Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

First theoretical framework of Al9N9 and B9N9 nanorings for unveiling their unique detection and sensing potential for SF6 decomposition gases (H2S, SO2, SOF2, and SO2F2): toward real-time gas sensing in high-voltage power systems

Hafiz Ali Rizwana, Muhammad Usman Khan*a, Abida Anwara, Sarah Alharthibc and Mohammed A. Aminb
aDepartment of Chemistry, University of Okara, Okara-56300, Pakistan. E-mail: usman.chemistry@gmail.com; usmankhan@uo.edu.pk
bDepartment of Chemistry, College of Science, Taif University, Taif 21944, Saudi Arabia
cResearch Center of Basic Sciences, Engineering and High Altitude, Taif University, Taif 21944, Saudi Arabia

Received 11th May 2025 , Accepted 5th June 2025

First published on 12th June 2025


Abstract

Sulfur hexafluoride (SF6) is widely used as an insulating gas in high-voltage electrical equipment due to its excellent dielectric properties. However, its decomposition under electrical discharges can generate toxic and corrosive byproducts such as H2S, SO2, SOF2, and SO2F2, which pose serious threats to insulation integrity and the reliability of power systems. Rapid and accurate sensing and detection of these decomposition products is thus critical for fault diagnosis and preventative maintenance. Despite various experimental advances, the development of efficient, sensitive, and real-time nanomaterial-based gas sensors remains a challenge. In this study, we systematically investigate the sensing capabilities of Al9N9 (AlN) and B9N9 (BN) nanorings for SF6 decomposition gases using density functional theory (DFT) with the B3LYP-D3/6-31G(d,p) method. Different key electronic and structural evaluations including adsorption energy (Eads) measurements, energy gap (Eg) determinations, natural bond orbital (NBO), density of states (DOS), thermodynamic studies, atom in molecules (AIM), non-covalent interactions (NCI) and sensing mechanism were carried out to assess the sensing performance. The adsorption of these gases on AlN nanoring results in higher adsorption energies ranging from −8.690 kcal mol−1 to −38.221 kcal mol−1 while these gases are weakly adsorbed on BN nanorings (−7.041 to −7.855 kcal mol−1). The reduction of the energy gap is observed after the adsorption of SF6 decomposed gases on both rings. The most notable reduction is observed after the adsorption of SO2 on AlN (1.103 eV) and BN (2.883 eV) nanorings. The study demonstrated that SO2 showed maximum sensitivity on BN nanorings (0.9797), accompanied by a substantial work function increase of 36.715% which confirmed BN as the most reactive material for SO2 detection. The adsorption of SF6 on AlN and BN nanorings produced fast recovery times, which shows their potential for real-time sensor applications, with the increase in temperature further decreasing the recovery time. Both AlN and BN nanorings showed better detection performance for SO2, while AlN nanorings proved more efficient for SOF2 and SO2F2 detection because of their superior electrical conductivity, better charge transfer, and quicker recovery times. These findings recommend the integration of Al9N9 and B9N9 nanorings in advanced gas sensor technologies for real-time, reliable detection of SF6 decomposition gases, crucial for enhancing the safety and efficiency of high-voltage power systems.


1. Introduction

Sulphur hexafluoride (SF6) is a widely used gas in the electrical power sector due to its high dielectric strength, thermal stability, and ability to quench electrical arcs.1–4 As a dielectric medium in gas-insulated switchgear (GIS), circuit breakers, transformers, and high-voltage power transmissions, it guarantees safe operation and efficiency.5–7 SF6 is capable of breaking down at high electrical discharges and at partial breakdown conditions to yield several toxic and corrosive byproducts such as H2S, SO2, SOF2, and SO2F2.7,8 These decomposition products are very dangerous to electrical personnel because when the decomposition products corrode components of the insulation of a variety of equipment, degradation is experienced, which in turn reduces the life and reliability of the equipment.9–11 Therefore, the early and accurate identification of these gases is essential to monitor insulation points and avoid equipment defects and risks to work and life.12

In recent years, nanomaterials have attracted extensive attention in the field of gas sensing attributable to their exceptional physicochemical properties, including large surface area, adjustable electronic structures, and excellent molecule adsorption capacity.13–15 These composite structures provided the structure with abundant active sites for gas adsorption, which improved the capacity and selectivity of sensors.16,17 These unique features may be useful for optical devices,18,19 catalysis,20 sensing materials (sensors),21 adsorption processes,22 medicinal therapies,23 and electronic devices24 since Kroto et al. synthesized the C60 carbon fullerene.25,26

In this context, carbonaceous nanomaterials including the recent graphene, carbon nanotubes (CNT), and fullerene derivatives are being used for gas-sensing properties due to the high electron mobilities and high surface-to-volume ratios.27,28 Nevertheless, pure CNT is not very useful for gas sensing since it has a weak reaction and is not very sensitive to gas molecules.29,30 The ability of pure graphene to detect some gas molecules, as H2, CO2, NO, NO2, CO, and NH3, is similarly diminished due to its very low sensitivity to these molecules.31,32 Graphene quantum dots and graphene nanoribbons are two other carbon allotropes that have been studied for gas detection. However, the findings also indicate that these materials exhibit poor interactions with gas molecules in their pure forms.33,34 The previous discussion makes it clear that pure carbon-based nanomaterials are not suited for developing better nano-sensors because they interact poorly with gas molecules.35

Additionally, several group III nitrides have been studied both experimentally and theoretically.36,37 However, boron and aluminium nitrides stand out for their exceptional physical and chemical characteristics, making them the favoured options. Interactions between NH3 molecules and aluminium nitride and boron nitride nanotubes are more substantial than those between CNTs and NH3 molecules.38–40 Also, B12N12 fullerene is great at detecting NH3 gas but isn't ideal for CO and NO gas.40 Material with Fe, Co, Ni, Cu, and Zn metals on Al12N12 nanoclusters were efficiently designed for hydrogen storage by Mehboob et al.41 After testing B12N12 fullerenes for CO sensing capabilities, Beheshtian et al. found that B12N12 might be used as a CO sensor.42

There is an immense amount of research going on for the detection of gas molecules using nanomaterials. Kartika et al. examined the adsorption behaviour, electronic properties and sensitivity of mustard chemical warfare agent (CWA) on Ca12O12 nanocages by using density functional theory (DFT), revealing that Ca12O12 nanocages may be utilised to detect mustard gas.43 The potential application of Ca12O12 nanocages as phosgene adsorption sensors was noted by Louis et al.44 One more work utilised DFT to examine the sensing characteristics of Al12Si12 after the adsorption of CH4, CO, H2, NO, and NH3 gases.45 After studying Mg12O12 nanocages coated with Zn, Hussain et al. found that they were effective in detecting COCl2 gas.46 Hussain et al. observed that Mg12O12 nanoclusters had enhanced adsorption properties for NO2 gas after gold treatment.47 Evidence from a density functional theory (DFT) investigation suggests that pure Al6N6 can detect CO, NO, and NH3 gas molecules, while B6N6 can detect just NH3 gas molecules properly.48 Sainda et al. work on gas sensing and removal using Mg9O9 nanorings based on density functional theory.49 Rahimi and Solimannejad worked on the CO2 capture and removal using B6N6 nanorings.50 Patel et al. utilized C12, B6N6 and Al6N6 nanorings for the sensing and detection of toxic gases using DFT.48 Panchal et al. used AlN and BN nanorings to investigate their sensing performance for CO, NO and NH3 hazardous gases.51 Al-Otaibi et al. used a different type of nanoring for the sensing of the resorcinol molecule.52 Al-Otaibi et al. also studied the sensing of cytosine on AlN, BN, and GaN nanorings using the solvent effect and the quantum theory of atoms in molecules.53 To the best of our knowledge, no research has been conducted on the detection of SF6 decomposed gases using AlN and BN nanorings.

Due to the need for real-time detection of SF6 decomposition byproducts and the state of the art of 3D nanomaterials in gas sensing, this work explores the adsorption of SF6 decomposed gases (H2S, SO2, SOF2 and SO2F2) on AlN and BN nanorings based on density functional theory (DFT). Specifically, we will explore the structural and electronic properties of AlN and BN nanorings, along with the adsorption energy, charge transfer, and electronic behaviour of SF6 decomposing gases on AlN and BN nanorings. We will assess the sensing capability of AlN and BN nanorings based on the change in electronic features following gas interactions. These key insights will lay the groundwork for the usability of AlN and BN nanorings as gas sensors to detect the presence of SF6 decomposition, which is important for ensuring the reliability of electrical insulation systems and for minimizing potential environmental hazards.

2. Computational methods

Density functional theory (DFT) calculations were performed using the Gaussian 09 (ref. 54) program to assess the feasibility of utilizing AlN and BN nanoring for adsorbing SF6 decomposed gases (H2S, SO2, SOF2, SO2F2) during this work. GaussView 5.0 (ref. 55) was used for the designing of gases, nanocage, and all the complexes after the adsorption process. We used the widely-used B3LYP-D3 functional, which was chosen as an appropriate approach for studying the non-covalent interactions between interacting substances, along with 6-31G(d,p) basis set to compute the necessary data. Adding Grimme's D3 dispersion correction56,57 approaches better for describing van der Waals interactions, which is important for the modelling of gas adsorption on AlN and BN nanoring. This approach predicts adsorption energy and charge transfer mechanism more reliably, resulting in a more accurate comprehensive gas-sensing analysis. To validate the structures at their local minima, frequency calculations were performed after geometry optimization. Moreover, for the interaction between gases and the pristine nanocage, the adsorption energies (Eads) were computed using eqn (1).58
 
Eads = Ecomplex − (Enanocage + Egas) + BSSE (1)
where Ecomplex is the energy of complex formed after the adsorption of gases on nanocage while Enanocage and Egas are the total energy of monomers. To further test the acquired Eads for basis set superposition error (BSSE), the counterpoise approach developed by Boys and Bernadi is used.59 Furthermore, the energy gap (Eg) of the complex was computed using HOMO–LUMO data60 to get more accurate results on the adsorption energies. The energy gap between HOMO and LUMO can be calculated using eqn (2).61
 
Eg = ELUMOEHOMO (2)

In addition to obtaining and verifying the optimized structures through vibrational frequency calculations, the models in this work were examined for molecular electrostatic potential (MEP) levels, molecular orbital distribution patterns, density of states (DOS) results, and quantities of transferred charges (QT). The distribution of HOMO and LUMO orbitals was studied using the Multiwfn57 and VMD58 software packages. The data for density of states (DOS) was extracted using Multiwfn62 and plotted using Origin63 software. Furthermore, quantum theory of atoms in molecules (QTAIM) and non-covalent interactions (NCI) investigations were used to identify the interactions between the gas molecules and the nanocages. The sensor mechanism was evaluated using various parameters, including electrical conductivity, work function, recovery time, and sensing response. The minimal energy needed to release an electron from the sensor's Fermi level to a distance effectively remote from the surface is represented as work function (Φ).64,65 The target gas may be detected when the Φ values change, which indicates a change in gate voltage that generates electrical signals.66 The work function (Φ) value can be determined using eqn (3).67

 
Φ = VEF (3)
EF in the equation is Fermi level energy, which can be calculated using eqn (4).68
 
image file: d5ra03312h-t1.tif(4)

3. Results and discussion

3.1 Optimization of geometry

Before conducting further computational analysis, geometry optimization represents a crucial step for optimizing the structure of the molecular systems and ensuring their structural stability.69 In density functional theory (DFT) studies, performing a geometric optimization of molecules and nanostructures allows one to find the most stable configuration, the closest to the minimum energy, as validation for further calculations, such as adsorption energies, charge transfer and other electronic properties.70 In this study, the density functional theory (DFT) with the B3LYP-D3 functional and 6-31G(d,p) basis set was used to individually optimize all molecular systems (H2S, SO2, SOF2, SO2F2, and AlN and BN nanorings). At the start, the H2S, SO2, SOF2, and SO2F2 molecules were placed about 2.0 Å away from the closest atoms in the AlN and BN nanorings. As a result, the interactions evolved organically during optimization, without the need for hypothetical bond constraints. The gases are placed outside the AlN and BN nanorings to prepare eight distinct systems, which were also optimized using the functional mentioned above to obtain low-energy structures. The optimized geometries of all the systems under study are given in Fig. 1.
image file: d5ra03312h-f1.tif
Fig. 1 The structure optimization of individual gases and nanorings along with all the systems after the adsorption process.

The bond lengths change from slight to significant by the adsorption of SF6 decomposition gases H2S, SO2, SOF2, and SO2F2 on AlN and BN nanorings as shown in Table 1. The bond length changes are indicative of the strength of interaction between the gas molecules and the nanorings. For AlN nanoring the largest shifts are observed in the SO2F2-out-AlN system where Al1–N18 bond bumps from 1.705 Å to 1.829 Å and N18–Al17 moves from 1.705 Å to 1.826 Å. The moderate shifts are to be found for SO2-out-AlN where N18–Al1 moves from 1.705 Å to 1.774 Å and H2S and SOF2 adsorption exhibit only minor shifts (≤0.02 Å). For BN nanoring, the bond lengths variations are not as sharp as AlN. The biggest variation is in SO2-out-BN, with the bond B3–N4 growing from 1.315 Å to 1.321 Å and the other systems (H2S, SOF2, SO2F2) only show small shifts (∼0.001–0.003 Å) in bond lengths. SO2F2 and SO2 adsorption produce considerable charge redistribution, resulting in more robust interactions and higher structural distortion in AlN. Since BN nanoring is more covalent in nature, the charge transfer interactions are less strong, resulting in smaller fluctuations of bond lengths. Due to larger fluctuation of bond lengths in AlN nanoring than that in BN nanoring, AlN nanoring shows higher gas sensitivity than the latter.

Table 1 The adsorption of SF6 decomposed gases on AlN and BN results in the variations of bond lengths after the adsorption process
Systems Connection Bond lengths (Å)
Before After
H2S-out-AlN Al1–N2 1.705 1.722
N18–Al1 1.705 1.722
SO2-out-AlN Al1–N2 1.705 1.714
N18–Al1 1.705 1.774
SOF2-out-AlN N18–Al1 1.705 1.722
Al1–N2 1.705 1.715
SO2F2-out-AlN Al1–N18 1.705 1.829
N18–Al17 1.705 1.826
H2S-out-BN B13–N14 1.315 1.314
N12–B13 1.315 1.313
SO2-out-BN B3–N4 1.315 1.321
N2–B3 1.315 1.316
SOF2-out-BN B7–N8 1.315 1.316
N6–B7 1.315 1.312
SO2F2-out-BN B9–N8 1.315 1.316
N10–B9 1.315 1.317


3.2 Adsorption of SF6 decomposed gases on AlN and BN

Herein, we report the interaction between AlN and BN nanorings with SF6 decomposition gases using density functional theory (DFT). The sensing performance of these nanorings is evaluated based on the adsorption energy, charge transfer and geometry deviations. We also study the impact of electronic effects, adsorption strength and configurations of binding to identify the optimal nanoring for selective gas detection.

One of the most important parameters to characterize how strongly and how the gas adsorbs to the surface is the adsorption energy (Eads). A higher negative adsorption energy with short interaction distance means a better interaction and is often attributed to chemisorption and a lower energy with long interaction distance is considered physisorption. The computed values of adsorption energy along with interaction distance are given in Table 2. The adsorption energy of SO2F2 is the highest at the value of −38.221 kcal mol−1 with lowest interaction distance (1.725 Å) therefore indicating a strong chemisorption. The moderate interaction of SO2 (−15.970 kcal mol−1) at the interaction distance of 1.881 Å proposes a combined character of physisorption and chemisorption. Weak adsorption is observed for H2S (−11.481 kcal mol−1), and SOF2 (−8.690 kcal mol−1), indicating that the adsorption involves physical interaction (physisorption). Adsorption energies of BN systems are however, lower than that of AlN, denoting weaker interactions. A range of −7.041 kcal mol−1 (H2S) to −7.855 kcal mol−1 (SO2F2) is found, which is consistent with typical physisorption values. That means BN nanoring would have lower adsorption energies for SF6 decomposition gases, in comparison with AlN. The calculated interaction distances for all BN cases are all greater than 3.0 Å, suggesting that adsorption is predominantly dependent on weak van der Waals force (physisorption). It was found that H2S has the lowest interaction distance (3.625 Å), which demonstrates the weak adsorption with BN nanoring.

Table 2 The calculated values of adsorption energy (Eads), distance of interaction (D), charge transfer (QT), changes in enthalpy (ΔH), and change in Gibbs free energy (ΔG)
Systems Eads (kcal mol−1) D (Å) BSSE (kcal mol−1) Eads past work QT ΔH ΔG
H2S-out-AlN −11.481 2.539 1.638 −6.69 (ref. 71) 0.205 −11.458 −1.597
SO2-out-AlN −15.970 1.881 7.073 −8.53 (ref. 48) −0.255 −22.524 −7.641
SOF2-out-AlN −8.690 2.079 5.502 −3.23 (ref. 48) 0.203 −13.315 5.143
SO2F2-out-AlN −38.221 1.725 10.858 −3.46 (ref. 51) −0.080 −47.862 −31.239
H2S-out-BN −7.041 3.625 1.581 −23.29 (ref. 51) 0.002 3.589 15.081
SO2-out-BN −7.549 3.055 5.099   0.024 −11.487 −1.391
SOF2-out-BN −7.294 3.468 7.948   0.059 −14.081 −3.901
SO2F2-out-BN −7.855 3.160 6.071   0.045 −12.755 −2.120


Moreover, the charge transfer (QT) between the adsorbed gas molecules and the nanorings can help understand information about the adsorption phenomenon, and the potential of these materials as a sensor. QT are positive, suggesting the gas molecule electron donation to the nanoring, and QT are negative, corresponding to electron transfer from the nanoring to the gas molecule. It can be calculated using eqn (5).

 
QT = Qgas in complexQgas (5)

The SO2 shows the highest degree of charge transfer, thus making it most electronically interactive with AlN as given in Table 2. Physisorption-dominated interactions are hinted by the electron donation from H2S and SOF2 to AlN. High adsorption energy of SO2F2 greatly leads to lower charge transfer, suggesting a more complex interaction mechanism. Low charge transfer values are consistent for BN systems with their physisorption on the surface for all gases. SOF2 has the most charge transfer to BN, but it is still weakly physisorbed. BN nanoring has the lower charge transfer, has intrinsic limits to its reactivity and is thus not suited for charge-based gas sensing.

3.3 Thermodynamics studies

The second law of thermodynamics states that chemical thermodynamics is the study of the heat-work connection as it pertains to chemical processes and physical state changes.72,73 Thermodynamic variables, such as the enthalpy of formation (ΔH) and Gibbs free energy (ΔG) of a system, are crucial parameters that provide insight into the feasibility and spontaneity of adsorption reactions. These parameters allow us to know if the interaction between the SF6 decomposition gases (H2S, SO2, SOF2, SO2F2) and both nanorings (AlN and BN) is thermodynamically favourable and if the adsorption process occurs spontaneously. ΔH value of enthalpy of formation indicates the type of adsorption process, whether it is exothermic or endothermic. A negative ΔH indicates that the adsorption of these molecules is exothermic, thus promoting stability as energy is released following adsorption. Positive ΔH corresponds to endothermic adsorption, which requires energy input from outside and is generally less favourable. Gibbs free energy (ΔG) is used to assess whether the adsorption process is spontaneous or not. A negative ΔG indicates that the reaction can occur spontaneously and proceeds in a thermodynamically favourable direction. In thermodynamic terms, ΔG > 0 is characteristic of a non-spontaneous reaction requiring external energy for adsorption to take place whereas ΔG equal to zero represents the equilibrium state of a reaction. Using the eqn (6)–(9), the enthalpy and Gibbs free energy values of the complexes were determined.
 
ΔH° = ΣΔfH°product − ΣΔfH°reactant (298.15 K) (6)
 
ΔfH° = Σ(εo + Hcorr)product − Σ(εo + Hcorr)reactant (298.15 K) (7)
 
ΔG° = ΣΔrG°product (298.15 K) − ΣΔrG°reactant (298.15 K) (8)
 
ΔrG° = Σ(εo + Gcorr)product − Σ(εo + Gcorr)reactant (298.15 K) (9)
here, εo stands for electronic energy, Hcorr for thermal correction to enthalpy, and Gcorr for thermal correction to Gibbs free energy. DfH° and DfG° are the standard enthalpy and Gibbs free energy of formation, respectively. Details on the system's enthalpy and Gibbs free energy are provided in Table 2, which is derived from density-functional theory (DFT) calculations conducted at a constant temperature of 298.15 K. The negative enthalpy of formation (ΔH < 0) for H2S, SO2, SOF2 and SO2F2 adsorption on AlN nanoring indicates that adsorption is an exothermic process and this type of process favours adsorption. Thus, the strongest interaction is found for SO2F2-out-AlN (ΔH = −47.86 kcal mol−1), denoting the most exothermic adsorption. The negative Gibbs free energy (ΔG < 0) indicates that the adsorption process is spontaneous in most cases, except for SOF2-out-AlN (ΔG = 5.14 kcal mol−1), which is a non-spontaneous process that would require external energy for adsorption to occur. Negative enthalpy values for BN systems (exothermic process) are consistent for SO2, SOF2, and SO2F2 systems, whereas H2S-out-BN (ΔH = 3.58 kcal mol−1) is endothermic, which suggests weak or unstable interaction. As seen for SO2, SOF2 and SO2F2 the Gibbs free energy values are less than zero (ΔG < 0) therefore all these adsorption processes are spontaneous while H2S-out-BN (ΔG = 15.08 kcal mol−1) is non-spontaneous. Stronger and more spontaneous adsorption of SF6 decomposition gases make the AlN nanoring a better candidate for sensing. Although BN nanoring is effective, but it has less adsorption for H2S, in particular.

3.4 Molecular orbital studies

Frontier Molecular Orbital (FMO) analysis is an important tool for elucidating the electronic properties and reactivity of materials. Highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) are believed to be the contributing factor to study the electronic behaviour, chemical stabilization, and sensing capability of nanomaterials. Energy gap (Eg) is an important parameter that governs the electrical conductivity and chemical reactivity of a system. Systems with small energy gap show higher reactivity, good conductivity, and sensing ability. Through FMO analysis we can evaluate the impact on the electronic properties of AlN and BN nanorings from the adsorption of SF6 decomposition gases. The decrease in energy gap following gas adsorption indicates improved electrical conductivity which makes it useful for sensor development as given in Table 3. The energy gap measurement of 3.690 eV for pure AlN reveals its moderate semiconducting properties. The adsorption of SO2 on AlN nanoring results in a deep reduction of the energy gap to 1.103 eV from its initial state, showing a −70.115% decrease and leading to improved electrical conductivity. The adsorption of SOF2 causes a minor energy gap reduction of −8.303% which points to a moderate level of electronic interaction. The slight increase in energy gap (1.578%) from SO2F2 adsorption demonstrates that AlN has only weak interaction with the nanoring which reduces its sensitivity to SO2F2 detection. The energy gap of pristine BN measures 6.267 eV which demonstrates its strong insulating properties. SO2 adsorption creates a drastic decrease in the energy gap by 2.883 eV which represents a −54.006% reduction showing strong interaction capabilities that enhance sensing functions. The adsorption of SOF2 results in a moderate −20.786% reduction of the energy gap which shows electronic structural changes. The change in energy gap after gas adsorption demonstrates AlN and BN nanorings as effective sensing materials. The change of energy gap is in Fig. 2(a) for AlN systems and Fig. 2(b) for BN systems. SO2 adsorption on AlN and BN nanorings causes the greatest reduction in the energy gap which demonstrates their outstanding sensitivity for SO2 detection. Both AlN and BN nanorings demonstrate significant responsiveness to SO2 which establishes them as prime materials for SO2 detection. Moreover, the adsorption of SO2 and SOF2 on AlN results in complete shift of LUMO orbitals on the surface of the gases due to the strong interaction between them and AlN nanoring as given in Fig. 3. Similarly, LUMO is also located on the surface of the gases after the adsorption of SO2, SOF2 and SO2F2 on BN nanoring. This redistribution of molecular orbitals after the adsorption process is indicative of electron transfer between analytes and nanorings.
Table 3 The results of HOMO, LUMO, energy gap (Eg), percentage change of energy gap (% ΔEg) and fermi energy (EF) for all the systems under study
Systems HOMO (eV) LUMO (eV) Eg % ΔEg EF
AlN −6.556 −2.866 3.690 −4.711
H2S-out-AlN −6.120 −2.629 3.492 −5.383 −4.375
SO2-out-AlN −6.401 −5.298 1.103 −70.115 −5.849
SOF2-out-AlN −5.984 −2.600 3.384 −8.303 −4.292
SO2F2-out-AlN −6.555 −2.806 3.748 1.578 −4.680
BN −7.385 −1.115 6.269 −4.250
H2S-out-BN −7.319 −1.052 6.267 −0.039 −4.186
SO2-out-BN −7.252 −4.369 2.883 −54.006 −5.810
SOF2-out-BN −7.309 −2.343 4.966 −20.786 −4.826
SO2F2-out-BN −7.355 −1.783 5.572 −11.115 −4.569



image file: d5ra03312h-f2.tif
Fig. 2 The variations of energy gap (Eg) for (a) AlN based systems and (b) BN based systems.

image file: d5ra03312h-f3.tif
Fig. 3 The distribution of molecular orbitals (HOMO and LUMO) before and after the adsorption of SF6 decomposed gases on AlN and BN nanorings.

3.5 Density of states (DOS)

The density of states (DOS) shows the number of possible energy levels for a given chemical system, making it an important statistic in solid physics. Its graph contains data that may be used to analyse the distribution of energy in molecular orbitals and the characteristics of electronic structure. Using the partial density of states (PDOS) and total density of states (TDOS), the quantum states in the studied nanostructured materials may be decomposed into the contributions from their respective atomic and molecular orbitals. The HOMO and LUMO energy states change at the Fermi level when SF6 decomposed gases interact with the AlN and BN nanorings, as seen in Fig. 4. This occurs because of the orbital overlap taking place in the complexes under study.74,75 The sensing capability of AlN and BN nanorings shifts due to changes in the Fermi level energy (EF) resulting from gas adsorption, which signals electron transfer and density of states (DOS) adjustments as given in Table 3. The Fermi level of pristine AlN at −4.711 eV serves as a baseline for comparison studies. The interaction of H2S with AlN nanorings causes the Fermi level (EF) to increase to −4.375 eV through electron donation from the gas molecule, leading to a slight enhancement in conductivity. The Fermi level drops to −5.849 eV after SO2 adsorption because the nanoring loses electrons to SO2. The Fermi level energy (EF) of pristine BN stands at −4.250 eV, which is higher than that of AlN showing that BN possesses more insulating characteristics. The adsorption of H2S leads to a slight Fermi level change to −4.186 eV, which indicates limited electron transfer. SO2 adsorption results in a significant Fermi level reduction to −5.810 eV, which replicates the effect observed on AlN. The Fermi level shows the weakest change when exposed to H2S, which reveals a poor interaction and low sensing capability with both AlN and BN nanorings. The moderate shifts in SOF2 and SO2F2 reveal weaker interactions compared to SO2. The adsorption of SO2 on AlN and BN produces the greatest downward Fermi level shift, which demonstrates their strong charge transfer properties and excellent sensing performance. SO2 demonstrates the highest detectability in both AlN and BN nanorings because its strong electron-withdrawing properties produce significant Fermi level shifts. These results are consistent with what we observed in energy gap changes after the adsorption of SO2 on both AlN and BN nanorings.
image file: d5ra03312h-f4.tif
Fig. 4 The density of states (DOS) spectra for pure AlN, BN and all the systems after the adsorption of SF6 decomposed gases.

3.6 Electrostatic potential (ESP)

Understanding how molecular systems interact is made much easier with the help of the ESP. This method determines the molecule's size, shape, and locations of positive, negative, and neutral electrostatic potential. White, blue, and red are the colours shown on the MEP surface, which correspond to neutral, positive, and negative potential areas, respectively.48,76,77 The AlN and BN nanorings have red regions in the centre and blue regions towards the outside ends, which represents their rich electron nature in the centre and electron deficient towards the ends as shown in Fig. 5. ESP results show that the adsorption of H2S and SO2F2 gases on AlN results in a very little change of colours due to the minor charge transfer taking place between them according to Mulliken charge analysis (QT). This suggests that the physisorption nature of the interaction between these gas molecules and the AlN nanoring. In contrast, a chemisorption interaction and Mulliken charge analysis are confirmed by the red area of negative electrostatic potential over the SO2 and SOF2 gas molecules and the blue region of positive electrostatic potential on the nanoring as a consequence of SO2 and SOF2 gas molecule adsorption. As a result of SO2 and SOF2 gas molecules adsorption, the system's non-uniform distribution of atom charges generates various reactive sites. The adsorption of SF6 decomposed gases on BN nanoring results in a more prominent distribution of red areas over the adsorbed gases. This suggests the electron donating nature of these gases towards the BN ring as confirmed by Mulliken charge analysis.
image file: d5ra03312h-f5.tif
Fig. 5 Electrostatic potential (ESP) analysis showing the distribution of red, blue, and white areas on the surface of AlN, BN, and all the systems.

3.7 Chemical reactivity parameters

The stability and reactivity of SF6 decomposed gases on the AlN and BN nanorings were investigated by calculating chemical reactivity parameters. The following parameters were identified by use of the B3LYP-D3/6-311(d,p) technique of density functional theory. Table 4 displays the anticipated chemical hardness (η), chemical softness (σ), chemical potential (μ), and electrophilicity (ω) values for each complex. Eqn (10) and (11) of the extended Koopman's theorem provide the values of ionization potential (IP) and electron affinity (EA), respectively.
 
IP = −EHOMO (10)
 
EA = −ELUMO (11)
chemical hardness (η), chemical softness (σ), chemical potential (μ), and electrophilicity (ω) values are computed using eqn (12)–(15) respectively.
 
image file: d5ra03312h-t2.tif(12)
 
image file: d5ra03312h-t3.tif(13)
 
image file: d5ra03312h-t4.tif(14)
 
image file: d5ra03312h-t5.tif(15)
Table 4 Chemical reactivity parameters including ionization potential (IP), electron affinity (EA), chemical hardness (η), chemical softness (σ), chemical potential (μ), and electrophilicity (ω)
Systems IP EA σ η μ ω
AlN 6.556 2.866 0.271 1.845 −4.711 6.014
H2S-out-AlN 6.120 2.629 0.286 1.746 −4.375 5.481
SO2-out-AlN 6.401 5.298 0.907 0.551 −5.849 31.026
SOF2-out-AlN 5.984 2.600 0.296 1.692 −4.292 5.443
SO2F2-out-AlN 6.555 2.806 0.267 1.874 −4.680 5.844
BN 7.385 1.115 0.160 3.135 −4.250 2.881
H2S-out-BN 7.319 1.052 0.160 3.133 −4.186 2.796
SO2-out-BN 7.252 4.369 0.347 1.442 −5.810 11.708
SOF2-out-BN 7.309 2.343 0.201 2.483 −4.826 4.690
SO2F2-out-BN 7.355 1.783 0.179 2.786 −4.569 3.747


According to data given in Table 4, there is a change in the values of chemical reactivity parameters after the adsorption process. BN exhibits greater chemical hardness (η) and stability compared to AlN based on their respective hardness values which are 3.135 eV for BN and 1.845 eV for AlN. Adsorption tends to decrease η values (with some exceptions) which results in complexes exhibiting higher reactivity compared to their pure surface counterparts. The lowest hardness for SO2-out-AlN (η = 0.551 eV) and SO2-out-BN (η = 1.442 eV) demonstrates that these interactions substantially modify the electronic stability of the nanorings. Since softness (σ) is calculated as the inverse of hardness, it demonstrates an opposite pattern of behaviour. The system SO2-out-AlN and SO2-out-BN demonstrates highest softness at 0.907 eV and 0.347 eV respectively, which denotes their increased polarizability and enhanced reactivity as compared to the remaining systems. More negative chemical potential (μ) values suggest stronger electron-attracting capability. The values −5.849 eV for SO2-out-AlN and −5.810 eV for SO2-out-BN represent the most negative μ values which demonstrate a stronger propensity to accept electrons and strong adsorption forces. Comparisons reveal only minor alterations in other complexes when matched against pure surfaces, which demonstrates intermediate interactions without severe changes in electronic structure. A system with higher ω values demonstrates stronger electrophilic characteristics, which leads to a greater electron acceptance tendency. The presence of SO2 causes AlN (ω = 31.026 eV) and BN (ω = 11.708 eV) to become strong electrophiles compared to pure AlN (ω = 6.014 eV) and BN (ω = 2.881 eV) which indicates that these SO2-treated systems attract electron-rich species more effectively. The limited increase in electrophilicity by other gases demonstrates their suitability for chemical sensing technologies. SO2-out-AlN and SO2-out-BN demonstrate major electronic changes which lead to their high softness and strong electrophilicity combined with low hardness thus creating highly reactive adsorption complexes. The softer and more reactive nature of AlN-based complexes makes AlN more suitable than BN-based complexes for adsorption-related applications. The variations between chemical hardness, chemical softness, chemical potential, and electrophilicity are given in Fig. 6.


image file: d5ra03312h-f6.tif
Fig. 6 Comparison between the different chemical reactivity parameters of all the systems including pristine AlN and BN nanorings.

3.8 Natural bond orbital (NBO) analysis

To verify the continued donor–acceptor orbital link between the adsorbed gases on the selected nanorings, we used natural bond orbital (NBO) analysis. It is an important method that shows how atomic species interact with each other and the distribution of electrons within molecular bonds.78 The complex Schrödinger equation is reduced to the simpler level of chemical bonding using the NBO approach.79 The investigation of the AlN and BN nanorings by NBO analysis reveals the delocalization of electrons and intermolecular charge transfer between the adsorbate gases. For NBO analysis, we used the B3LYP-D3/6-31G(d,p) method of density functional theory (DFT). One way to quantify the strength of the interaction between the orbitals of the donor and acceptor is by looking at the stabilization energy (E2). The stabilization energies of the most interacting donor–acceptor orbitals in the systems that were studied are shown in Table 5. To calculate stabilization/perturbation energy, the following eqn (16) is used.
 
image file: d5ra03312h-t6.tif(16)
Table 5 The most prominent interactions between donor and acceptor orbitals of SF6 decomposed gases and nanorings
Systems Donor Acceptor E2 kcal mol−1 E(j) − E(i) a.u. F(j, i) a.u.
H2S-out-AlN LP S19 σ* Al1–N2 9.38 0.69 0.073
σ S19–H20 π* Al1–N2 1.91 0.7 0.034
π Al1–N2 σ* S19–H21 0.18 0.43 0.008
SO2-out-AlN LP O20 σ* N18–S19 12.2 0.13 0.072
π Al11–N12 π* Al13–N14 17.02 0.29 0.063
σ N14–Al15 σ* N12–Al13 8.49 0.91 0.079
SOF2-out-AlN LP O20 σ* Al1–N18 9.06 0.72 0.074
σ S19–O20 π* Al1–N2 2.63 1.1 0.05
π Al1–N2 σ* S19–O20 0.48 0.43 0.013
SO2F2-out-AlN σ S19–F22 σ* N18–S19 30.11 1.06 0.16
LP O20 σ* Al17–O20 28.92 0.67 0.125
π Al13–N14 π* Al15–N16 17.18 0.31 0.066
H2S-out-BN LP S19 π* B9–N10 0.98 0.38 0.018
σ S19–H20 π* B13–N14 0.23 0.68 0.012
π B13–N14 σ* S19–H21 0.07 0.52 0.006
SO2-out-BN LP O20 π* N12–B13 1.23 0.94 0.032
σ S19–O21 π* N4–B5 0.12 1.01 0.01
π S19–O21 π* N4–B5 0.54 0.61 0.017
SOF2-out-BN LP O20 π* N8–B9 0.5 0.45 0.014
π S19–O20 π* N4–B5 0.42 0.49 0.013
SO2F2-out-BN LP O21 π* B1–N18 0.83 0.47 0.018
π N6–B7 σ* S19–F23 0.09 0.27 0.004


In the equation Ei and Ej are the diagonal elements, F(i, j) is the Fock Matrix, and q is the donor occupancy. The interactions between donor and acceptor orbitals are visualized in Fig. 7 for all the systems. SO2F2-out-AlN demonstrates the strongest interaction among studied complexes because the σ S19–F22 → σ* N18–S19 transition produces a stabilization energy of 30.11 kcal mol−1. The H2S-out-AlN complex features a key interaction between the lone pair (LP) of S19 and the anti-bonding σ orbital from the Al1–N2 bond that produces an E2 energy value of 9.38 kcal mol−1. Charge transfer between sulphur's lone pair and the anti-bonding orbital of the Al–N bond makes the Al–N bond weaker, which demonstrates moderate H2S adsorption on AlN. A LP O20 → σ* N18–S19 interaction in SO2-out-AlN delivers 12.20 kcal mol−1 of stabilization energy which demonstrates strong conjugation effects that improve the complex's electronic properties. The oxygen lone pair (LP O20) transfers charge to the σ* Al1–N18 orbital and achieves 9.06 kcal mol−1 of stabilization energy in the SOF2-out-AlN system revealing moderate interaction strength. BN-based complexes demonstrate weaker interactions in comparison to their AlN-based counterparts. SO2-out-BN achieves the maximum stabilization energy for BN complexes through charge donation from LP O20 into the π* N12–B13 orbital with an E2 energy value of 1.23 kcal mol−1. The bonding shows low adsorption capability along with restricted electronic stabilization. The H2S-out-BN complex features a lone pair (LP S19) to π* B9–N10 transition which has E2 value of 0.98 kcal mol−1 indicating minimal charge transfer. The observed stabilization energy in SOF2-out-BN and SO2F2-out-BN is even weaker than other reported values, which suggests minimal interaction between BN and the adsorbed gases. The results show AlN-based systems possess stronger charge transfer interactions than BN-based systems, which makes AlN more appropriate for applications in gas adsorption and sensing. The SO2F2-out-AlN complex emerges as the most stable system because of its high stabilization energy which marks it as an excellent option for further study in gas sensing applications.


image file: d5ra03312h-f7.tif
Fig. 7 Donor and acceptor interactions between natural bond orbitals (NBO) of AlN and BN-based systems.

3.9 Atoms in molecules (AIM) analysis

Topological research was conducted using the atoms in molecules (AIM) analysis. Microelectronic and structural investigations alone may not be enough to decipher the complex system's intermolecular interactions. After the adsorption process AIM reveals information on the interactions between SF6 decomposed gases and nanorings. This study examined various topological parameters, including total electronic density (ρ(r)), Laplacian of electron densities (∇2ρ(r)), Lagrangian kinetic energy (G(r)), total electronic energy (H(r)), potential energy (V(r)), electron localisation function (ELF), and bond ellipticity (ε) at bond critical points (BCP). Bond critical point is a saddle point between two bonded atoms, which has the maximum electron density as shown in Fig. 8. Table 6 lists the topological features of the AlN and BN nanorings and how they interact with H2S, SO2, SOF2 and SO2F2 gases at bond critical points (BCPs). Closed-shell interactions at a BCP are predicted by AIM theory when the total electron energy density (ρ) is positive. The presence of weak covalent contacts, also known as strong electrostatic interactions, occurs when ∇2ρ > 0 and H > 0, while the presence of a strong covalent bond occurs when ∇2ρ < 0 and H < 0. In contrast, a partial or medium covalent bond is indicated by H < 0 and ∇2ρ > 0. According to Table 6, all of the complexes that were tested had positive ∇2ρ values and the majority of them had positive H(r) values. Additionally, there exist a few negative H(r) values. In other words, most of the bonds in our complexes are weak covalent interactions, while there are a few medium or partial covalent connections as well. Based on the literature study, mixed character interactions are present when the ratio of V(r) to G(r) is less than 2, but more than 1.80 An indication of covalent bonding is when the value of |V(r)|/G(r) > 2 while the value of |V(r)|/G(r) < 1 indicates ionic bonding and van der Waals interactions. The ratio of V(r) to G(r) is mostly less than 1 indicating van der Waals interactions between analytes and nanorings. Bond ellipticity (ε) is a significant measure for assessing interaction stability. When the value of ellipticity (ε) is less than 1, structural stability is suggested, whereas structural instability is indicated when the value of ellipticity is greater than 1. The bond ellipticity values are mostly less than 1 for AlN based systems while BN systems are mostly unstable due to higher ellipticity values. The delocalization of electronic density is studied using electron localization function (ELF) values. A number for ELF below 0.5 indicates that the electron is delocalised, whilst a value between 0.5 and 1 indicates that the electron is localised for bonding and non-bonding. Adsorption results in complexes with values less than 0.5, as seen in Table 6, suggesting that electrons have been delocalised in studied systems.
image file: d5ra03312h-f8.tif
Fig. 8 Visual study of atom in molecules (AIM) analysis showing the important bond critical points (BCPs) between SF6 decomposed gases and nanorings.
Table 6 Different parameters of atoms in molecules (AIM) analysis performed for all the systems
BCP Connection ρ(r) 2ρ(r) G(r) V(r) H(r) V(r)/G(r) ELF ε
H2S-out-AlN
34 Al1–S19 0.032 0.088 0.027 −0.033 −0.005 1.192 0.101 0.461
[thin space (1/6-em)]
SO2-out-AlN
28 Al1–O20 0.058 0.429 0.098 −0.089 0.009 0.909 0.060 0.064
37 O21–Al17 0.058 0.429 0.098 −0.089 0.009 0.909 0.060 0.064
[thin space (1/6-em)]
SOF2-out-AlN
32 N18–F22 0.007 0.029 0.006 −0.005 0.001 0.829 0.013 1.051
35 N18–F21 0.007 0.029 0.006 −0.005 0.001 0.828 0.013 1.162
42 Al1–O20 0.034 0.194 0.046 −0.044 0.002 0.949 0.049 0.331
[thin space (1/6-em)]
SO2F2-out-AlN
30 O20–Al17 0.066 0.458 0.110 −0.106 0.004 0.962 0.075 0.024
37 S19–N18 0.227 −0.509 0.093 −0.313 −0.220 3.369 0.871 0.063
43 O21–Al1 0.066 0.458 0.110 −0.106 0.004 0.962 0.074 0.026
[thin space (1/6-em)]
H2S-out-BN
35 B7–S19 0.004 0.012 0.002 −0.002 0.001 0.696 0.016 5.243
43 S19–B17 0.004 0.012 0.002 −0.002 0.001 0.695 0.016 5.330
53 H20–N12 0.004 0.015 0.003 −0.002 0.001 0.618 0.009 2.303
[thin space (1/6-em)]
SO2-out-BN
28 N6–S19 0.007 0.022 0.004 −0.003 0.001 0.726 0.029 0.936
37 B3–O21 0.007 0.025 0.005 −0.005 0.001 0.872 0.019 3.013
44 O20–B13 0.007 0.024 0.005 −0.005 0.001 0.867 0.017 3.181
[thin space (1/6-em)]
SOF2-out-BN
30 B15–F22 0.006 0.026 0.005 −0.004 0.001 0.789 0.011 2.175
34 B3–F21 0.006 0.026 0.005 −0.004 0.001 0.790 0.011 2.165
45 S19–B9 0.004 0.017 0.003 −0.002 0.001 0.646 0.010 1.971
[thin space (1/6-em)]
SO2F2-out-BN
36 O20–N10 0.006 0.022 0.005 −0.004 0.001 0.821 0.014 4.388
56 N4–F22 0.005 0.025 0.005 −0.004 0.001 0.744 0.009 2.606
57 F23–N16 0.005 0.025 0.005 −0.004 0.001 0.745 0.009 2.534


3.10 Non-covalent interactions (NCI)

NCI analysis functions as an effective computational method for displaying and analyzing weak intermolecular forces within molecular structures. Non-covalent interactions such as hydrogen bonding, van der Waals forces and dipole–dipole interactions stabilize molecular complexes unlike covalent bonds which involve strong electron sharing.81 The Reduced Density Gradient (RDG) approach is the standard method used to carry out NCI analysis, which identifies spatial regions where non-covalent interactions happen.82 The colours on the RDG map use blue to indicate strong attractive interactions like hydrogen bonding, while green represents weak van der Waals interactions and strong repulsive interactions from steric effects appear as red. The use of NCI analysis on our systems enables enhanced understanding of the non-covalent interactions between nanorings and gas molecules and verifies their stability and reactivity. The x-axis shows sign(λ2)ρ which identifies various interaction types while the y-axis plots the Reduced Density Gradient (RDG) to detect interaction zones as shown in Fig. 9. In AlN based systems the strong blue regions (−0.020 to −0.050) provide evidence that hydrogen bonding and charge transfer interactions play major roles during adsorption. Green areas (near sign(λ2)ρ ≈ 0) of van der Waals interactions indicate that weak physisorption contributes to the stabilization of gas molecules upon the AlN surface. The red repulsive interactions demonstrate areas where steric hindrance blocks additional stabilization. The BN-based system displays weaker attractive forces because the blue regions show less intensity and fail to reach as far into negative values. The BN-based system displays a broader and stronger green region that indicates weak van der Waals interactions compared to the AlN-based system as also seen on their 3D iso-surface maps in Fig. 9. The reliance of BN-based systems on dispersion forces exceeds their dependence on strong electrostatic interactions and charge transfer mechanisms. The stronger attractive interactions from hydrogen bonding and charge transfer combined with reduced steric hindrance make AlN-based systems superior for gas adsorption. The analysis aligns with NBO findings because AlN systems demonstrate increased charge transfer stabilization while BN systems present reduced donor–acceptor interactions.
image file: d5ra03312h-f9.tif
Fig. 9 Reduced density gradient (RDG) maps and 3D iso-surfaces for the designed systems based on non-covalent interaction (NCI) studies.

3.11 Sensing mechanism

The fundamental purpose for this study is to get a better understanding of the capability of AlN and BN nanorings to detect SF6 decomposed gases. Various factors such as electrical conductivity (σ), sensing response (S), work function (Φ), and recovery time (τ) were assessed to achieve this goal.83,84
3.11.1 Electrical conductivity (σ). The gas sensor mechanism is better understood by comparing the electrical conductivity fluctuations of designed complexes before and after gas adsorption. For an electrical material to be electrically conductive, its capacity to transport electrons between the valence and conduction bands is critical.85 Eqn (17), which involves examining the changes in energy gap both before and after gas adsorption, was used to compute the electrical conductivities of the complexes.
 
image file: d5ra03312h-t7.tif(17)
herein, A represents the Richardson constant 6 × 105 A m−2, T is the working temperature (298 K), and k is Boltzmann constant (8.318 × 10−3 kJ mol−1 K−1). The adsorption of SF6 decomposed gases on nanorings results in the decrease of energy gap that increases the electrical conductivities as shown in Table 7. The variations between energy gap and electrical conductivity are given in Fig. 10(a). The conductivity of pure AlN nanoring reaches 3.77 × 1012 S m−1. The electrical conductivity of the AlN system reaches its peak at 6.36 × 1012 S m−1 when modified with SO2 to create the SO2-out-AlN complex. The SOF2-out-AlN complex shows improved conductivity with a value of 4.01 × 1012 S m−1 although it remains less conductive than the SO2-out-AlN complex. The pure BN nanoring system demonstrates a conductivity level of 2.24 × 1012 S m−1. The SO2-out-BN complex shows maximum conductivity at 4.44 × 1012 S m−1 representing a major enhancement compared to pure BN which mirrors the behaviour found in AlN systems. The conductivity of the SOF2-out-BN complex reaches 2.92 × 1012 S m−1 although it remains less impressive when compared to the SO2-out-BN complex. SO2 systems stand out as they display the highest conductivity levels among AlN and BN nanorings which positions them as top choices for high-conductivity applications. AlN-based systems typically demonstrate greater conductivity compared to BN-based systems, which suggests AlN as the superior choice for sensor applications.
Table 7 Values of electrical conductivity (σ), sensing response (S), work function (Φ), percentage change of work function (% ΔΦ) and recovery time (τ) for all the systems
Systems σ Φ % ΔΦ τ (298 K) τ (373 K) S
AlN 3.77 × 1012 4.711
H2S-out-AlN 3.93 × 1012 4.375 −7.139 1.03 × 10−10 4.05 × 10−11 0.0409
SO2-out-AlN 6.36 × 1012 5.849 24.170 6.28 × 10−10 1.72 × 10−10 0.6852
SOF2-out-AlN 4.01 × 1012 4.292 −8.895 3.33 × 10−11 1.65 × 10−11 0.0638
SO2F2-out-AlN 3.73 × 1012 4.680 −0.647 4.97 × 10−6 2.24 × 10−7 −0.0117
BN 2.24 × 1012 4.250
H2S-out-BN 2.24 × 1012 4.186 −1.508 1.71 × 10−11 9.67 × 10−12 0.0005
SO2-out-BN 4.44 × 1012 5.810 36.715 2.10 × 10−11 1.14 × 10−11 0.9797
SOF2-out-BN 2.92 × 1012 4.826 13.557 1.90 × 10−11 1.05 × 10−11 0.3006
SO2F2-out-BN 2.58 × 1012 4.569 7.513 2.38 × 10−11 1.26 × 10−11 0.1509



image file: d5ra03312h-f10.tif
Fig. 10 Comparative study of (a) energy gap and electrical conductivity and (b) work function values of all the systems.
3.11.2 Work function (Φ). The work function is defined as the minimum energy needed to remove an electron from the Fermi level (EF).86 The values of the work functions were obtained by using the eqn (3) and the changes in work function values are calculated using eqn (18).
 
image file: d5ra03312h-t8.tif(18)
here Φcomplex and Φnanoring are the work function values of complexes and pure nanorings. The molecular adsorption leads to substantial changes in work function as it demonstrates major electronic modifications to the nanoring surface as listed in Table 7. The work function measurement of pure AlN nanoring stands at 4.711 eV. The SO2-out-AlN complex stands out among modified AlN systems by achieving the highest work function at 5.849 eV, which represents a 24.170% rise as compared to pure AlN. Investigation of the SOF2-out-AlN and H2S-out-AlN systems reveals decreased Φ values of 4.292 eV (−8.895%) and 4.375 eV (−7.139%), respectively, demonstrating a reduced electron attraction potential. The SO2-out-BN complex demonstrates the highest Φ value at 5.810 eV, which represents the most substantial enhancement across all systems examined at 36.715% as compared to 4.250 eV of pure BN nanoring. The SOF2-out-BN complex demonstrates a significant rise in Φ by 4.826 eV and 13.557% while the SO2F2-out-BN system shows a moderate Φ increase of 4.569 eV and 7.513%. The H2S-out-BN complex shows a minor Φ decrease to 4.186 eV with a −1.508% change that follows the same trend found in the AlN system. SO2 complexes stand out as the top-performing systems for both AlN and BN nanorings because they demonstrate major enhancements in work function. SO2 adsorption triggers substantial electronic modifications that position these complexes as top choices for electronic and sensor applications. The comparison of function for all the systems is given in Fig. 10(b).
3.11.3 Recovery time (τ) and effect of temperature. The bare minimum of time needed for a gas sensor to return to its initial form after gas sensing is known as recovery time (τ). More research on the reusability of nanomaterials is necessary due to the significant benefit of their low cost when used for gas sensing. The hypothetical study of recovery time for the adsorption of SF6 decomposed gases on AlN and BN was conducted at two different temperatures. Eqn (19), in which Eads is related to the recovery time, was used for the calculations of recovery time.87
 
image file: d5ra03312h-t9.tif(19)

In this equation, the variables Vo is the attempt frequency (1012 s−1), Eads is the adsorption energy, K is the Boltzmann constant, and T is the working temperature (298 K and 373 K). The results of recovery time at both these temperatures are given in Table 7. H2S-out-BN achieves the fastest recovery from all systems with its shortest reaction time of 1.71 × 10−11 s. Both SOF2-out-AlN (3.33 × 10−11 s) and SO2-out-BN (2.10 × 10−11 s) systems display quick recovery times. The SO2F2-out-AlN system has the longest recovery time at 4.97 × 10−6 s which suggests slow sensor recovery and possible saturation challenges. All system recovery times decrease when temperatures rise from 298 K to 373 K and the most pronounced change is observed in SO2F2-out-AlN where RT reduces from 4.97 × 10−6 s to 2.24 × 10−7 s demonstrating that thermal energy speeds up desorption. H2S-out-BN shows minimal changes in recovery time because it already possesses very short RT at 298 K which demonstrates its natural efficiency in gas desorption even at room temperature. Gas sensors that need quick response and reset times perform best with BN-based systems since they demonstrate shorter RTs compared to AlN-based systems. The largest improvement in recovery times appears in SO2F2-out-AlN when temperatures increase.

3.11.4 Sensing response (S). The sensing response is a quantitative value that describes the degree of a surface to detect the adsorbed analyte molecules.88,89 The sensing response is related to the electrical conductivity by the eqn (20).
 
image file: d5ra03312h-t10.tif(20)

Gas sensor performance relies on sensitivity, which reflects the strength of the gas adsorption response. The maximum sensitivity of SO2-out-BN at 0.9797 demonstrates an outstanding detection capability for SO2 gas as shown in Table 7. AlN also demonstrates high sensitivity to SO2 but performs less effectively compared to BN, which makes both materials suitable for SO2 sensing with BN being the better option. The performance of SOF2-out-BN at 0.3006 and SOF2-out-AlN at 0.0638 is moderate but BN shows better results than AlN. Both nanorings show a poor performance in detecting H2S indicating that they are unsuitable for this application. SO2F2-out-AlN proves unsuitable as a detection material because of its negative sensitivity.

4. Conclusion

Our research examined how Al9N9 (AlN) and B9N9 (BN) nanorings interact with different SF6 decomposed gases such as H2S, SO2, SOF2 and SO2F2 through their adsorption behaviour and sensing abilities using density functional theory (DFT) calculations. We investigated several electronic and structural features such as adsorption energy (Eads), energy gap (Eg), natural bond orbitals (NBO), chemical reactivity parameters, non-covalent interaction (NCI) analysis and sensing mechanism such as electrical conductivity, work function, recovery time, and sensitivity. Analysis of adsorption energy values demonstrates strong binding between AlN nanoring and both SO2 and SO2F2 with adsorption energy values of −15.970 kcal mol−1 and −38.221 kcal mol−1 respectively, which indicates their high interaction stability as compared to BN based systems. Adsorption of SO2 on both AlN and BN nanorings resulted in substantial band gap reductions in SO2-out-AlN (−70.115%) and that SO2-out-BN (−54.006%) as compared to pure AlN and BN nanorings. The reduction of energy gap caused improved electrical conductivity and sensor response during energy gap analysis. The NBO analysis validated substantial charge transfer within AlN based systems as compared to BN systems which amplified their adsorption-driven electronic structure changes. The systems with adsorbed molecules exhibited enhanced softness according to chemical reactivity parameter calculations while SO2-out-BN (0.347 eV) and SO2-out-AlN (0.907 eV) demonstrated the greatest softness increase which improved their reactivity and sensing application potential. The NCI analysis demonstrated that van der Waals forces and hydrogen bonding primarily govern interactions which lead to the stabilization of adsorbed complexes. The analysis revealed that SO2 adsorption on BN nanoring (SO2-out-BN) achieved the best results by displaying high sensitivity (0.9797) and substantial work function changes (36.715% increase), demonstrating its strong suitability for gas sensing devices. The SO2-out-AlN system showed substantial changes in its electronic structure by achieving superior conductivity at 6.36 × 1012 S m−1 and a notable work function increase of 24.170%, positioning it as an excellent choice for gas sensor applications. The performance of a sensor heavily depends on recovery time, which defines how fast the sensor can revert to its original state following gas desorption. All the systems show fast recovery time with the fastest being reported for BN-based systems due to the weak adsorption of gases with BN nanoring. The raised temperature levels decreased recovery time, which validated thermal desorption as a feasible process. Our investigation identifies BN and AlN as good choices for SO2 gas sensing applications because they provide strong adsorption properties alongside substantial electronic changes and improved conductivity while maintaining ideal recovery times. The research shows that BN and AlN nanorings function as highly effective materials for the detection of SF6 decomposed gases with the superior performance of AlN nanoring based on better electrical conductivity, charge transfer and adsorption energy. This enables the use of AlN nanoring in environmental monitoring and industrial safety systems.

Data availability

The data are available from the corresponding author on reasonable request.

Author contributions

HAR contributed to the research design, writing of the original draft, investigation, validation, visualization, formal analysis, acquisition, and interpretation of data. MUK made a substantial contribution to the research design, conceptualization, methodology, project administration, investigation, data curation, supervision, review, and editing, and approval of the submitted version of the manuscript. AA had substantial contributions to the formal analysis, visualization, data curation, validation, writing, review & editing. SA had substantial contributions to the formal analysis, visualization, data curation, validation, funding, acquisition, writing, review & editing. MAA had a substantial contribution to the formal analysis, interpretation of data, validation, writing, review, and editing.

Conflicts of interest

The authors declare that they have no conflict of interest.

Acknowledgements

The authors are thankful to Taif University, Saudi Arabia, for supporting this work through project number TU-DSPP-2024-29.

References

  1. J. Chen, Q. Zhou, L. Jia, X. Cui and W. Zeng, The gas-sensing mechanism of Pt3 cluster doped SnS2 monolayer for SF6 decomposition: a DFT study, Appl. Surf. Sci., 2022, 597, 153693 CrossRef CAS.
  2. L. Xu, Y. Gui, W. Li, Q. Li and X. Chen, Gas-sensing properties of Pt n-doped WSe2 to SF6 decomposition products, J. Ind. Eng. Chem., 2021, 97, 452–459 CrossRef CAS.
  3. Z. Liu, Y. Gui, L. Xu and X. Chen, Adsorption and sensing performances of transition metal (Ag, Pd, Pt, Rh, and Ru) modified WSe2 monolayer upon SF6 decomposition gases (SOF2 and SO2F2), Appl. Surf. Sci., 2022, 581, 152365 CrossRef CAS.
  4. B. Li, Q. Zhou, R. Peng, Y. Liao and W. Zeng, Adsorption of SF6 decomposition gases (H2S, SO2, SOF2 and SO2F2) on Sc-doped MoS2 surface: a DFT study, Appl. Surf. Sci., 2021, 549, 149271 CrossRef CAS.
  5. Y. Liao, Q. Zhou, R. Peng and W. Zeng, Adsorption properties of InP3 monolayer toward SF6 decomposed gases: a DFT study, Phys. E, 2021, 130, 114689 CrossRef CAS.
  6. H. Cui, H. Zhu and P. Jia, Adsorption and sensing of SO2 and SOF2 molecule by Pt-doped HfSe2 monolayer: a first-principles study, Appl. Surf. Sci., 2020, 530, 147242 Search PubMed.
  7. H. Liu, L. Xu, Y. Gui, L. Ran and X. Chen, Adsorption properties of Ag2O–MoSe2 towards SF6 decomposed products, Vacuum, 2021, 189, 110248 CrossRef CAS.
  8. A. Dong and M. Liu, A DFT study on the adsorption properties of Ti3C2O2 MXene towards SF6 decomposition gases, Surf. Sci., 2023, 734, 122317 CrossRef CAS.
  9. M. Istad and M. Runde, Thirty-six years of service experience with a national population of gas-insulated substations, IEEE Trans. Power Delivery, 2010, 25(4), 2448–2454 Search PubMed.
  10. L. G. Christophorou, J. K. Olthoff and R. J. Van Brunt, Sulfur hexafluoride and the electric power industry, IEEE Electrical Insulation Magazine, 1997, 13(5), 20–24 Search PubMed.
  11. O. Bender, H. J. Boehmer, D. Jens and K. P. Schumacher, Using GIS to analyse long-term cultural landscape change in Southern Germany, Landscape and Urban Planning, 2005, 70(1–2), 111–125 Search PubMed.
  12. D. Chen, X. Zhang, J. Tang, Z. Cui and H. Cui, Pristine and Cu decorated hexagonal InN monolayer, a promising candidate to detect and scavenge SF6 decompositions based on first-principle study, J. Hazard. Mater., 2019, 363, 346–357 CrossRef CAS PubMed.
  13. S. Rozas, R. Alcalde, M. Atilhan and S. Aparicio, A theoretical study on the adsorption of acid gases by boron nitride-based nanomaterials, Appl. Surf. Sci., 2019, 480, 83–95 CrossRef CAS.
  14. R. Geetha Sadasivan Nair, A. K. Narayanan Nair and S. Sun, Adsorption of Gases on Fullerene-like X12Y12 (X = Be, Mg, Ca, B, Al, Ga, C; Y = C, Si, N, P, O) Nanocages, Energy Fuels, 2023, 37(18), 14053–14063 CrossRef CAS.
  15. S. Hussain, R. Hussain, M. Y. Mehboob, S. A. S. Chatha, A. I. Hussain and A. Umar, et al., Adsorption of phosgene gas on pristine and copper-decorated B12N12 nanocages: a comparative DFT study, ACS Omega, 2020, 5(13), 7641–7650 CrossRef CAS PubMed.
  16. R. G. S. Nair, A. K. N. Nair and S. Sun, Adsorption of gases on B12N12 and Al12N12 nanocages, New J. Chem., 2024, 48(18), 8093–8105 RSC.
  17. F. S. Ghoreishi, A. Mahmoudi and S. M. Dehaghi, DFT insights into the electronic properties and gas sensing performance of C19X (X = C, N, and Si) nanocages for the detection of CO2, CO, N2, and H2 gases, Diamond Relat. Mater., 2024, 148, 111348 CrossRef.
  18. J. Shi, Y. Zhu, X. Zhang, W. R. G. Baeyens and A. M. García-Campaña, Recent developments in nanomaterial optical sensors, TrAC, Trends Anal. Chem., 2004, 23(5), 351–360 CrossRef CAS.
  19. J. Wang, Y. Chen and W. J. Blau, Carbon nanotubes and nanotube composites for nonlinear optical devices, J. Mater. Chem., 2009, 19(40), 7425–7443 RSC.
  20. Z.-Y. Zhou, N. Tian, J.-T. Li, I. Broadwell and S.-G. Sun, Nanomaterials of high surface energy with exceptional properties in catalysis and energy storage, Chem. Soc. Rev., 2011, 40(7), 4167–4185 RSC.
  21. C. Zhu, G. Yang, H. Li, D. Du and Y. Lin, Electrochemical Sensors and Biosensors Based on Nanomaterials and Nanostructures, Anal. Chem., 2015, 87(1), 230–249 CrossRef CAS PubMed.
  22. K. Yang, L. Zhu and B. Xing, Adsorption of Polycyclic Aromatic Hydrocarbons by Carbon Nanomaterials, Environ. Sci. Technol., 2006, 40(6), 1855–1861 CrossRef CAS PubMed.
  23. D. F. Emerich and C. G. Thanos, Nanotechnology and medicine, Expet Opin. Biol. Ther., 2003, 3(4), 655–663 CrossRef CAS PubMed.
  24. A. Kamyshny and S. Magdassi, Conductive Nanomaterials for Printed Electronics, Small, 2014, 10(17), 3515–3535 CrossRef CAS PubMed.
  25. H. W. Kroto, J. R. Heath, S. C. O'Brien, R. F. Curl and R. E. Smalley, C60: Buckminsterfullerene, Nature, 1985, 318(6042), 162–163 CrossRef CAS.
  26. J. Liu, A. G. Rinzler, H. Dai, J. H. Hafner, R. K. Bradley and P. J. Boul, et al., Fullerene Pipes, Science, 1998, 280(5367), 1253–1256 CrossRef CAS PubMed.
  27. M. L. Terranova, S. Orlanducci and M. Rossi, Carbon Nanomaterials for Gas Adsorption, CRC Press, 2012 Search PubMed.
  28. D. J. Babu, D. Puthusseri, F. G. Kühl, S. Okeil, M. Bruns and M. Hampe, et al., SO2 gas adsorption on carbon nanomaterials: a comparative study, Beilstein J. Nanotechnol., 2018, 9(1), 1782–1792 CrossRef CAS PubMed.
  29. R. Wang, D. Zhang, W. Sun, Z. Han and C. Liu, A novel aluminum-doped carbon nanotubes sensor for carbon monoxide, J. Mol. Struct.: THEOCHEM, 2007, 806(1–3), 93–97 CrossRef CAS.
  30. X. Jia, L. An and T. Chen, Adsorption of nitrogen oxides on Al-doped carbon nanotubes: the first principles study, Adsorption, 2020, 26(4), 587–595 CrossRef CAS.
  31. W. Wang, Y. Zhang, C. Shen and Y. Chai, Adsorption of CO molecules on doped graphene: a first-principles study, AIP Adv., 2016, 6(2), 025317 CrossRef.
  32. R. Gholizadeh and Y.-X. Yu, N2O + CO reaction over Si-and Se-doped graphenes: an ab initio DFT study, Appl. Surf. Sci., 2015, 357, 1187–1195 CrossRef CAS.
  33. S. Kumar, M. Malhotra and H. Sharma, Adsorption of gas molecules on ultra-thin pristine and doped graphene nanoribbons, Mater. Res. Express, 2018, 5(10), 105007 CrossRef.
  34. T. C. Nguyen and M. T. Nguyen, First-principles Studies of CO2 and NH3 Gas Molecules Adsorbed on Graphene Nanoribbons, VNU Journal of Science: Mathematics-Physics, 2016, 32(2), 15–21 Search PubMed.
  35. E. Salih and A. I. Ayesh, Enhancing the sensing performance of zigzag graphene nanoribbon to detect NO, NO2, and NH3 gases, Sensors, 2020, 20(14), 3932 CrossRef CAS PubMed.
  36. S. F. Rastegar, A. A. Peyghan, H. R. Ghenaatian and N. L. Hadipour, NO2 detection by nanosized AlN sheet in the presence of NH3: DFT studies, Appl. Surf. Sci., 2013, 274, 217–220 CrossRef CAS.
  37. J. Beheshtian, Z. Bagheri, M. Kamfiroozi and A. Ahmadi, A theoretical study of CO adsorption on aluminum nitride nanotubes, Struct. Chem., 2012, 23, 653–657 CrossRef CAS.
  38. A. Ahmadi, J. Beheshtian and N. L. Hadipour, Interaction of NH3 with aluminum nitride nanotube: electrostatic vs. covalent, Phys. E, 2011, 43(9), 1717–1719 CrossRef CAS.
  39. Y. Li, Z. Zhou and J. Zhao, Transformation from chemisorption to physisorption with tube diameter and gas concentration: computational studies on NH3 adsorption in BN nanotubes, J. Chem. Phys., 2007, 127(18), 184705 CrossRef PubMed.
  40. H. Y. Ammar, H. M. Badran and K. M. Eid, 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 CrossRef CAS.
  41. M. Y. Mehboob, F. Hussain, R. Hussain, S. Ali, Z. Irshad and M. Adnan, et al., Designing of Inorganic Al12N12 Nanocluster with Fe, Co, Ni, Cu and Zn Metals for Efficient Hydrogen Storage Materials, J. Comput. Biophys. Chem., 2021, 20(04), 359–375 CrossRef CAS.
  42. J. Beheshtian, Z. Bagheri, M. Kamfiroozi and A. Ahmadi, Toxic CO detection by B12N12 nanocluster, Microelectron. J., 2011, 42(12), 1400–1403 CrossRef CAS.
  43. R. Kartika, F. H. Alsultany, A. Turki Jalil, M. Z. Mahmoud, M. N. Fenjan and H. Rajabzadeh, Ca12O12 nanocluster as highly sensitive material for the detection of hazardous mustard gas: density-functional theory, Inorg. Chem. Commun., 2022, 137, 109174 CrossRef CAS.
  44. H. Louis, I. O. Amodu, T. O. Unimuke, T. E. Gber, B. B. Isang and A. S. Adeyinka, Modeling of Ca12O12, Mg12O12, and Al12N12 nanostructured materials as sensors for phosgene (Cl2CO), Mater. Today Commun., 2022, 32, 103946 CrossRef CAS.
  45. L.-K. Li, Y.-Q. Ma, K.-N. Li, W.-L. Xie and B. Huang, Structural and electronic properties of H2, CO, CH4, NO, and NH3 adsorbed onto Al12Si12 nanocages using density functional theory, Front. Chem., 2023, 11, 1143951 CrossRef CAS PubMed.
  46. S. Hussain, S. A. S. Chatha, A. I. Hussain, R. Hussain, M. Yasir Mehboob and A. Mansha, et al., A Theoretical Framework of Zinc-Decorated Inorganic Mg12O12 Nanoclusters for Efficient COCl2 Adsorption: A Step Forward toward the Development of COCl2 Sensing Materials, ACS Omega, 2021, 6(30), 19435–19444 CrossRef CAS PubMed.
  47. S. Hussain, S. A. S. Chatha, A. I. Hussain, R. Hussain, M. Yasir Mehboob and A. Mansha, et al., In Silico Designing of Mg12O12 Nanoclusters with a Late Transition Metal for NO2 Adsorption: An Efficient Approach toward the Development of NO2 Sensing Materials, ACS Omega, 2021, 6(22), 14191–14199 CrossRef CAS PubMed.
  48. S. Patel, P. Patel, D. Chodvadiya, N. N. Som and P. K. Jha, 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 CrossRef CAS.
  49. R. Sainda, D. Chodvadiya and P. K. Jha, A DFT study of Mg9O9 nanoring for gas sensing and removal applications, J. Mol. Liq., 2024, 397, 124121 CrossRef CAS.
  50. R. Rahimi and M. Solimannejad, A study combining DFT and molecular dynamics simulations into the performance of B6N6 nanosheets for CO2 capture and separation, Appl. Phys. A: Mater. Sci. Process., 2023, 130(1), 21 CrossRef.
  51. J. Panchal, A. Gauswami, D. Chodvadiya, H. Jadeja and P. K. Jha, Adsorption performance of CO, NO and NH3 hazardous gas molecules over B9N9 and Al9N9 nanorings: Acumen from density functional theory, Mater. Chem. Phys., 2024, 311, 128565 CrossRef CAS.
  52. J. S. Al-Otaibi, Y. S. Mary, Y. S. Mary, J. N. C. Mishma and A. Manikandan, Investigation of the adsorption and sensor properties of resorcinol (RSL) on X9N9 (X = Al, B, Ga, In) nanorings by DFT and QTAIM analysis with solvent effects, J. Mol. Liq., 2024, 398, 124285 CrossRef CAS.
  53. J. S. Al-Otaibi, Y. S. Mary, U. Jethawa, B. Chakraborty and M. C. Gamberini, Examining the adsorption and sensing characteristics of cytosine (CTE) on Y9N9 (Y = Al, B, Ga) nanorings using solvent effects, DFT, AIM and SERS analyses, Spectrochim. Acta, Part A, 2025, 337, 126148 CrossRef CAS PubMed.
  54. M. J. Frisch, Gaussian09, 2009, https://www.gaussian.com/ Search PubMed.
  55. R. Dennington, T. Keith and J. Millam, GaussView, Version 5, 2009 Search PubMed.
  56. S. Grimme, S. Ehrlich and L. Goerigk, Effect of the damping function in dispersion corrected density functional theory, J. Comput. Chem., 2011, 32(7), 1456–1465 CrossRef CAS PubMed.
  57. S. Grimme and M. Steinmetz, Effects of London dispersion correction in density functional theory on the structures of organic molecules in the gas phase, Phys. Chem. Chem. Phys., 2013, 15(38), 16031–16042 RSC.
  58. H. Liu, F. Wang, K. Hu, T. Li and Y. Yan, Pd4 cluster decorated SnO2 nanowire for detecting characteristic gases in oil-immersed transformers: a theoretical and experimental study, Appl. Surf. Sci., 2022, 590, 153122 CrossRef CAS.
  59. S. F. Boys and F. Bernardi, The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors, Mol. Phys., 1970, 19(4), 553–566 CrossRef CAS.
  60. J.-i. Aihara, Reduced HOMO–LUMO gap as an index of kinetic stability for polycyclic aromatic hydrocarbons, J. Phys. Chem. A, 1999, 103(37), 7487–7495 CrossRef CAS.
  61. Z. Chen, X. Zhang, H. Xiong, D. Chen, H. Cheng and J. Tang, et al., Dissolved gas analysis in transformer oil using Pt-doped WSe2 monolayer based on first principles method, IEEE Access, 2019, 7, 72012–72019 Search PubMed.
  62. T. Lu and F. Chen, Multiwfn: a multifunctional wavefunction analyzer, J. Comput. Chem., 2012, 33(5), 580–592 CrossRef CAS PubMed.
  63. Origin V, OriginLab Corporation, Northampton, MA, USA, 2018 Search PubMed.
  64. G. Korotcenkov and G. Korotcenkov, Sensing layers in work-function-type gas sensors, in Handbook of Gas Sensor Materials: Properties, Advantages and Shortcomings for Applications Volume 1: Conventional Approaches, 2013, pp. 377–388 Search PubMed.
  65. G. Halek, I. D. Baikie, H. Teterycz, P. Halek, P. Suchorska-Woźniak and K. Wiśniewski, Work function analysis of gas sensitive WO3 layers with Pt doping, Sens. Actuators, B, 2013, 187, 379–385 CrossRef CAS.
  66. P. Davydovskaya, A. Tawil and R. Pohle, Ethanol sensing with Cu-BTC Metal Organic Framework: mass sensitive, work function based and IR investigations, Key Eng. Mater., 2014, 605, 87–90 Search PubMed.
  67. F. Zhuang, G. Hao, R. Zhou, W. Lei, Y. Deng and X. Ye, et al., Electric fields and strains effect on the electronic and optical properties of Zr2CO2/MoSSe van der Waals heterostructure, Mater. Today Commun., 2022, 33, 104295 CrossRef CAS.
  68. H. A. Rizwan, M. U. Khan, A. Anwar, M. Idrees and N. A. Siddiqui, A molecular modeling study of pristine and Li-doped B16N16 nanocages for sensing G-series nerve agents using DFT-D3, J. Mol. Graphics Modell., 2025, 139, 109069 CrossRef CAS PubMed.
  69. C. J. Cramer, Essentials of Computational Chemistry: Theories and Models, John Wiley & Sons, 2013 Search PubMed.
  70. F. Jensen, Introduction to Computational Chemistry, John Wiley & Sons, 2017 Search PubMed.
  71. S. Vadalkar, D. Chodvadiya, N. N. Som, K. N. Vyas, P. K. Jha and B. Chakraborty, An ab-initio study of the C18 nanocluster for hazardous gas sensor application, ChemistrySelect, 2022, 7(3), e202103874 CrossRef CAS.
  72. T. O. Unimuke, H. Louis, E. A. Eno, E. C. Agwamba and A. S. Adeyinka, Meta-hybrid density functional theory prediction of the reactivity, stability, and IGM of azepane, oxepane, thiepane, and halogenated cycloheptane, ACS Omega, 2022, 7(16), 13704–13720 CrossRef CAS PubMed.
  73. A. Mardiana-Idayu and S. B. Riffat, An experimental study on the performance of enthalpy recovery system for building applications, Energy and Buildings, 2011, 43(9), 2533–2538 CrossRef.
  74. V. Nagarajan, R. Bhuvaneswari and R. Chandiramouli, Interaction studies of propylene and butadiene on tricycle graphane nanosheet–a DFT outlook, J. Mol. Graphics Modell., 2023, 121, 108449 CrossRef CAS PubMed.
  75. M. S. Jyothi, V. Nagarajan and R. Chandiramouli, Chemisorption of atrazine and diuron molecules on γ-arsenene nanosheet-a first-principles study, Chem. Phys. Lett., 2022, 794, 139484 CrossRef CAS.
  76. M. A. Hossain, M. R. Hossain, M. K. Hossain, J. I. Khandaker, F. Ahmed and T. Ferdous, et al., An ab initio study of the B35 boron nanocluster for application as atmospheric gas (NO, NO2, N2O, NH3) sensor, Chem. Phys. Lett., 2020, 754, 137701 CrossRef CAS.
  77. M. Arjmandi, A. Arjmandi, M. Peyravi and A. K. Pirzaman, First-principles study of adsorption of XCN (X = F, Cl, and Br) on surfaces of polyaniline, Russ. J. Phys. Chem. A, 2020, 94, 2148–2154 CrossRef.
  78. J. A. Agwupuye, H. Louis, T. O. Unimuke, P. David, E. I. Ubana and Y. L. Moshood, Electronic structure investigation of the stability, reactivity, NBO analysis, thermodynamics, and the nature of the interactions in methyl-substituted imidazolium-based ionic liquids, J. Mol. Liq., 2021, 337, 116458 CrossRef CAS.
  79. R. Gangadharan and S. Sampath Krishnan, Natural bond orbital (NBO) population analysis of 1-azanapthalene-8-ol, Acta Phys. Pol. A, 2014, 125(1), 18–22 CrossRef.
  80. H. A. Rizwan, M. U. Khan, A. Hamid, J. Yaqoob, R. Hussain and S. Ahmed, et al., Molecular modelling of Al24N24 nanocage for the chemical sensing of phosgene and mustard chemical warfare agents: first theoretical framework, Comput. Theor. Chem., 2023, 1230, 114349 CrossRef.
  81. E. R. Johnson, S. Keinan, P. Mori-Sánchez, J. Contreras-García, A. J. Cohen and W. Yang, Revealing Noncovalent Interactions, J. Am. Chem. Soc., 2010, 132(18), 6498–6506 CrossRef CAS PubMed.
  82. G. Saleh, C. Gatti and L. Lo Presti, Non-covalent interaction via the reduced density gradient: independent atom model vs. experimental multipolar electron densities, Comput. Theor. Chem., 2012, 998, 148–163 CrossRef CAS.
  83. M. M. Edim, O. C. Enudi, B. B. Asuquo, H. Louis, E. A. Bisong and J. A. Agwupuye, et al., Aromaticity indices, electronic structural properties, and fuzzy atomic space investigations of naphthalene and its aza-derivatives, Heliyon, 2021, 7(2), e06138 CrossRef CAS PubMed.
  84. V. Nagarajan and R. Chandiramouli, Sensing response of novel ε-antimonene nanosheet towards ethyl acetate and isopropyl acetate–a DFT insight, Chem. Phys., 2022, 562, 111632 CrossRef CAS.
  85. N. F. Mott, Conduction in non-crystalline materials: III. Localized states in a pseudogap and near extremities of conduction and valence bands, Philos. Mag., 1969, 19(160), 835–852 CrossRef CAS.
  86. J. Chen, G. Zhu, F. Wang, Y. Xu, C. Wang and Y. Zhu, et al., Design of flexible strain sensor with both ultralow detection limit and wide sensing range via the multiple sensing mechanisms, Compos. Sci. Technol., 2021, 213, 108932 CrossRef.
  87. V. Nagarajan, R. Ramesh and R. Chandiramouli, Chemical sensing properties of σ-PXene sheets towards dioxin vapours–a first-principles insight, Comput. Theor. Chem., 2023, 1228, 114300 CrossRef CAS.
  88. S. Saravanan, V. Nagarajan, A. Srivastava and R. Chandiramouli, Blue phosphorene nanoribbon for detection of chloroform vapours–a first-principles study, Int. J. Environ. Anal. Chem., 2021, 101(12), 1697–1709 CrossRef CAS.
  89. H. A. Rizwan, M. U. Khan, A. Anwar, M. U. Khan, A. Sohail and S. Ahmed, et al., Deciphering the adsorption and sensing performance of Al24N24 and B24N24 nanoclusters as a drug delivery system for nitrosourea anticancer drug: a DFT insight, Surf. Interfaces, 2024, 51, 104779 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra03312h

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.