DOI: 
10.1039/D5RA04777C
(Paper)
RSC Adv., 2025, 
15, 39604-39615
Surface tailoring of porphyrin via phosphorus-doping and chromium encapsulation towards the detection of CO2, SO2, and NO2 gas pollutants: a computational study
Received 
4th July 2025
, Accepted 13th October 2025
First published on 20th October 2025
Abstract
As burning of fossil fuels is a major contributor to air pollution, as a result, both acute and chronic releases of these toxic chemical gases into the air can cause significant damage to the cardiovascular and pulmonary systems, potentially leading to death. This therefore calls for environmental remediation, through the detection of these gases, since they cannot be completely eradicated from the atmosphere. Herein, this research presents a newly tailored phosphorus-doped chromium encapsulation porphyrin (P–Cr@PPR) for the detection of CO2, SO2, and NO2 gas pollutants. The potential of this material as a suitable detector for these gases was studied through various computational analyses, carried out using the DFT/HSEH1PBE/LANL2DZ level of theory. Due to interaction, slight changes in the surface morphology were observed, showing the effect of detection on the surface. All systems showcased high perturbation energy, with the greatest perturbation energies of 909.58 and 481.60 kcal mol−1 observed for the surface upon the adsorption of SO2 gas, showing that the Cr–P@PPR–SO2 complex will be easily stabilized as compared to its counterparts studied. Majorly, slight changes in the surface morphology show the effect of the detection of the surface. Among the systems, non-covalent and partial covalent forms of interactions were observed, with a positive value of ellipticity (ε) showing some degree of ionic character with greater ionic contribution. Lastly, detecting strength follows an increasing order of: NO2–Cr–P@PPR < SO2–Cr–P@PPR < CO2–Cr–P@PPR, with the detection energies recorded as 10.471, 6.503, and 4.581 eV, respectively. Based on the characteristic properties exhibited by this material, it can be a potential candidate to be considered when selecting a material to engineer a detection device for these gas pollutants.
1 Introduction
Human activities significantly contribute to air pollution through various sources like burning, smoking, and industrial processes. This pollution is caused by harmful gases like CO2, SO2, and NO2, amongst others, posing a significant threat to human health and the environment.1–3 These gases originate from various sources, including domestic appliances, power plants, and burning fossil fuels.4,5 They can cause respiratory problems, acid rain, and contribute to climate change.6 Nitrogen dioxide (NO2), sulfur dioxide (SO2), and carbon monoxide (CO) are some of the important gases that serve as ambient air pollutants. Exposure to these gases, particularly NO2, can cause catastrophic injuries and even death in humans. It can also increase the risk of respiratory tract infections by weakening the immune system.7,8 Sulfur dioxide (SO2) can cause respiratory symptoms in patients, especially those with underlying pulmonary disease. Research shows that experimental exposure to SO2 can lead to changes in airway physiology, including increased airway resistance.9,10 Both acute and chronic releases of these toxic chemical gases into the air can cause significant damage to the cardiovascular and pulmonary systems, potentially leading to death. In addition, high concentrations of NO2 in confined spaces can be extremely harmful, even fatal to humans.11–13 Human activities, such as burning fossil fuels (coal, oil, and natural gas), are major contributors to air pollution, including CO2 and NO2 emissions.14 Also, sulfur dioxide (SO2) is known as a key component of acid rain. When sulfuric acid falls as rain, it acidifies both aquatic and terrestrial ecosystems, harming plants and animals.15,16 Due to their harmful effects, the need for environmental remediation brings about the detection of these gases.
Nanomaterials (where some call it a tiny powerhouse) are a fascinating class of materials with at least one dimension less than 100 nanometers.17 From this, they are significantly smaller than microscale materials, which typically range from 1 to 100 micrometers. In simpler terms, nanomaterials are about one billionth of a meter in size.18,19 These materials come in different shapes, namely: nanoparticles, nanorods, and nanosheets, amongst others.20 Their dimensionality is often used in determining their classification, wherein zero-dimensional nanomaterials are called nanoparticles. In furtherance, these materials can be categorized distinctively based on their unique properties and structures: metal nanomaterials use divalent or trivalent metal ions as starting materials and can be prepared through various methods like chemical or photochemical techniques; semiconductor nanomaterials bridges the gap between metals and non-metals, these materials have wide band gaps that influence their properties when modified; polymeric nanomaterials encompasses nanocomposites, polymer nanoparticles, block copolymers, and dendrimers; and lastly carbon-based nanomaterials as the name suggests, are carbon-based and include fullerenes and carbon nanotubes.21–23 Porphyrin, a molecule with just two nitrogen atoms linked by a carbon chain, is generating excitement as a powerful gas sensor. It is important to note that porphyrin can not only capture gases but also selectively choose its targets. Scientists are customizing its structure using different functional groups, like adding hydroxyl groups to attract SO2 or alkyl chains for CO2. This targeted approach is even being used to tackle a real-world problem: capturing harmful SO2 emissions from industrial sources. Research by Gajjar and Roy (2024), studied gas sensing through transport characteristics and analyzed NH3, SO2, NO2, and H2O for adsorption of energy, charge density, and recovery time. Their findings indicate that the porphyrin's electronic transport changes significantly with gas adsorption. This suggests the potential for low-power, ultrasensitive, recyclable gas sensors using NEGF formalism and DFT.24
Various research has been conducted on the adsorption of CO2, SO2, and NO2 gases on different nano surfaces. However, a few works have been done using the porphyrin surface for the adsorption of these gases. In an adsorption of carbon dioxide, ethane, and methane on porphyrin-based nanoporous organic polymers carried out by Jun Yan and team, a beneficial result towards designing and constructing better surfaces derived from tetrahedral-structured building blocks for simple molecular gas adsorption in the future.25 In another study conducted by Yeongran et al. on gas capture and precious metal adsorption using alkyl-linked porphyrin porous polymers, inexpensive and scalable porous porphyrin polymers were found to offer great potential in capture, separation, and precious metal recovery.26 Haang and Khang investigated the adsorption properties of the Sc2CF2 monolayer towards the adsorption of various gases, including CO2, NO2, and SO2. It was found that the adsorption of these gases does not change the conductive behavior of the monolayer. Also, the adsorption energy indicates that the monolayer exhibits low adsorption selectivity. This is due to the molecule's physiosorbed on the Sc2CF2 monolayer.27 Furthermore, Saini and Pandey (2025) studied 5,10,15,20-tetrakis-(4-hydroxyphenyl)porphyrin (TPPH) and its nickel (NiTPPH) and zinc (ZnTPPH) forms as gas sensors for CO2, COCl2, CO, and HCN using density functional theory (DFT). The results show NiTPPH and ZnTPPH have strong affinity for CO2 and COCl2, with NiTPPH-COCl2 having the most stable interaction (adsorption energy ∼ 0.326 eV). Thus, both porphyrins are minimally sensitive to CO and HCN, confirming selectivity. Overall, NiTPPH and ZnTPPH are established as promising candidates for selective CO2 and COCl2 sensing. This provides a theoretical basis for developing advanced porphyrin-based materials for environmental monitoring.28
Furthermore, recent Density Functional Theory (DFT) investigations have explored the absorption of these industrial gases (CO2, SO2, NO2) on carbon-based materials, revealing that surface modification influences adsorption strength and electronic sensitivity. Akhmetsadyk et al. (2023), examined SO2 adsorption on pristine and N, GA, and Ga–N co-doped graphene using the GGA-PBE functional with a DNP basis set in DMol3, and reported weak physisorption on pristine graphene (Eads = −0.32 eV) but much stronger chemisorption after Ga-doping, accompanied by significant electronic structure changes.29 Similarly, Lin et al. (2023) investigated NO2, SO2, and NO gases adsorption on Janus MoSeTe monolayers decorated with Fe, Co, and Ni transition metals.30 The findings revealed that TM-decoration converts weak physisorption on pristine MoSeTe into strong, mainly chemisorptive binding for NO2/SO2 (approx −2.4 eV for TM-decorated systems and 1.7–1.9 eV for pristine MOSeTe). Faghihnasiri & Branicio (2025) used DFT to assess CO2 vs. N2 on graphene; the authors report weak physisorption with CO2 adsorption energy of −0.168 eV (N2 = −0.120 eV), i.e., small intrinsic selectivity on pristine graphene.31 However, these studies are limited either to a single gas species or to non-metal doped carbons, offering little insight into how transition-metal centers influence the simultaneous adsorption of different industrial gases.
This current research utilizes density functional theory (DFT) to investigate how phosphorus (P) doping and chromium (Cr) encapsulation enhance the adsorption and sensing performance of a single-layer porphyrin (PPR) material for CO2, SO2, and NO2 gases. This work aims to investigate the performance of phosphorus-doped chromium-encapsulated porphyrin (Cr–P@PPR) material towards the adsorption of the labelled gases. The modification of porphyrin will be valuable in synergistic modulation of the electronic structure and behavior that is achieved through a combined effect of phosphorus doing chromium encapsulation, which has not been explored in previous metal-doped porphyrin-based gas sensors. This will result in the first reported Chromium-phosphorous engineered porphyrin nanostructure, enabling enhanced sensitivity and selectivity for CO2, SO2, and NO2 compared to previously reported metal-doped porphyrins and other carbon materials. Cr metal creates a 3D environment and steric confinement, leading to specific site for orbital overlap and selective trapping of the gases. The effect of surface modification by phosphorus-doping and chromium-encapsulation on the surface was determined through various computational analyses, such as geometry optimization, electronic properties (FMO and NBO analyses), visual study (QTAIM and NCI analyses), adsorption energy computation, and, lastly, the use of the selected sensor properties. This exploration opens doors for developing PPR-based gas detectors applicable to monitor operating conditions in CO2, SO2, and NO2 gases.
2 Computational details
2.1 DFT approach
All software programs computationally utilized for the design and calculations of phosphorus-doped chromium-encapsulated porphyrin (Cr–P@PPR) are as follows: Gaussview 6.0.16,32 Gaussian 16 for optimization,33 Multiwfn software 3.7 for visualizing the reduced density gradient (RDG) and derivation of the QTAIM descriptor,34 visual molecular dynamics (VMD) software for plotting the RDG,35 and ChemCraft software was employed in the visualization of HOMO–LUMO electronic isosurfaces of the systems.36
2.1.1 Engineering of phosphorus-doped chromium-encapsulated porphyrin (Cr–P@PPR). The studied Cr–P@PPR was modeled by firstly doping the porphyrin surface with the phosphorus atom and then encapsulating it with chromium, forming phosphorus-doped chromium-encapsulated porphyrin (Cr–P@PPR), which was optimized for geometry equilibration.
2.1.2 Geometry optimization of the surface and adsorbate. Single-point energy (SPE) calculations and geometry optimizations have been performed within the framework of density functional theory (DFT), utilizing the Gaussian 16 code and GaussView 6.0.16 software. The calculations were carried out using the GENECP/LANL2DZ/Def2svp basis set and incorporating the HSEH1PBE functional due to its proven accuracy in predicting the electronic and adsorption properties of a system.37 The HSEH1PBE hybrid functional was employed because its screened Hartree–Fock exchange improves the description of localized d-orbitals in transition-metal systems, such as Cr-based adsorbents, where conventional GGA functionals often underestimate electronic interactions. Prior studies have shown that HSE-type functionals yield reliable adsorption energies and electronic properties for gas-surface systems.38 Grimme's D3 dispersion correction effect was included to ensure accurate treatment of long-range van der Waals interactions critical for gas-surface interactions. This consideration was consistently applied across all computations and analyses presented in this current study. For the basis sets, LANL2DZ with effective core potentials was applied to Cr to account for relativistic effects efficiently. Def2-SVP was used for lighter atoms (C, H, O, N, S) to achieve balanced accuracy in geometry optimization and adsorption studies. The GENECP combination ensures accurate treatment of the transition metal while maintaining computational efficiency.39 The calculations herein included the quantum descriptors, natural bond orbital (NBO), the topology analysis for inter and intramolecular interactions (QTAIM and NCI), and adsorption studies, which were integral aspects of this study. 
2.1.3 Adsorption strategy of the gases (CO2, SO2, and NO2) on phosphorus-doped chromium-encapsulated porphyrin. The optimization of the surface was carried out before modification, then after modification, it was optimized again to attain a stable geometry structure. The gases are positioned randomly at the center of the surface for adsorption. To ascertain the best configuration, equilibration of the surfaces close to the gases was carried out during optimization.
3 Results and discussions
3.1 Structural geometry analysis
The studied Cr–P@PPR surface and CO2–Cr–P@PPR, NO2–Cr–P@PPR, and SO2–Cr–P@PPR complexes were optimized using the DFT/HSEH1PBE/LANL2DZ level of theory. Optimization plays a very important role in the computation of adsorption energy between the investigated surface and the gases (CO2, NO2, and SO2). This present study involves doping the porphyrin surface with chromium and phosphorus atoms to form phosphorus-doped chromium-encapsulated porphyrin (Cr–P@PPR). Upon modification, the newly tailored surfaces were positioned to ensure the best adsorption configuration towards the adsorption of CO2, NO2, and SO2 gas molecules. Upon adsorption, the resulting complexes are CO2–Cr–P@PPR, NO2–Cr–P@PPR, and SO2–Cr–P@PPR corresponding to the adsorption of CO2, NO2, and SO2. Fig. 1 presents the optimized structures of the gases, adsorbent, and complexes formed.
|  | 
|  | Fig. 1  Optimized structures of CO2, SO2, and NO2 adsorption on the Cr–P@PPR surface, showing the configurations before and after adsorption for each gas molecule. |  | 
The gases shown in the figure showcase geometries that agree with previous computational research. Herein, the bonds C–O, S–O, and N–O with the bond lengths of 1.157, 1.145, and 1.193 Å are computationally intact, as they correspond to previous theoretical work.40,41 The pictorial representation of the gases helps illustrate the increase in the bond length after adsorption. Table 1 shows the summary of the computed bond lengths during the pre- and post-adsorption. This increment in bond lengths, as observed, is nearly insignificant when compared with those obtained before adsorption. For instance, the complex CO2–Cr–P@PPR has a bond length of 2.077 Å (Cr39–N19) before adsorption, and then that of 2.081 Å upon adsorption. Upon adsorption, the bonds within the surface have been observed to slightly increase and decrease, wherein the SO2–Cr–P@PPR complex showcased a decrease. In furtherance, another decrement was observed in the NO2–Cr–P@PPR complex. Lastly, the CO2–Cr–P@PPR complex shows an increment and a decrement at different instances in its bond lengths. This result indicates that adsorption of the labelled gases causes slight changes in the surface morphology, thereby elucidating the presence of the mentioned gases on the adsorbent.
Table 1 Cr–N bond lengths (Å) in the Cr–P@PPR complex before and after adsorption of CO2, SO2, and NO2 gas molecules, showing structural changes upon gas interaction
		
| Complex | Bond label | Pre-adsorption | Post-adsorption | 
| CO2–Cr–P@PPR | Cr39–N5 | 2.090 | 2.081 | 
| Cr39–N19 | 2.077 | 2.081 | 
| Cr–N12 | 2.054 | 2.042 | 
| Cr–N26 | 2.041 | 2.042 | 
| SO2–Cr–P@PPR | Cr–N5 | 2.090 | 2.068 | 
| Cr–N19 | 2.077 | 2.075 | 
| Cr–N12 | 2.054 | 2.033 | 
| Cr–N26 | 2.041 | 2.025 | 
| NO2–Cr–P@PPR | Cr–N5 | 2.090 | 2.062 | 
| Cr–N19 | 2.077 | 2.099 | 
| Cr–N12 | 2.054 | 2.019 | 
| Cr–N26 | 2.041 | 1.996 | 
3.2 Electronic properties
3.2.1 FMO analysis: energy gap fluctuation and the chemical quantum descriptors. Quantum descriptors are very important parameters that enable one to understand the reactivity, stability, and conductivity of a complex. It therefore becomes a crucial approach used in computational chemistry to understand the electronic and optical properties, taking cognizance of the energies of the highest occupied orbital (EHOMO) and that of the lowest unoccupied orbital (ELUMO), where they are fundamentally applied in obtaining the energy gap.42,43 The energy gap is the difference between the HOMO and the LUMO, and the results are summarized in Table 2.
Table 2 Quantum chemical descriptors of Cr–P@PPR and its gas-adsorbed complexes (CO2, SO2, and NO2), with all units in electron Volt (eV), except for chemical softness (S) with a unit of eV−1
		
| Complexes | EHOMO eV | ELUMO eV | IP eV | EA eV | EGAP eV | μ (eV) | χ (eV) | η (eV) | S (eV−1) | 
| Cr–P@PPR | −4.352 | −3.054 | 4.352 | 3.054 | 1.297 | −3.703 | 3.703 | 0.648 | 0.770 | 
| CO2–Cr@PPR | −5.269 | −3.421 | 5.269 | 3.421 | 1.848 | −4.343 | 4.343 | 0.924 | 0.540 | 
| SO2–Cr–P@PPR | −5.668 | −3.845 | 3.668 | 3.845 | 1.823 | −4.757 | 4.757 | 0.911 | 0.548 | 
| NO2–Cr–P@PPR | −4.749 | −3.677 | 4.749 | 3.677 | 1.071 | −4.213 | 4.213 | 0.535 | 0.933 | 
The highest energy gap was observed in the complex CO2–Cr@PPR with a value of 1.848 eV. Although the pure surface has a small energy gap of 1.297 eV, it increases upon adsorption in CO2–Cr@PPR and SO2–Cr@PPR. A study by Saini and Pandey (2025), that the adsorption of gases, including CO2, resulted in a significant increase in the energy gap.44 A decrease in the energy gap was obtained in NO2–Cr@PPR, where it decreased to 1.071 eV. The effect of CO2 adsorption differs from NO2 and SO2. This is explained by the chemical activities of the P-block elements (non-metals). The results showed higher reactivity of the modified porphyrin surface in adsorbing gases compared to those obtained from Ammar and Badran's work (2019), whose transition metal-doped porphyrin surfaces had an energy gap within the range of 2.827–3.124 eV.45 This implies a significant influence of the dopant on the change in the electronic state of the material. Further investigation of the stability and reactivity of the complexes and surface can be scrutinized with the help of the chemical descriptors. Improvement in the chemical reactivity and kinetic stability of the studied complexes has been achieved through the reactivity descriptors.46 The subject encompassing the frontier molecular orbital envelops crucial information about the studied complexes and surface. These descriptors include the ionization potential (IP), electron affinity (EA), chemical softness (S), chemical potential (μ), electronegativity (X), and global hardness (η), which have been theoretically deployed to evaluate the reactivity properties of the studied complexes. The Koopmann hypothesis47 has been applied to calculate these descriptors as shown in the following eqn (1)–(6).
|  | |  | (3) | 
|  | |  | (4) | 
|  | |  | (5) | 
|  | |  | (6) | 
Referring to the Koopmann approximation, the ionization potential (IP) can be directly proportional to the energy of HOMO. The electron affinity (EA) has a relationship with LUMO energy. In conjunction, global hardness and the electrophilicity index account for greater stability and low reactivity of the complexes studied. The ionization potential is the energy needed for the transfer of electrons from HOMO to LUMO, and electron affinity is the energy released when molecules, atoms, etc. accept electron(s). Chemical hardness defines the tendency of a chemical species to resist electron charge transfer within its chemical environment.48 On the other hand, knowing that chemical hardness is the direct opposite of chemical softness it means the ability of a chemical species to attract electron charge transfer. These two parameters (chemical hardness and chemical softness) depend on how firmly the atoms are held together. From Table 2, a higher global hardness value of 0.924 eV was recorded for CO2–Cr@PPR. As chemical hardness displays small or mild resistance to electron transfer, this suggests relatively little resistance.49 Moreover, the least value of 0.535 eV was observed in complex NO2–Cr–P@PPR, suggesting a relatively greater resistance among its studied counterparts. Finally, the HOMO and LUMO plots showing the distribution of HOMO–LUMO for the surface before and after interaction are presented in Fig. 2.
|  | 
|  | Fig. 2  The visualization of the HOMO and LUMO plots for of Cr–P@PPR, CO2–Cr–P@PPR, NO2–Cr–P@PPR, and SO2–Cr–P@PPR. |  | 
 
3.2.2 NBO analysis: perturbation study. Natural bond orbital analysis plays a significant role in describing intermolecular and intramolecular interactions, arising from chemical bonds in molecules, making it an effective tool in elucidating hyperconjugation and electron density transfer from filled lone-pair electrons. According to Weinhold et al.,50 NBO analysis provides detailed information regarding the distribution of electrons with molecular bonds that exist between atomic species. The stabilization energy (second-order perturbation energy, E2) was theoretically calculated and expressed in eqn (7).|  | |  | (7) | 
where qi denotes the donor occupancy, F(i,j) is the off-diagonal NBO Fock Matrix elements, and Ei and Ej represent the diagonal elements. Reviews from literature show that the higher the second-order perturbation of energy corresponds with the stronger the interaction between the donor and acceptor orbital, with a greater level of conjugation.51 The summarized results of the NBO calculation for the studied surface and its complexes are presented in Table 3, illustrating charge delocalization from donor to acceptor orbitals, second-order perturbation energies, diagonal elements, and off-diagonal NBO Fock matrix elements. For the pristine Cr–P@PPR surface, electron transition occurs from donor to acceptor orbitals (π* → π*), with a perturbation energy of 126.42 kcal mol−1. Upon gas adsorption, the results reveal that SO2 exhibits the highest stabilization energies (909.58 and 481.60 kcal mol−1), indicating the strongest charge delocalization and interaction with the Cr–P@PPR surface. In comparison, CO2 shows moderate stabilization energies (320.06 and 160.04 kcal mol−1), while NO2 displays comparatively lower values (168.89 and 140.47 kcal mol−1). Specifically, for the Cr–P@PPR–NO2 complex, charge transitions occur from π*(2) C8–N12 → π*(2) C18–C34 and C22–N26 → π*(2) C11–C29, with corresponding energies of 168.89 and 140.47 kcal mol−1, respectively. Similarly, the CO2–Cr–P@PPR complex exhibits relatively higher stabilization energies (320.06 and 160.04 kcal mol−1), primarily due to charge delocalization involving lone pair → lone pair interactions.52 These findings suggest that the Cr–P@PPR surface possesses promising adsorption characteristics for gas capture, with SO2 forming the most stable complex, thereby demonstrating the strongest adsorption behavior compared to NO2 and CO2.
Table 3 Second-order perturbation stabilization energies (E2, kcal mol−1) from NBO analysis for Cr–P@PPR and its gas-adsorbed complexes (CO2, SO2, and NO2)
		
| System | Donor | Acceptor | E2 kcal mol−1 | E(i) − E(j) | F(ij) | 
| Cr–P@ PPR | π*(2)C1–C2 | π*(2)C25–C31 | 126.42 | 0.01 | 0.078 | 
| π*(2) C3–C4 | π*(2)C1–C2 | 120.53 | 0.01 | 0.073 | 
| CO2–Cr–P@PPR | LP* (1) C11 | LP (1) C29 | 320.06 | 0.03 | 0.138 | 
| LP*(5) Cr37 | LP*(6) Cr37 | 160.04 | 0.05 | 0.228 | 
| SO2–Cr–P@PPR | LP*(1) C31 | LP (1) C25 | 909.58 | 0.01 | 0.135 | 
| LP (1) C11 | LP*(1) C29 | 481.60 | 0.02 | 0.138 | 
| NO2–Cr–P@PPR | π*(2) C8–N12 | π (2) C18–C34 | 168.89 | 0.03 | 0.093 | 
| π*(2) C22– N26 | π*(2) C11–C29 | 140.47 | 0.04 | 0.091 | 
 
3.3 Visual study
3.3.1 The 3D-RDG maps analysis. The concept of non-covalent interaction (NCI) helps to explain in detail the structural properties as well as the forces taking place during the interaction. It enables us to effectively visualize the structures, interpret, and distinguish the kinds of interactions that occur. The NCI includes the use of electron density and its gradient (RDG), thus presented as a plot and the second eigenvalues of the Hessian matrix (λ2).53 The forces involved here are the hydrogen bonds, electrostatic forces, and van der Waals (repulsive forces), which are elucidated using the NCI study. During this study, the interaction between the surface Cr–P@PPR was investigated to determine the non-covalent interaction and how it affects the adsorption ability of the labelled gases. Using the computational software Visual Molecular Dynamics (VMD), the NCI was viewed, and the RDG scatter plots were obtained and incorporated for the evaluation of the readings obtained through this software application. From Fig. 3, it was observed that the complex CO2–Cr–P@PPR is characterized by a hydrogen bond, denoting a strong interaction. On the other hand, the complex SO2–Cr–P@PPR has a constituent property of a strong repulsive force caused by steric. This distinguishable characteristic of this complex is traceable to the features of its adsorbent (gas), although the complex NO2–Cr–P@PPR is more repulsive than attractive because of the higher steric effect.
|  | 
|  | Fig. 3  The 3D RDG maps for the studied complexes display the designated colour peaks, which explain the intermolecular interaction between the modified porphyrin surface and the adsorbed gases. |  | 
 
3.3.2 Quantum theory of atoms-in-molecules (AIMs) analysis. The theory of quantum molecules was developed to decrypt the intermolecular interaction that exists between molecules, which involves ionic bonding, van der Waals forces of attraction, and covalent bonding of structures.54 This analysis is often carried out to determine whether these molecules can interact between species. The phrase “quantum molecular theory” is best treated with various topological parameters to validate the authenticity of the nature of interaction arising because of adsorption. These parameters include: the density of all electrons ρ(r), the Lagrangian Kinetic Energy G(r), Laplacian of electron density ∇2ρ(r), potential energy density V(r), Energy density E(r) or H(r), Eigenvalues of Hessian (λ1, λ2, and λ3), Electron Localization Function (ELF), and Ellipticity of Electron Density ε.55 Ultimately, the basis for which these analyses were carried out was to ensure the strength of the interactions at different critical points (CPs) for each of the complexes and their possible significance as captured in sensor materials. Table 4 presents these topological parameters computed at bond critical points (BCPs). Literature review shows that the expression ρ(r) > 0.1is usually attributed to higher stability, which connotes the complexes studied at various CPs.56 The reverse implication of the latter is that, for every CP having ρ(r) < 0.1, there will be a likelihood that such a complex may not be stable.57 Furthermore, the density of electrons H(r) is responsible for the nature of interaction, which signifies that if this parameter has negative values, there is a likelihood of a covalent attraction of molecules. In furtherance, the Laplacian of Electron Density ∇2ρ(r) accounts elaborately on the nature of the interaction, wherein a Laplacian of Electron Density ∇2ρ(r) > 0 and Energy Density H(r) < 0, indicates non-covalent interaction.58 From Table 4, all complexes exhibit a non-covalent and partial covalent form of interaction. These properties are those found to align with sensor materials. Positive ellipticity of electron density ε > 0 suggests elongated electron density along the bond axis, thus signifying a polar covalent bond with some degree of ionic character.59 The larger the positive value, the greater the ionic contribution to the bond. This statement is significant to our case, as captured in Table 1, where the positive values show some degree of ionic character with greater ionic contribution.
Table 4 The summarized result of the topological parameters computed at bonds critical points (BCPs) and calculateda
		 
| System | Bonds | CPs | P(r) | ∇2 ρ(r) | G(r) | V(r) | K(r) | H(r) | ELF | LOL | λ1 | λ2 | λ3 | λ1/λ3 | ε | 
| The units of the parameters in Table 5 are as follows: P(r), e Ang−3; ∇2 ρ(r), e Ang−3; λ1, λ2, and λ3 have a unit of e Ang−5 respectively; G(r) and V(r) have units of eV; H(r) has a unit of eV Ang3 e−1; ellipticity of electron density (ε), electron localization function (ELF) and λ1/λ3 are dimensionless. The unit of G/|V| is Ang3. | 
| CO2–Cr_P@PPR | Cr37–O39 | 67 | 0.378 | 0.258 | 0.538 | −0.430 | −0.108 | 0.108 | 0.492 | 0.185 | −0.598 | −0.574 | 0.376 | −1.590 | 0.041 | 
| N26–C22 | 63 | 0.420 | 0.308 | 0.749 | −0.142 | 0.672 | −0.672 | 0.449 | 0.474 | 0.256 | −0.113 | −0.113 | −2.265 | 0.000 | 
| C4–N5 | 68 | 0.299 | −0.413 | 0.109 | −0.413 | 0.304 | −0.304 | 0.927 | 0.779 | −0.579 | 0.275 | −0.476 | 1.216 | 0.217 | 
| NO2–Cr_P@PPR | Cr37–N5 | 71 | 0.969 | 0.525 | 0.102 | −0.735 | −0.289 | 0.289 | 0.150 | 0.110 | 0.162 | 0.423 | −0.593 | −0.273 | −1.367 | 
| C15–P36 | 68 | 0.133 | 0.184 | 0.132 | −0.218 | 0.863 | −0.863 | 0.361 | 0.429 | −0.154 | 0.463 | −0.124 | 1.242 | 0.242 | 
| C18–N19 | 56 | 0.291 | −0.578 | 0.172 | −0.489 | 0.317 | −0.317 | 0.819 | 0.680 | −0.555 | 0.460 | −0.483 | 1.149 | 0.150 | 
| SO2–Cr_P@PPR | Cr37–S40 | 69 | 0.517 | 0.135 | 0.381 | −0.425 | 0.438 | −0.438 | 0.234 | 0.355 | −0.693 | −0.617 | 0.266 | 2.319 | 0.121 | 
| N26–C22 | 64 | 0.297 | −0.600 | 0.187 | −0.524 | 0.337 | −0.337 | 0.805 | 0.670 | 0.467 | −0.568 | −0.499 | 0.069 | 0.137 | 
| C4–N5 | 57 | 0.310 | −0.591 | 0.219 | −0.586 | 0.367 | −0.367 | 0.775 | 0.650 | 0.499 | −0.579 | −0.512 | −0.975 | 0.130 | 
 
3.4 Adsorption study
To gain elaborate insight into the subject of adsorption studies, the studied complexes were optimized before and after detection. On the cluster, we have the initial structural equilibration LANL2DZ as the employed basis set. Also, all computations were carried out using DFT/HSEH1PBE/LANL2DZ level of theory. The adsorption behaviors of the Cr–P@PPR surface in the detection of CO2, NO2, and SO2 have been effectively studied using the calculated adsorption energies of the studied complexes. Mathematically, the following eqn (8) was used to account for the adsorption of the various gases:|  | | Eads = EComplex − ESurface − EGas | (8) | 
Various theoretical literatures assert that greater adsorption of energy is beneficial in comparative studies across different complexes.60–62 That implies the greater the detection energy, the greater the likelihood that the adsorbent material best detects the adsorbate during the study. Although the detection energy in our case, as revealed in Table 5, shows physisorption with small values. According to their detecting abilities, the complexes follow the order: NO2–Cr–P@PPR < SO2–Cr–P@PPR < CO2–Cr–P@PPR, with the detection energies recorded as 10.471, 6.503, and 4.581, respectively. Although the adsorption energies for this study showed higher physisorption, the result is comparable to the findings presented in the research of Saini and Pandey (2025). Whose result showed approximately 0.326 eV energy for the adsorption of CO2 by metalated porphyrin material.28 Likewise, NO2 adsorption by porphyrin in the research of Shah et al. (2024) demonstrated chemisorption adsorption energy of −0.78 eV.63 BSSE correction was performed for the systems, and the adsorption energy was observed to range from 4.589 eV (CO2) to 10.481 eV (NO2), which implies that gas adsorption consistently occurs via weak electrostatic and van der Waals forces rather than chemical bonding. The inclusion of BSSE correction increased the adsorption energies by 0.008–0.111 eV, confirming that basis set overlap has only a minor numerical impact on the computed adsorption strengths. This minor correction validates the accuracy of the LANL2DZ/Def2svp basis sets and the capacity of the HSEH1PBE functional in describing non-covalent interactions in transition-metal systems. The physisorption of these gases depends on temperature and the surface. Physisorption of a gas increases with a decrease in temperature and surface area, which is due to the formation of van der Waals forces of attraction.64 Due to the physical nature of the adsorption of the three gases on the Cr–P@PPR surface, the investigated surface could be used for the detection of the labeled gases. This will lead to a quick recovery of the gases (reversible sensing), and reusability of the surface, as adsorption is weak. In conclusion, this surface may be suitable for detection rather than the removal of these gases. Thus, the weak yet stable physisorption highlights the potential of Cr–P@PPR for selective, reusable gas sensing applications.
Table 5 The adsorption energy of the studied complexes in Hartree (H) and electron Volt (eV) with energies >0 eV, depicting physisorption of the gases
		
| Complexes | EComplex | EAdsorbent | EAdsorbate | Eads Hartree | Eads(eV) | BSSE energy | Eads + BSSE (eV) | 
| NO2–Cr–P@PPR | −1580.741 | −1375.985 | −205.140 | 0.384 | 10.471 | 0.01004 | 10.481 | 
| SO2–Cr–P@PPR | −1924.126 | −1375.985 | −548.380 | 0.239 | 6.503 | 0.1105 | 6.6135 | 
| CO2–Cr–P@PPR | −1564.277 | −1375.985 | −188.460 | 0.168 | 4.581 | 0.0077 | 4.5887 | 
3.5 Sensor properties
Detection of gases on a newly engineered surface can be further analyzed through sensor properties such as conductivity, electrical back donation, and fraction of electron transfer, and charge transfer mechanism. These properties would provide further insights into the detecting or adsorbing behaviors of the new surface. This section will therefore focus on the applicability of a Cr–P@PPR adsorbent material towards the detection of some selected gases like CO2, SO2, and NO2 gas molecules.
3.5.1 Fraction of electron transfer (FET) and electrical back-donation. The usefulness of the global hardness in the realm of the adsorption mechanism cannot be overemphasized. In here, the electrical back donation has been computed using the values of the global hardness as calculated in the section on electronic properties. Generally, a good detector or sensor is one with an attribute of global hardness value greater than zero (η > 0) and a back donation less than zero (ΔEBack-donation < 0).65,66 The ΔEBack-donation and ΔN have been calculated using the eqn (9) and (10) respectively as follows:|  | |  | (9) | 
|  | |  | (10) | 
As observed, the electrical back donations are all negative (see Table 6), showing the characteristic of a promising adsorbent material. The calculated back donation presented in the table was observed to be in a small range of −0.231 to −0.134 eV, elucidating that the detection of these gases on the modified surface exhibits close and similar adsorbing properties across all complexes formed. FET values are within a close range of −0.0288 to 0.1386, with the least and greatest values attributed to SO2–Cr–P@PPR and NO2–Cr–P@PPR, respectively. Our result in this section agrees with those obtained in the previous section and can therefore suggest a good adsorption pattern.
Table 6 Summary of the sensor performance computed at DFT/B3LYP/6-31+G(d) computational methoda
		 
| Systems | Qt (e) | ΔN | ΔEBack-donation | ΔEg | % ΔEg | 
| The units of ΔEg and ΔEback-donation are in electron volt (eV). The Qt has a unit of electron (e), and finally, the % ΔEg has its unit in percentage. ΔN is dimensionless. | 
| CO2–Cr–P@PPR | 1.367 | 0.0883 | −0.231 | 0.425 | 42.483 | 
| SO2–Cr–P@PPR | 1.422 | −0.0288 | −0.134 | −0.174 | −17.425 | 
| NO2–Cr–P@PPR | 0.675 | 0.1386 | −0.228 | 0.406 | 40.555 | 
Notably, Wei et al. (2018), in their DFT studies of gas adsorption on doped MoS2 monolayers, specifically Ni-doped MoS2, reported ΔE values around −0.2 eV for SO2 adsorption, along with fractional charge transfers ΔQ in the range of approximately 0.10–0.15 e.67 Similarly, a study using phosphorene-based sensors by Cai et al. (2017) shows strong NO2 adsorption with significant charge transfer, approximately 0.10–0.20 e, and comparable adsorption energies.68 Thus, the results from our back-donation values (−0.231 to −0.134 eV) and FET values (approximately −0.028 to 0.14) fall well within the ranges reported for established 2D-material gas-sensor platforms, supporting the viability of Cr–P@PPR surfaces as competitive sensor candidates.
 
3.5.2 Charge transfer mechanism. The natural charge on the gases and the newly tailored material after adsorption were considered when computing the mechanism of charge transfer (Qt). This was calculated using eqn (11).69 The distribution of electron density makes it feasible to investigate the charge transfer mechanism of a complex.|  | | Qt = Qadsorption − Qisolated | (11) | 
From the equation, the charge transfer, charge transfers on the gases, and that on the surface are represented by Qt, Qisolated, and Qadsorption, respectively. Previous studies reported that a charge transfer from the gas to the surface is seen through the negative magnitude of charge transfer, whereas that from the surface to the gas is described by a positive charge transfer.70 Table 6 presents the summarized result of the charge transfer. For all three complexes, positive Qt values were obtained (see Table 6), showcasing electron transfer from the surface to the gases. As was observed, this is an indication of a good detecting material for CO2, SO2, and NO2 gas pollutants. Therefore, the newly engineered phosphorus-doped chromium-encapsulated porphyrin (Cr–P@PPR) possesses a good sensitivity and selectivity towards gas adsorption.
 
4 Conclusions
The detection of toxic gases for investigating the potential of a newly engineered phosphorus-doped chromium-encapsulated porphyrin (Cr–P@PPR) material was carried out using the DFT/B3LYP/LANL2DZ level of theory. Various theoretical/computational analyses were carried out to inquire into the inherent properties of these materials, and the following deductions have been drawn:
(I) Detection of the CO2, SO2, and NO2 gas molecules caused slight changes in the surface morphology, showing the effect of detection on the surface.
(II) Higher global hardness value of 0.924 eV in CO2–Cr@PPR suggests a relatively smaller resistance. The least value of 0.535 eV was observed in complex NO2–Cr–P@PPR, suggesting a relatively greater resistance among its studied counterparts.
(III) All systems showcased high perturbation energy. However, the greatest perturbation energies of 909.58 and 481.60 Kcal mol−1 observed for the surface upon the adsorption of SO2 gas show that the Cr–P@PPR–SO2 complex will be easily stabilized as compared to its studied counterparts.
(IV) AIM analysis showed that all complexes exhibit majorly non-covalent and partial covalent forms of interactions, which are found to align with sensor materials. Also, the positive values of ellipticity (ε) show some degree of ionic character with a greater ionic contribution.
(V) An increasing order of detecting strength follows: NO2–Cr–P@PPR < SO2–Cr–P@PPR < CO2–Cr–P@PPR, with the detection energies recorded as 10.471, 6.503, and 4.581 eV respectively.
Author contributions
Ededet A. Eno: conceptualization, project administration and supervision. Miracle N. Ogbogu: data curation and formal analysis. Opeyemi M. Oyebanji and Morenikeji A. Ajayi: manuscript writing, editing and interpretation of data. Wrote the paper. Stephen A. Adalikwu and Ernest G. Irilochuwe: validation and visualization. Jeremiah E. Ochepo: methodology and manuscript proofreading.
Conflicts of interest
We declare no conflict of any sort.
Data availability
All data are contained within the manuscript.
Acknowledgements
This work was not funded by any agency.
References
- I. B. Abaje, Y. Bello and S. A. Ahmad, A review of air quality and concentrations of air pollutants in Nigeria, J. Appl. Sci. Environ. Manag., 2020, 24(2), 373–379 CAS.
- J. Lewtas, Air pollution combustion emissions: characterization of causative agents and mechanisms associated with cancer, reproductive, and cardiovascular effects, Mutat. Res., Rev. Mutat. Res., 2007, 636(1–3), 95–133 CrossRef.
- D. K. Saini, S. K. Garg and M. Kumar, Major air pollutants and their effects on plant and human health: a review, Plant Arch., 2019, 19(2), 09725210 Search PubMed.
- R. Saidur, H. H. Masjuki, M. Y. Jamaluddin and S. Ahmed, Energy and associated greenhouse gas emissions from household appliances in Malaysia, Energy Policy, 2007, 35(3), 1648–1657 CrossRef.
- S. Ozgen, S. Cernuschi and S. Caserini, An overview of nitrogen oxides emissions from biomass combustion for domestic heat production, Renew. Sustain. Energy Rev., 2021, 135, 110–113 CrossRef.
- W. C. Clark, J. Jäger, J. Cavender-Bares and N. M. Dickson, Acid Rain, Ozone Depletion, and Climate Change: an Historical Overview,  2001 Search PubMed.
- H. W. T. Mapoma, C. Tenthani, M. Tsakama and I. B. M. Kosamu, Air quality assessment of carbon monoxide, nitrogen dioxide and sulfur dioxide levels in Blantyre, Malawi: a statistical approach to a stationary environmental monitoring station, Afr. J. Environ. Sci. Technol., 2014, 8(6), 330–343 CrossRef.
- S. Prasad, L. Zhao and J. Gomes, Methane and natural gas exposure limits, Epidemiology, 2011, 22(1), S251 CrossRef.
- E. Wigenstam, L. Elfsmark, A. Bucht and S. Jonasson, Inhaled sulfur dioxide causes pulmonary and systemic inflammation leading to fibrotic respiratory disease in a rat model of chemical-induced lung injury, Toxicology, 2016, 368, 28–36 CrossRef PubMed.
- D. B. Peden, Mechanisms of pollution-induced airway disease: in vivo studies, Allergy, 1997, 52, 37–44 CrossRef.
- B. Z. Simkhovich, M. T. Kleinman and R. A. Kloner, Air pollution and cardiovascular injury: epidemiology, toxicology, and mechanisms, J. Am. Coll. Cardiol., 2008, 52(9), 719–726 CrossRef PubMed.
- R. D. Brook, Cardiovascular effects of air pollution, Clin. Sci., 2008, 115(6), 175–187 CrossRef.
- L. Curtis, W. Rea, P. Smith-Willis, E. Fenyves and Y. Pan, Adverse health effects of outdoor air pollutants, Environ. Int., 2006, 32(6), 815–830 CrossRef PubMed.
- S. B. Sharma, S. Jain, P. Khirwadkar and S. Kulkarni, The effects of air pollution on the environment and human health, Indian J. Res. Pharm. Biotechnol., 2013, 1(3), 391–396 Search PubMed.
- P. M. Irving and J. E. Miller, Productivity of field-grown soybeans exposed to acid rain and sulfur dioxide alone and in combination, Agron. J., 1981, 10(4), 473–478 Search PubMed.
- A. Singh and M. Agrawal, Acid rain and its ecological consequences, J. Environ. Biol., 2007, 29(1), 15 Search PubMed.
- P. N. Sudha, K. Sangeetha, K. Vijayalakshmi and A. Barhoum, Nanomaterials history, classification, unique properties, production and market. in Emerging Applications of Nanoparticles and Architecture Nanostructures, Elsevier,  2018, pp. 341–384 Search PubMed.
- J. K. Patel, A. Patel and D. Bhatia, Introduction to nanomaterials and nanotechnology. in Emerging Technologies for Nanoparticle Manufacturing, Cham, Springer Int Publ,  2021, pp. 3–23 Search PubMed.
- M. A. Ratner and D. Ratner, Nanotechnology: A Gentle Introduction to the Next Big Idea, Upper Saddle River, Prentice Hall Prof,  2003 Search PubMed.
- J. Shen, Y. Zhu, H. Jiang and C. Li, 2D nanosheets-based novel architectures: synthesis, assembly and applications, Nano Today, 2016, 11(4), 483–520 CrossRef.
- S. Saha, S. Bansal and M. Khanuja, Classification of nanomaterials and their physical
and chemical nature, in Nano-enabled Agrochemicals in Agriculture, Academic Press,  2022, pp. 7–34 Search PubMed.
- A. Mazumder, S. Bhattacharya and C. Bhattacharjee, Role of nano-photocatalysis in heavy metal detoxification, in Nanophotocatalysis and Environmental Applications: Detoxification and Disinfection, Elsevier,  2020, pp. 1–33 Search PubMed.
- W. Ramadan, Y. AlSalka, O. Al-Madanat and D. W. Bahnemann, Synthesis of magnetic ferrite and TiO2-based nanomaterials for photocatalytic water splitting applications, in Synthesis and Applications of Nanomaterials and Nanocomposites, Singapore, Springer Nature Singapore,  2023, pp. 293 - 329 Search PubMed.
- J. V. Gajjar and D. R. Roy, Porphyrin and its co-doped complex-based sensor devices for nitrogen and sulfur containing environmental toxic gases, IEEE Sens. J., 2024, 24(7), 9411–9418 Search PubMed.
- J. Yan, B. Zhang, S. Guo and Z. Wang, Porphyrin-based nanoporous organic polymers for adsorption of carbon dioxide, ethane, and methane, ACS Appl. Nano Mater., 2021, 4(10), 10565–10574 CrossRef.
- Y. Hong, V. Rozyyev and C. T. Yavuz, Alkyl-linked porphyrin porous polymers for gas capture and precious metal adsorption, Small Sci., 2021, 1(6), 2000078 CrossRef.
- H. Van Ngoc and K. D. Pham, First-principles study on N2, H2, O2, NO, NO2, CO, CO2, and SO2 gas adsorption properties of the Sc2CF2 monolayer, Phys. E, 2022, 141, 115–162 Search PubMed.
- L. K. Saini and M. Pandey, Porphyrin-based gas sensors: a computational investigation of CO2 and COCl2 detection using pure and metallic derivatives, Phys. Scr., 2025, 100(7), 075922 CrossRef.
- D. Akhmetsadyk, A. Ilyin, N. Guseinov and G. Beall, Adsorption of SO2 molecule on pristine, N-, Ga-doped and Ga–N co-doped graphene: a DFT study, Computation, 2023, 11(12), 235 CrossRef.
- L. Lin, C. Xue, X. Li, H. Tao and L. Su, Adsorption and sensing of NO2, SO2, and NH3 on a Janus MoSeTe monolayer decorated with transition metals (Fe, Co, and Ni): a first-principles study, Langmuir, 2023, 39(36), 12662–12670 CrossRef.
- M. Faghihnasiri and P. S. Branicio, Selective adsorption of CO2 over N2 on graphene: insights from first-principles calculations, J. Phys. Chem. C, 2025, 129(5), 2851–2864 CrossRef.
- G. Qu, Z. Ma and T. Jia, Influence of hydroxyl groups on the oxidative reaction characteristics of active groups in lignite at room temperature, ACS Omega, 2024, 9(14), 16237–16248 CrossRef PubMed.
- J. Zhang and T. Lu, Efficient evaluation of electrostatic potential with computerized optimized code, Phys. Chem. Chem. Phys., 2021, 23(36), 20323–20328 RSC.
- A. R. Kumar, L. Ilavarasan, G. S. Mol, S. Selvaraj, M. Azam and P. Jayaprakash, et al., Spectroscopic (FT-IR, FT-Raman, UV–Vis and NMR) and computational (DFT, MESP, NBO, NCI, LOL, ELF, RDG and QTAIM) profiling of 5-chloro-2-hydroxy-3-methoxybenzaldehyde: a promising antitumor agent, J. Mol. Struct., 2024, 1298, 136–974 Search PubMed.
- T. Mackoy, B. Kale, M. E. Papka and R. A. Wheeler, viewSq, a Visual Molecular Dynamics (VMD) module for calculating, analyzing, and visualizing X-ray and neutron structure factors from atomistic simulations, Comput. Phys. Commun., 2021, 264, 107–881 CrossRef.
- R. A. Massoud and M. A. Makhyoun, Theoretical study of the interaction of ethanol with the (3,5-dimethylpyrazole-N2)(N-salicylidenephenyl-alaninato-O,N,O′) copper(II) complex, J. Struct. Chem., 2019, 60, 882–889 CrossRef.
- M. R. Poopari, Z. Dezhahang and Y. Xu, Stereochemical properties of multidentate nitrogen donor ligands and their copper complexes by electronic CD and DFT, Chirality, 2016, 28(7), 545–555 CrossRef.
- S. Ghoshal, P. Roy, A. Pramanik and P. Sarkar, Ru/Rh catalyzed selective hydrogenation of CO2 to formic acid: a first principles microkinetics analysis, Catal. Sci. Technol., 2022, 12(23), 7219–7232 RSC.
- M. Abdel-Megid, O. M. Adly, M. A. Ibrahim and R. Fouad, New metal complexes based on chromone and triazine moieties as potential antitumor agents: full structural elucidation, theoretical calculations, and biological
studies, Appl. Organomet. Chem., 2025, 39(6), e70202 CrossRef.
- Q. Yang, D. Liu, C. Zhong and J. R. Li, Development of computational methodologies for metal–organic frameworks and their application in gas separations, Chem. Rev., 2013, 113(10), 8261–8323 CrossRef.
- Y. Zhou, Dilithium and Disodium Salts of Pseudoephedrine-Derived Myers Enolates: Stereoselectivity and Mechanism of Alkylation [dissertation], Ithaca (NY), Cornell Univ,  2019 Search PubMed.
- B. A. Ademola, First-Principles Density Functional Theory Studies of the Structural and Electronic Properties of Prodigiosin as a Cancer Drug, PhD Thesis, African University of Science and Technology, Abuja,  2019.
- H. Hadi, H. Louis, T. E. Gber and F. O. Ogungbemiro, Molecular modeling of the structural, electronic, excited state dynamic, and photovoltaic properties of oligomers of n-corannulene (n = 1–4), Heliyon, 2023, 9(10), e2230 CrossRef.
- H. Y. Ammar and H. M. Badran, Effect of CO adsorption on properties of transition metal doped porphyrin: a DFT and TD-DFT study, Heliyon, 2019, 5(10), e02540 CrossRef.
- J. Abdul Nasir, A. Munir, N. Ahmad, T. U. Haq, Z. Khan and Z. Rehman, Photocatalytic Z-scheme overall water splitting: recent advances in theory and experiments, Adv. Mater., 2021, 33(52), 2105–2195 CrossRef.
- A. U. Hassan and S. H. Sumrra, Exploration of pull–push effect for novel photovoltaic dyes with A–π–D design: a DFT/TD-DFT investigation, J. Fluoresc., 2022, 32(6), 1999–2014 CrossRef.
- R. G. Pearson, Chemical hardness and density functional theory, J. Chem. Sci., 2005, 117, 369–377 CrossRef.
- J. Wang, L. N. Guo, W. M. Lin, J. Chen, S. Zhang, T. T. Zhen and Y. Y. Zhang, The effects of graphene content on the corrosion resistance, and electrical, thermal and mechanical properties of graphene/copper composites, New Carbon Mater., 2019, 34(2), 161–169 CrossRef CAS.
- R. K. Castellano, Progress toward understanding the nature and function of CH···O interactions, Curr. Org. Chem., 2004, 8(10), 845–865 CrossRef CAS.
- M. Khalid, G. Wu, A. Umber, M. Khan, A. A. Braga, N. Alhokbany and S. C. Ojha, Impact of extended π-conjugated acceptor moieties on non-linear optical behavior in polar medium for designed 4,4-dimethyl-4H-indeno[1,2-b]thiophene-based materials, J. Mol. Liq., 2024, 401, 124608 CrossRef CAS.
- B. Szcześniak, J. Choma and M. Jaroniec, Gas adsorption properties of graphene-based materials, Adv. Colloid Interface Sci., 2017, 243, 46–59 CrossRef.
- A. Otero-De-La-Roza, E. R. Johnson and J. Contreras-García, Revealing non-covalent interactions in solids: NCI plots revisited, Phys. Chem. Chem. Phys., 2012, 14(35), 12165–12172 RSC.
- C. Tantardini, A Study of Chemical Bonding through Quantum Chemical Topology, PhD Thesis, Skolkovo Institute of Science and Technology,  2020.
- M. R. Xavier, M. M. Marinho, M. S. Julião, E. S. Marinho, F. W. Almeida-Neto, K. K. de Castro, J. P. da Hora, M. N. da Rocha, A. C. Barreto, G. D. Saraiva and P. N. Bandeira, Structural, topological, vibrational, and electronic analysis, and ADMET study of methyl-2-(4-isobutylphenyl) propanoate, J. Mol. Struct., 2024, 1309, 138019 CrossRef CAS.
- A. J. Irving, Geochemical and High-Pressure Experimental Studies of Xenoliths, Megacrysts and Basalts from Southeastern Australia, PhD Thesis, Canberra, Aust Natl Univ,  1971 Search PubMed.
- S. J. Grabowski, What is the covalency of hydrogen bonding?, Chem. Rev., 2011, 111(4), 2597–2625 CrossRef PubMed.
- A. M. Donald, A. H. Windle and S. Hanna, Liquid Crystalline Polymers. Cambridge, Cambridge Univ Press,  2006 Search PubMed.
- K. S. Kim, P. Tarakeshwar and J. Y. Lee, Molecular clusters of π-systems: theoretical studies of structures, spectra, and origin of interaction energies, Chem. Rev., 2000, 100(11), 4145–4186 CrossRef.
- Y. Shah, M. T. Hussain, A. Mansha and M. R. S. A. Janjua, Exploring the potential of quinacridone-porphyrin-based materials for nitrogen oxides sensing: quantum chemical design, mechanism and future prospects, Adsorption, 2024, 30(8), 2053–2067 CrossRef.
- K. R. S. Chandrakumar and S. K. Ghosh, Alkali-metal-induced enhancement of hydrogen adsorption in C60 fullerene: an ab initio study, Nano Lett., 2008, 8(1), 13–19 CrossRef PubMed.
- E. Mete, M. Yortanlı and M. F. Danışman, A van der Waals DFT study of chain length dependence of alkanethiol adsorption on Au(111): physisorption vs. chemisorption, Phys. Chem. Chem. Phys., 2017, 19(21), 13756–13766 RSC.
- Y. Guo, K. Xu, C. Wu, J. Zhao and Y. Xie, Surface chemical modification for engineering the intrinsic physical properties of inorganic two-dimensional nanomaterials, Chem. Soc. Rev., 2015, 44(3), 637–646 RSC.
- N. Mercier, S. Poiroux, A. Riou and P. Batail, Unique hydrogen bonding correlating with a reduced band gap and phase transition in the hybrid perovskites (HO(CH2)2NH3)2PbX4 (X = I, Br), Inorg. Chem., 2004, 43(26), 8361–8366 CrossRef PubMed.
- R. Meenakshi, Spectral investigations, DFT-based global reactivity descriptors, inhibition efficiency and analysis of 5-chloro-2-nitroanisole as π-spacer with donor–acceptor variations effect for DSSC performance, J. Mol. Struct., 2017, 1127, 694–707 CrossRef.
- R. Vuilleumier and M. Sprik, Electronic properties of hard and soft ions in solution: aqueous Na+ and Ag+ compared, J. Chem. Phys., 2001, 115(8), 3454–3468 CrossRef.
- H. Wei, Y. Gui, J. Kang, W. Wang and C. Tang, A DFT study on the adsorption of H2S and SO2 on Ni-doped MoS2 monolayer, Nanomaterials, 2018, 8(9), 646 CrossRef.
- Y. Cai, G. Zhang and Y. W. Zhang, Charge transfer and functionalization of monolayer InSe by physisorption of small molecules for gas sensing, J. Phys. Chem. C, 2017, 121(18), 10182–10193 CrossRef.
- J. Z. Ou, W. Ge, B. Carey, T. Daeneke, A. Rotbart and W. Shan, et al., Physisorption-based charge transfer in two-dimensional SnS2 for selective and reversible NO2 gas sensing, ACS Nano, 2015, 9(10), 10313–10323 CrossRef PubMed.
- A. M. Gutiérrez-Vílchez, C. V. Ileperuma, V. Navarro-Pérez, P. A. Karr, F. Fernández-Lázaro and F. D’Souza, Excited Charge Transfer Promoted Electron Transfer in all Perylenediimide Derived, Wide-Band Capturing Conjugates: A Mimicry of the Early Events of Natural Photosynthesis, ChemPlusChem, 2024, 89(11), e202400348 CrossRef PubMed.
- D. Cappelletti, E. Ronca, L. Belpassi, F. Tarantelli and F. Pirani, Revealing charge-transfer effects in gas-phase water chemistry, Acc. Chem. Res., 2012, 45(9), 1571–1580 CrossRef CAS PubMed.
| 
 | 
| This journal is © The Royal Society of Chemistry 2025 | 
Click here to see how this site uses Cookies. View our privacy policy here.