversible detection, and the experimental synthesis of encapsulated Cr within the nanocage remains a challenge. Therefore, additional studies, particularly experimental validation and stability assessments, are necessary to support the practical implementation of this system in real-world sensor technologies.
Acknowledgments
The authors acknowledge financial support by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPESGrant No. 88887.472618/2019-00-PROCAD-AM).
The Article Processing Charge for the publication of this research was funded by the Coordenacao de Aperfeicoamento de Pessoal de Nivel Superior (CAPES), Brazil (ROR identifier: 00x0ma614).
The authors declare no competing financial interest.
§.
Date of death: June 10, 2024
References
- Stern, A. C. ; Boubel, R. W. ; Turner, D. B. ; Fox, D. L. . Fundamentals of Air Pollution, 2nd ed.; Academic Press: Warsaw, Poland, 1997. [Google Scholar]
- Xue R., Guo Y., He H., Zhang Y., Yang N., Xie G., Nie Z.. Bimetallic Phthalocyanine Monolayers as Promising Materials for Toxic H2S, SO2, and SOF2 Gas Detection: Insights from DFT Calculations. Langmuir. 2025;41:4059–4075. doi: 10.1021/acs.langmuir.4c04401. [DOI] [PubMed] [Google Scholar]
- Zhou Q., Zeng W., Chen W., Xu L., Kumar R., Umar A.. High sensitive and low-concentration sulfur dioxide (SO2) gas sensor application of heterostructure NiO-ZnO nanodisks. Sens. Actuators, B. 2019;298:126870. doi: 10.1016/j.snb.2019.126870. [DOI] [Google Scholar]
- Obeso J. L., Flores C. V., Peralta R. A., Viniegra M., Martín-Guaregua N., Huxley M. T., Solis-Ibarra D., Ibarra I. A., Ibarra I. A., Janiak C.. Metal–organic frameworks (MOFs) toward SO2 detection. Chem. Soc. Rev. 2025;54:4135–4163. doi: 10.1039/d4cs00997e. [DOI] [PubMed] [Google Scholar]
- Tang Y., Zhao Y., Liu H.. Room-temperature semiconductor gas sensors: challenges and opportunities. ACS Sens. 2022;7:3582–3597. doi: 10.1021/acssensors.2c01142. [DOI] [PubMed] [Google Scholar]
- Gond R., Barala S., Shukla P., Bassi G., Kumar S., Kumar M., Kumar M., Rawat B.. Fe2O3-Functionalized MoS2 Nanostructure Sensor for High-Sensitivity and Low-Level SO2 Detection. ACS Sens. 2025;10:3412–3422. doi: 10.1021/acssensors.4c03297. [DOI] [PubMed] [Google Scholar]
- Steinhauer S.. Gas Sensors Based on Copper Oxide Nanomaterials: A Review. Chemosensors. 2021;9:51. doi: 10.3390/chemosensors9030051. [DOI] [Google Scholar]
- Silva A. L. P., Sousa N. S., Júnior J. J. G. V.. Theoretical studies with B12N12 a toxic gas sensor: a review. J. Nanopart. Res. 2023;25:22. doi: 10.1007/s11051-023-05667-9. [DOI] [Google Scholar]
- Raad N. H., Manavizadeh N., Frank I., Nadimi E.. Gas sensing properties of a two-dimensional graphene/h-BN multi-heterostructure toward H2O, NH3 and NO2: A first principles study. Appl. Surf. Sci. 2021;565:150454. doi: 10.1016/j.apsusc.2021.150454. [DOI] [Google Scholar]
- Patel S., Patel P., Chodvadiya D., Som N. N., Jha P. K.. Adsorption performance of C12, B6N6 and Al6N6 nanoclusters towards hazardous gas molecules: A DFT investigation for gas sensing and removal application. J. Mol. Liq. 2022;352:118702. doi: 10.1016/j.molliq.2022.118702. [DOI] [Google Scholar]
- Palomino-Asencio L., García-Hernández E., Shakerzadeh E., Chigo-Anota E.. Boron-nitride nanostructures for the detection of harmful gases (CO, CO2, H2S, N2O, and SO2) Comput. Theor. Chem. 2024;1236:114613. doi: 10.1016/j.comptc.2024.114613. [DOI] [Google Scholar]
- Laraib S. R., Patel P., Chodvadiya D., Som N. N., Jha P. K.. et al. Intermolecular interaction of Al8O12 oxymetallic clusters in the detection of atmospheric pollutants: a DFT exploration of CO, CO2, H2, N2, NO, NO2, O2, and SO2, binding mechanisms. RSC Adv. 2025;15:7489–7508. doi: 10.1039/d4ra07985j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li L.. Investigating the adsorption and sensing of H2S, SO2, NO, and NO2 on transition metal atom doped C7N3 using DFT. Comput. Theor. Chem. 2025;1244:115074. doi: 10.1016/j.comptc.2025.115074. [DOI] [Google Scholar]
- Benjamin I., Louis H., Okon G. A., Qader S. W., Afahanam L. E., Fidelis C. F., Eno E. A., Ejiofor E. E., Manicum A. L. E.. Transition Metal-Decorated B12N12–X (X = Au, Cu, Ni, Os, Pt, and Zn) Nanoclusters as Biosensors for Carboplatin. ACS Omega. 2023;8:10006–10021. doi: 10.1021/acsomega.2c07250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abrar A., Ahmed Q. M., Arshad F., Shahzad N., Ayub K., Sheikh N. S., Jadoon T., Ullah F.. Enhanced sensitivity in bromochlorodifluoromethane detection: a comparative study of B12N12 and B12P12 nanocages. Adsorption. 2025;31:70. doi: 10.1007/s10450-025-00623-6. [DOI] [Google Scholar]
- Hidayat E. F., Amelia S. R., Fitriani N. D., Wardhani I. P., Pradila R., Wafi A. I., Chadiza N. M., Sudiarti T., Kusumawati Y., Muttaqien F., Ivansyah A. L.. Revealing the Role of X12Y12 Nanocages (X = B, In; Y = N, Sb) in NH3 Gas Adsorption: toward Gas Sensor Application. ACS Omega. 2025;10:3361–3374. doi: 10.1021/acsomega.4c06143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- de Sousa Sousa N., Silva A. L. P., Silva A. C. A., Júnior J. J. G. V.. Cu-modified B12N12 nanocage as a chemical sensor for nitrogen monoxide gas: a density functional theory study. J. Nanopart. Res. 2023;25:248. doi: 10.1007/s11051-023-05898-w. [DOI] [Google Scholar]
- Silva A. L. P., Júnior J. J. G. V.. MB11N12 (M = Fe–Zn) Nanocages for Cyanogen Chloride Detection: A DFT Study. J. Inorg. Organomet. Polym. Mater. 2024;34:302–312. doi: 10.1007/s10904-023-02824-4. [DOI] [Google Scholar]
- Soltani A., Raz S. G., Taghartapeh M. R., Moradi A. V., Mehrabian R. Z.. Ab initio study of the NO2 and SO2 adsorption on Al12N12 nano-cage sensitized with gallium and magnesium. Comput. Mater. Sci. 2013;79:795–803. doi: 10.1016/j.commatsci.2013.07.011. [DOI] [Google Scholar]
- Rad A. S., Ayub K.. O3 and SO2 sensing concept on extended surface of B12N12 nanocages modified by Nickel decoration: a comprehensive DFT study. Solid State Sci. 2017;69:22–30. doi: 10.1016/j.solidstatesciences.2017.05.007. [DOI] [Google Scholar]
- Ding S., Gu W.. Evaluate the potential utilization of B24N24 fullerene in the recognition of COS, H2S, SO2, and CS2 gases (environmental pollution) J. Mol. Liq. 2022;345:117041. doi: 10.1016/j.molliq.2021.117041. [DOI] [Google Scholar]
- Badran H. M., Eid K. M., Baskoutas S., Ammar H. Y.. Mg12O12 and Be12O12 nanocages as sorbents and sensors for H2S and SO2 gases: A theoretical approach. Nanomaterials. 2022;12:1757. doi: 10.3390/nano12101757. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mohammadi M. D., Louis H., Chukwu U. G., Bhowmick S., Rasaki M. E., Biskos G.. Gas-phase interaction of CO, CO2, H2S, NH3, NO, NO2, and SO2 with Zn12O12 and Zn24 atomic clusters. ACS Omega. 2023;8:20621–20633. doi: 10.1021/acsomega.3c01177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Albargi H., Ammar H. Y., Badran H. M., Algadi H., Umar A.. p-CuO/n-ZnO heterojunction structure for the selective detection of hydrogen sulphide and sulphur dioxide gases: A theoretical approach. Coatings. 2021;11:1200. doi: 10.3390/coatings11101200. [DOI] [Google Scholar]
- Hussain S., Chatha S. A. S., Hussain A. I., Hussain R., Mehboob M. Y., Muhammad S., Ahmad Z., Ayub K.. Zinc-Doped Boron Phosphide Nanocluster as Efficient Sensor for SO2 . J. Chem. 2020;2020:2629596. doi: 10.1155/2020/2629596. [DOI] [Google Scholar]
- Siddique S. A., Sajid H., Gilani M. A., Ahmed E., Arshad M., Mahmood T.. Sensing of SO3, SO2, H2S, NO2 and N2O toxic gases through aza-macrocycle via DFT calculations. Comput. Theor. Chem. 2022;1209:113606. doi: 10.1016/j.comptc.2022.113606. [DOI] [Google Scholar]
- Ashiq M., Shehzad R. A., Iqbal J., Ayub K.. Sensing applications of graphitic carbon nitride (g-C3N4) for sensing SO2 and SO3 – A DFT study. Phys. B. 2024;676:415661. doi: 10.1016/j.physb.2024.415661. [DOI] [Google Scholar]
- Rad A. S., Zareyee D.. Adsorption properties of SO2 and O3 molecules on Pt-decorated graphene: a theoretical study. Vacuum. 2016;130:113–118. doi: 10.1016/j.vacuum.2016.05.009. [DOI] [Google Scholar]
- Shamim S. U. D., Siddique A., Dash B. K., Ahmed T., Shaha S., Islam M., Piya A. A.. Exploring the Sensing Performance of T-Graphene, T-Boron Nitride, and Their Lateral Heterostructure for Toxic CO, NO, NO2, and SO2 Gas Molecules. Langmuir. 2025;41:8726–8739. doi: 10.1021/acs.langmuir.4c05324. [DOI] [PubMed] [Google Scholar]
- An L., Jia X., Liu Y.. Adsorption of SO2 molecules on Fe-doped carbon nanotubes: the first principles study. Adsorption. 2019;25:217–224. doi: 10.1007/s10450-019-00026-4. [DOI] [Google Scholar]
- Parkar P., Chaudhari A., Sonawane M. R., Nagare B. J.. SO2 sensing performance of silicon substitutional doped (8,0) carbon nanotube: A density functional theory study. Talanta Open. 2025;11:100403. doi: 10.1016/j.talo.2025.100403. [DOI] [Google Scholar]
- Ahmed M. T., Roman A. A., Roy D., Islam S., Ahmed F.. Phosphorus-doped T-graphene nanocapsule toward O3 and SO2 gas sensing: a DFT and QTAIM analysis. Sci. Rep. 2024;14:3467. doi: 10.1038/s41598-024-54110-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hou T., Zhou Q., Zeng W.. Utilizing armchair and zigzag nanoribbons for improved detection of SO2 Toxicity with graphene biosensor. Phys. B. 2025;696:416599. doi: 10.1016/j.physb.2024.416599. [DOI] [Google Scholar]
- Shao L., Chen G., Ye H., Niu H., Wu Y., Zhu Y., Ding B.. Sulfur dioxide molecule sensors based on zigzag graphene nanoribbons with and without Cr dopant. Phys. Lett. A. 2014;378:667–671. doi: 10.1016/j.physleta.2013.12.042. [DOI] [Google Scholar]
- Yan Z., Bai Y., Sun L.. Adsorption of thiophene and SOx molecules on Cr-doped and Ti-doped graphene nanosheets: a DFT study. Mater. Res. Express. 2019;6:125067. doi: 10.1088/2053-1591/ab599d. [DOI] [Google Scholar]
- Wu D., Ma A., Liu Z., Wang Z., Xu F., Fan G., Xu H.. Adsorption of sulfur-containing contaminant gases by pristine, Cr and Mo doped NbS2 monolayers based on density functional theory. Nanotechnology. 2023;34:505708. doi: 10.1088/1361-6528/acfb13. [DOI] [PubMed] [Google Scholar]
- Tang H., Xiang Y., Zhan H., Zhou Y., Kang J., Zhou Y.. Density functional theory analysis of MoTe2 decorated with transition metals (V, Cr, Mn) for hazardous gases adsorption. Comput. Theor. Chem. 2024;1238:114716. doi: 10.1016/j.comptc.2024.114716. [DOI] [Google Scholar]
- Etiese D., Amodu I. O., Edet H. O., Adeyinka A. S., Louis H.. Novel engineering of single-metals (TM: Cr, Mo, W) chemical tailoring of Pt-encapsulated fullerenes (Pt@C59TM) as dual sensors for H2CO and H2S gases: A theoretical study. Chem. Pap. 2024;78:6053–6067. doi: 10.1007/s11696-024-03525-z. [DOI] [Google Scholar]
- Tunalı Ö. F., Yuksel N., Gece G., Fellah M. F.. A DFT study of H2S adsorption and sensing on Ti, V, Cr and Sc doped graphene surfaces. Struct. Chem. 2024;35:759–775. doi: 10.1007/s11224-023-02265-2. [DOI] [Google Scholar]
- Neese F.. Software update: The ORCA program system–Version 5.0. WIREs Comput. Mol. Sci. 2022;12:e1606. doi: 10.1002/wcms.1606. [DOI] [Google Scholar]
- Farrokhpour H., Jouypazadeh H., Sohroforouzani S. V.. Interaction of different types of nanocages (Al12N12, Al12P12, B12N12, Be12O12, Mg12O12, Si12C12 and C24) with HCN and ClCN: DFT, TD-DFT, QTAIM, and NBO calculations. Mol. Phys. 2020;118:e1626506. doi: 10.1080/00268976.2019.1626506. [DOI] [Google Scholar]
- Dhali P., Hossain A. K. M. A.. Investigating the adsorption potential and sensitivity of pristine along with carbon, boron, and nitrogen substituted hetero-nanocages towards azacitidine drug. J. Mol. Liq. 2024;396:124051. doi: 10.1016/j.molliq.2024.124051. [DOI] [Google Scholar]
- Grimme S.. Density functional theory with London dispersion corrections. WIREs Comput. Mol. Sci. 2011;1:211–228. doi: 10.1002/wcms.30. [DOI] [Google Scholar]
- De Sousa Sousa N., Silva A. L. P., Silva A. C. A., Júnior J. J. G. V.. DFT Analysis of Structural, Energetic and Electronic Properties of Doped, Encapsulated, and Decorated First-Row Transition Metals on B12N12 Nanocage: Part 1. J. Inorg. Organomet. Polym. 2024;34:4082–4099. doi: 10.1007/s10904-024-03071-x. [DOI] [Google Scholar]
- De Sousa Sousa N., Silva A. L. P., Silva A. C. A., Júnior J. J. G. V.. DFT Analysis of Dynamic, Charge, and TD-DFT Properties of Doped, Encapsulated, and Decorated First-Row Transition Metals on B12N12 Nanocage: Part 2. J. Inorg. Organomet. Polym. 2024;34:3576–3588. doi: 10.1007/s10904-024-03025-3. [DOI] [Google Scholar]
- Sousa N. S., Nascimento W. C. L., Silva A. L. P., Júnior J. J. G. V.. DFT study of TM (Sc – Zn) modified B12N12 nanocage as sensor for N2O gas selective detection. Sens. Actuators, A. 2024;378:115841. doi: 10.1016/j.sna.2024.115841. [DOI] [Google Scholar]
- Asif M., Hamid M. H. S. A., Bayach I., Sheikh N. S., Ayub K.. Understanding the stabilities and non-linear optical response of transition metal endo-doped B12N12 and B12P12 nanocages; a density functional theory investigation. Phys. Scr. 2024;99:105522. doi: 10.1088/1402-4896/ad72a7. [DOI] [Google Scholar]
- Ammar H. Y., Badran H. M., Eid K. M.. TM-doped B12N12 nano-cage (TM = Mn, Fe) as a sensor for CO, NO, and NH3 gases: A DFT and TD-DFT study. Mater. Today Commun. 2020;25:101681. doi: 10.1016/j.mtcomm.2020.101681. [DOI] [Google Scholar]
- Hossain M. R., Hasan M. M., Nishat M., Noor-E-Ashrafi, Ahmed F., Ferdous T., Hossain M. A.. DFT and QTAIM investigations of the adsorption of chlormethine anticancer drug on the exterior surface of pristine and transition metal functionalized boron nitride fullerene. J. Mol. Liq. 2021;323:114627. doi: 10.1016/j.molliq.2020.114627. [DOI] [Google Scholar]
- Arshad Y., Asghar M., Yar M., Bibi T., Ayub K.. Transition Metal Doped Boron Nitride Nanocages as High Performance Nonlinear Optical Materials: A DFT Study. J. Inorg. Organomet. Polym. 2023;33:943–955. doi: 10.1007/s10904-023-02546-7. [DOI] [Google Scholar]
- Silva A. L. P., Júnior J. J. G. V.. Density Functional Theory Study of Cu-Modified B12N12 Nanocage as a Chemical Sensor for Carbon Monoxide Gas. Inorg. Chem. 2023;62:1926–1934. doi: 10.1021/acs.inorgchem.2c01621. [DOI] [PubMed] [Google Scholar]
- Baei M. T., Taghartapeh M. R., Lemeski E. T., Soltani A.. A computational study of adenine, uracil, and cytosine adsorption upon AlN and BN nano-cages. Phys. B. 2014;444:6–13. doi: 10.1016/j.physb.2014.03.013. [DOI] [Google Scholar]
- Koopmans T.. Über die Zuordnung von Wellenfunktionen und Eigenwerten zu den Einzelnen Elektronen Eines Atoms. Physica. 1934;1:104–113. doi: 10.1016/S0031-8914(34)90011-2. [DOI] [Google Scholar]
- Pearson R. G.. Absolute electronegativity and hardness correlated with molecular orbital theory. Proc. Natl. Acad. Sci. U.S.A. 1986;83:8440–8441. doi: 10.1073/pnas.83.22.8440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parr R. G., Szentpaly L. V., Liu S.. Electrophilicity index. J. Am. Chem. Soc. 1999;121:1922–1924. doi: 10.1021/ja983494x. [DOI] [Google Scholar]
- Pearson R. G.. The principle of maximum hardness. Acc. Chem. Res. 1993;26:250–255. doi: 10.1021/ar00029a004. [DOI] [Google Scholar]
- Lu T., Chen F.. Multiwfn: a multifunctional wavefunction analyzer. J. Comput. Chem. 2012;33:580–592. doi: 10.1002/jcc.22885. [DOI] [PubMed] [Google Scholar]
- Pettersen E. F., Goddard T. D., Huang C. C., Meng E. C., Couch G. S., Croll T. I., Morris J. H., Ferrin T. E.. UCSF ChimeraX: Structure visualization for researchers, educators, and developers. Protein Sci. 2021;30:70–82. doi: 10.1002/pro.3943. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Patterson, J. D. ; Bailey, B. C. . Semiconductors in Solid-State Physics. In Introduction to the Theory; Springer: Berlin, 2007; pp 293–351. [Google Scholar]
- Pineda-Reyes A. M., Herrera-Rivera M. R., Rojas-Chávez H., Cruz-Martínez H., Medina D. I.. Recent Advances in ZnO-Based Carbon Monoxide Sensors: Role of Doping. Sensors. 2021;21:4425. doi: 10.3390/s21134425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cui H., Zhang X., Zhang G., Tang J.. Pd-doped MoS2 monolayer: a promising candidate for DGA in transformer oil based on DFT method. Appl. Surf. Sci. 2019;470:1035–1042. doi: 10.1016/j.apsusc.2018.11.230. [DOI] [Google Scholar]
- Redondo A., Zeiri Y., Lowand J. J., Goddard W. A.. Application of transition state theory to desorption from solid surfaces: Ammonia on Ni(111) J. Chem. Phys. 1983;79:6410–6415. doi: 10.1063/1.445748. [DOI] [Google Scholar]
- Zahedi E., Seif A., Ahmadi T. S.. Structural and Electronic Properties of Ammonia Adsorption on the C30B15N15 Heterofullerene: A Density Functional Theory Study. J. Comput. Theor. Nanosci. 2011;8:2159–2165. doi: 10.1166/jctn.2011.1938. [DOI] [Google Scholar]
- Moghadami R., Vessally E., Babazadeh M., Es’haghi M., Bekhradnia A.. Electronic and work function-based sensors for acetylsalicylic acid based on the AlN and BN nanoclusters: DFT studies. J. Clust. Sci. 2019;30:151–159. doi: 10.1007/s10876-018-1466-3. [DOI] [Google Scholar]
- Silva A. L. P., Silva A. C. A., Navis C. N., Júnior J. J. G. V.. Theoretical study of putrescine and X12Y12 (X = Al, B and Y = N, P) nanocage interactions. J. Nanopart. Res. 2021;23:108. doi: 10.1007/s11051-021-05211-7. [DOI] [Google Scholar]
- Bano A., Zacherle T., Grieshammer S., Martin M.. An ab initio study of sensing applications of MoB2 monolayer: a potential gas sensor. Phys. Chem. Chem. Phys. 2019;21:4633–4640. doi: 10.1039/C8CP07038E. [DOI] [PubMed] [Google Scholar]
- Fan G., Wang X., Tu X., Xu H., Wang Q., Chu X.. Density functional theory study of Cu-doped BNNT as highly sensitive and selective gas sensor for carbon monoxide. Nanotechnology. 2021;32:075502. doi: 10.1088/1361-6528/abc57a. [DOI] [PubMed] [Google Scholar]
- Zhang X., Lei Y., Wu X., Hu W.. Experimental sensing and density functional theory study of H2S and SOF2 adsorption on Au-modified graphene. Adv. Sci. 2015;2:1500101. doi: 10.1002/advs.201500101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma S., Li D., Rao X., Xia X., Su Y., Lu Y.. Pd-doped h-BN monolayer: a promising gas scavenger for SF6 insulation devices. Adsorption. 2020;26:619–626. doi: 10.1007/s10450-020-00226-3. [DOI] [Google Scholar]
- Grimme S., Bannwarth C., Shushkov P.. A robust and accurate tight-binding quantum chemical method for structures, vibrational frequencies, and noncovalent interactions of large molecular systems parametrized for all spd-block elements (Z = 1–86) J. Chem. Theory Comput. 2017;13:1989–2009. doi: 10.1021/acs.jctc.7b00118. [DOI] [PubMed] [Google Scholar]
- Goel, S. ; Masunov, A. E. . Pairwise Spin-Contamination Correction Method and DFT Study of MnH and H2 Dissociation Curves. In Lecture Notes in Computer Science; Springer: Berlin, 2009; pp 141–150. [Google Scholar]
- Beheshtian J., Bagheri Z., Kamfroozi M., Ahmadi A.. Toxic CO detection by B12N12 nanocluster. Microelectron. J. 2011;42:1400–1403. doi: 10.1016/j.mejo.2011.10.010. [DOI] [Google Scholar]
- Silva A. L. P., Silva A. C. A., Júnior J. J. G. V.. Putrescine adsorption on pristine and Cu-decorated B12N12 nanocages: a density functional theory study. Comput. Theor. Chem. 2022;1215:113836. doi: 10.1016/j.comptc.2022.113836. [DOI] [Google Scholar]
- Escobedo-Morales A., Tepech-Carrillo L., Bautista-Hernández A., Camacho-García J. H., Cortes-Arriagada D., Chigo-Anota E.. Effect of Chemical Order in the Structural Stability and Physicochemical Properties of B12N12 Fullerenes. Sci. Rep. 2019;9:16521. doi: 10.1038/s41598-019-52981-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silva A. L. P., Júnior J. J. G. V.. Carbon monoxide interaction with B12N12 nanocage with and without an external electric field: a DFT study. J. Nanopart. Res. 2022;24(1):1. doi: 10.1007/s11051-021-05382-3. [DOI] [Google Scholar]
- Ghosh D. C., Biswas R.. Theoretical Calculation of Absolute Radii of Atoms and Ions. Part 1. The Atomic Radii. Int. J. Mol. Sci. 2002;3:87–113. doi: 10.3390/i3020087. [DOI] [Google Scholar]
- Zhao Z., Li Z., Wang Q.. Structures, electronic and magnetic properties of transition metal atoms encapsulated in B12N12 cage. Chem. Phys. Lett. 2020;739:136922. doi: 10.1016/j.cplett.2019.136922. [DOI] [Google Scholar]
- Oku T., Nishiwaki A., Narita I.. Formation and atomic structure of B12N12 nanocage clusters studied by mass spectrometry and cluster calculation. Sci. Technol. Adv. Mater. 2004;5:635–638. doi: 10.1016/j.stam.2004.03.017. [DOI] [Google Scholar]
- Oku T., Hirano T., Kuno M., Kusunose T., Niihara K., Suganuma K.. Synthesis, atomic structures and properties of carbon and boron nitride fullerene materials. Mater. Sci. Eng., B. 2000;74:206–217. doi: 10.1016/S0921-5107(99)00563-2. [DOI] [Google Scholar]
- Oku T., Kuno M., Kitahara H., Narita I.. Formation, atomic structures and properties of boron nitride and carbon nanocage fullerene materials. Int. J. Inorg. Mater. 2001;3:597–612. doi: 10.1016/S1466-6049(01)00169-6. [DOI] [Google Scholar]
- Oku T., Narita I., Nishiwaki A., Koi N.. Atomic structures, electronic states and hydrogen storage of boron nitride nanocage clusters, nanotubes and nanohorns. Defect Diffus. Forum. 2004;226-228:113–140. doi: 10.4028/www.scientific.net/DDF.226-228.113. [DOI] [Google Scholar]
- Baei M. T., Bagheri Z., Peyghan A. A.. Transition metal atom adsorptions on a boron nitride nanocage. Struct. Chem. 2013;24:1039–1044. doi: 10.1007/s11224-012-0132-x. [DOI] [Google Scholar]
- Peng S., Cho K., Qi P., Dai H.. Ab initio study of CNT NO2 gas sensor. Chem. Phys. Lett. 2004;387:271–276. doi: 10.1016/j.cplett.2004.02.026. [DOI] [Google Scholar]
- Kaewmaraya T., Ngamwongwan L., Moontragoon P., Jarernboon W., Singh D., Ahuja R., Karton A., Hussain T.. Novel green phosphorene as a superior chemical gas sensing material. J. Hazard. Mater. 2021;401:123340. doi: 10.1016/j.jhazmat.2020.123340. [DOI] [PubMed] [Google Scholar]
- Liu X., Cheng S., Liu H., Hu S., Zhang D., Ning H.. A Survey on Gas Sensing Technology. Sensors. 2012;12:9635–9665. doi: 10.3390/s120709635. [DOI] [PMC free article] [PubMed] [Google Scholar]