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

Aluminium phosphide (Al12P12) nanocage as a potential sensor for volatile organic compounds: A DFT study

Mahmoud A. A. Ibrahim*ab, Manar H. A. Hamada, Nayra A. M. Moussaa, Omar H. Abd-Elkaderc, Shaban R. M. Sayedd, Muhammad Naeem Ahmede, Ahmed M. Awadf and Tamer Shoeib*g
aComputational Chemistry Laboratory, Chemistry Department, Faculty of Science, Minia University, Minia 61519, Egypt. E-mail: m.ibrahim@compchem.net
bSchool of Health Sciences, University of KwaZulu-Natal, Westville Campus, Durban 4000, South Africa
cDepartment of Physics and Astronomy, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
dDepartment of Botany and Microbiology, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
eDepartment of Chemistry, The University of Azad Jammu and Kashmir, Muzaffarabad 13100, Pakistan
fDepartment of Chemistry, California State University Channel Islands, Camarillo, California 93012, USA
gDepartment of Chemistry, The American University in Cairo, New Cairo 11835, Egypt. E-mail: t.shoeib@aucegypt.edu

Received 9th March 2024 , Accepted 19th April 2024

First published on 29th April 2024


Abstract

The efficacy of aluminium phosphide (Al12P12) nanocage toward sensing methanol (MeOH) and ethanol (EtOH) volatile organic compounds (VOCs) was herein thoroughly elucidated utilizing various density functional theory (DFT) computations. In this perspective, MeOH⋯ and EtOH⋯Al12P12 complexes were investigated within all plausible configurations. According to the energetic features, the EtOH⋯Al12P12 complexes exhibited larger negative values of adsorption and interaction energies with values up to −27.23 and −32.84 kcal mol−1, respectively, in comparison to the MeOH⋯Al12P12 complexes. Based on the symmetry-adapted perturbation theory (SAPT) results, the electrostatic forces were pinpointed as the predominant component beyond the adsorption process within the preferable MeOH⋯ and EtOH⋯Al12P12 complexes. The findings of the noncovalent interaction (NCI) index and quantum theory of atoms in molecules (QTAIM) outlined the closed-shell nature of the interactions within the studied complexes. Substantial variations were found in the molecular orbitals distribution patterns of MeOH/EtOH molecules and Al12P12 nanocage, outlining the occurrence of the adsorption process within the complexes under investigation. Thermodynamic parameters were denoted with negative values, demonstrating the spontaneous exothermic nature of the most favorable complexes. New energy states were observed within the extracted density of states plots, confirming the impact of adsorbing MeOH and EtOH molecules on the electronic properties of the Al12P12 nanocage. The appearance of additional peaks in Infrared Radiation (IR) and Raman spectra revealed the apparent effect of the adsorption process on the features of the utilized sensor. The emerging results declared the potential uses of Al12P12 nanocage as a promising candidate for sensing VOCs, particularly MeOH and EtOH.


1. Introduction

In the contemporary world, scientists have focused significantly on developing sustainable nanomaterials for various applications, including energy, environment, and drug delivery. Indeed, the continued life of our planet depends significantly on advancements in sophisticated materials science. Several nano-based structures have been accordingly developed and were widely investigated, comprising fullerenes,1 nanotubes,2 and nanocones.3

More recently, the utilization of phosphide nanocages has attracted the attention of scientists.4–6 It is worth mentioning that inorganic aluminium phosphide (Al12P12) nanocage was earlier distinguished by its distinctive chemical features, including a high energy gap and low electron attraction.7–9 The potential applications of Al12P12 in non-linear optics,10 drug carriers,11,12 and sensors13,14 have garnered tremendous attention. As an appropriate sensing material, the adsorption of cyanogen chloride and hydrogen cyanide toxic gases on the surface of Al12P12 was investigated.15 Moreover, the sensitivity of the Al12P12 nanocarrier toward detecting phosgene gas was revealed.16

The increase in energy consumption and industrial activities made monitoring air pollution an essential priority for several countries and organizations.17–23 As a point of departure, volatile organic compounds (VOCs) are categorized as organic compounds with low water solubility, low boiling points, and high vapor pressure.24,25 Detailedly, VOCs are considered hazardous air pollutants and are among the most frequent air pollutants released by industrial chemical processes, standard household products, and construction materials.26–28 The natural environment and human health are significantly threatened by VOCs, which act as precursors to ozone and photochemical smog.29–31 VOCs are the main contributor to the greenhouse effect and also have the possibility to damage the human nervous and circulatory systems.32,33 Considering the risks of VOCs exposure, these compounds need to be eliminated from the environment. Several VOCs purification techniques have been developed, including adsorption,34,35 biodegradation,36 and membrane separation.37 Numerous alcohols, including methanol (MeOH), ethanol (EtOH), isopropanol, ethylene glycol, etc.,38 are commonly found as VOCs in various indoor air conditions. The adsorption amplitude of the MoSe2 nanosheet and carbon nanopores toward adsorbing the MeOH and EtOH was earlier divulged.39,40 Notwithstanding the promising properties of aluminium-bearing nanocages, no solid investigation revealed their efficiency in detecting the MeOH and EtOH molecules.

In this regard, the principal purpose of this study was to elucidate the potentiality of Al12P12 toward sensing the MeOH and EtOH molecules within all plausible configurations of the MeOH⋯ and EtOH⋯Al12P12 complexes. Geometrical optimization and frequency computations were executed for all the investigated systems, accompanied by adsorption and interaction energy calculations. To shed light on the physical forces dominating the adsorption process, the SAPT method was employed. Subsequently, the thermodynamic features, global indices of reactivity, and electronic parameters were assessed. This study will provide significant principles for the design and enhancement of Al12P12 nanocage applications in detecting toxic molecules, particularly for VOCs.

2. Results and discussion

2.1. ESP analysis

Molecular electrostatic potential (MEP) maps were portrayed for the optimized structures to clarify the electrophile and nucleophile sites, as recommended in literature.41 Different colors were utilized to depict the difference in ESP at the molecular surface. In the colored scale, the red/orange/yellow, green, and blue colors-refer to electron-rich, neutral, and electron-deficient sites, respectively. A quantitative insight was subsequently provided by computing the surface electrostatic potential extrema values (Vs,min/Vs,max) over the surface of the optimized monomers. Fig. 1 illustrates the MEP maps and Vs,min/Vs,max values.
image file: d4ra01828a-f1.tif
Fig. 1 Optimized structures of VOCs and Al12P12 nanocage combined with the MEP maps and Vs,min/Vs,max values (in kcal mol−1).

As displayed in Fig. 1, the pure Al12P12 nanocage exhibits Th symmetry and is composed of eight hexagonal rings in combination with six tetragonal rings.42 At first glance, two different bonds in the Al12P12 nanocage were denoted. The first Al⋯P bond is shared by tetragonal and hexagonal rings with an average bond length of 2.28 Å. Meanwhile, the other Al⋯P bond is located between two hexagonal rings with a bond distance of 2.33 Å. The MEP plots of the MeOH and EtOH clarified the nucleophilic site at the region enclosing the O atoms. Such pictorial outcomes were ensured by the existence of negative Vs,min values of −42.8 and −42.5 kcal mol−1 over the surface of the MeOH and EtOH, respectively. At the same time, the electrophilic sites were observed via the existence of blue-colored regions around C and H atoms in the MeOH and EtOH molecules. On the surface of the Al12P12 nanocage, the red-colored nucleophilic and blue-colored electrophilic regions were around P and Al atoms with Vs,min and Vs,max values of −8.7 and 59.2 kcal mol−1, respectively.

2.2. Adsorption features

To gain an extensive comprehension of the adsorption process, the VOCs were placed on the surface of the Al12P12 nanocage. Geometrical optimization calculations were executed for the VOC⋯Al12P12 complexes within all plausible configurations (Fig. 2). No imaginary frequency was identified, confirming that the optimized structures are true minima. On the optimized VOC⋯Al12P12 complexes, the MEP maps were extracted and are displayed in Fig. 2. The adsorption (Eads) and interaction (Eint) energies were accordingly assessed. Table 1 shows the computed Eads and Eint values along with the intermolecular distances between the VOC molecule and Al12P12 nanocage.
image file: d4ra01828a-f2.tif
Fig. 2 Optimized structures and MEP maps of the MeOH⋯ and EtOH⋯Al12P12 complexes within all plausible configurations. The intermolecular distances are in Å.
Table 1 Calculated values of the adsorption energies (Eads, kcal mol−1) and the interaction energies (Eint, kcal mol−1) of the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within all plausible configurations in conjunction with intermolecular distances (d, Å)
Complexes Configuration Bond d Eads Eint
MeOH⋯Al12P12 A1 O⋯Al 1.97 −26.01 −30.76
A2 O⋯Al 1.97 −25.37 −30.21
A3 H⋯Al 2.64 −3.66 −3.93
H⋯P 4.45
A4 H⋯Al 2.96 −3.66 −3.91
H⋯P 3.35
A5 Al⋯H 2.78 −1.90 −1.97
EtOH⋯Al12P12 B1 O⋯Al 1.95 −27.23 −32.84
H⋯P 3.13
B2 O⋯Al 1.97 −26.59 −32.45
H⋯P 2.94
B3 O⋯Al 1.95 −27.03 −32.39
B4 O⋯Al 1.95 −26.44 −32.17
H⋯P 3.14
B5 O⋯Al 1.96 −26.21 −31.50
B6 H⋯Al 2.69 −4.47 −4.91
H⋯P 2.76
B7 H1⋯P1 2.65 −4.68 −4.84
H2⋯P2 3.30
B8 H⋯Al 2.94 −2.81 −2.91
H⋯P 3.20
B9 C⋯Al 3.26 −2.44 −2.52
B10 C⋯Al 3.26 −2.43 −2.51


According to data presented in Table 1, the intermolecular distances were observed with values ranging from 4.45 to 1.97 and 3.30 to 1.95 Å for the optimized MeOH⋯ and EtOH⋯Al12P12 complexes, respectively. Notably, the VOC⋯Al12P12 complexes exhibited significant negative Eads and Eint values, ensuring the efficacy of the Al12P12 nanocage toward adsorbing VOC molecules. The EtOH⋯Al12P12 complexes showed higher negative Eads and Eint values relative to the MeOH⋯Al12P12 complexes. Numerically, Eads/Eint of the interactions within the MeOH⋯ and EtOH⋯Al12P12 complexes showed values ranging from −1.90/−1.97 to −26.01/−30.76 and from −2.43/−2.51 to −27.23/−32.84 kcal mol−1, respectively. It is worth noting that the selectivity of Al12P12 nanocage toward adsorbing EtOH over MeOH molecules is not guaranteed where the energy differences between MeOH⋯ and EtOH⋯Al12P12 complexes is about 2 kcal mol−1.

For the sake of comparison, more favorability was denoted in the case of configurations A1 ↔ A2 and B1 ↔ B5 within the MeOH⋯ and EtOH⋯Al12P12 complexes, respectively. In the abovementioned configurations, the investigated VOCs were adsorbed on the surface of the Al12P12 nanocage via the interactions of their O atoms and the nanocage's Al atom. This finding was in line with the ESP results (Fig. 2) that confirmed the predominant nucleophilic character around the O atom in the VOCs. It was noticeable that the most preferred MeOH⋯ and EtOH⋯Al12P12 complexes were denoted in the case of the configurations A1 and B1, which had Eads values of −26.01 and −27.23 kcal mol−1, respectively.

2.3. SAPT calculations

To unveil the contributions of the physical energetic components to the inspected adsorption process, SAPT analysis was carried out for the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within all plausible configurations. Fig. 3 illustrates the four main components of the total SAPT0 energy, namely induction (Eind), electrostatic (Eelst), exchange (Eexch), and dispersion (Edisp).
image file: d4ra01828a-f3.tif
Fig. 3 Graphical representation demonstrating the energetic components of total SAPT0 energies of the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within all plausible configurations.

Looking at Fig. 3, the Eelst, Eind, and Edisp exhibited negative values, revealing their favorable role as attractive forces between the interacted species within the inspected complexes. It is worth mentioning that the Eelst forces were dominant within the most preferable configurations of the VOC⋯Al12P12 complexes. Such findings could be attributed to the interaction of the electron-rich oxygen atom in MeOH and EtOH with the electron-deficient aluminium atom in the Al12P12 nanocage. For instance, the Eelst values of the configurations A1 and B1 were −62.18 and −63.84 kcal mol−1, respectively. For the other configurations (i.e., A3 ↔ A5 and B6 ↔ B10), the Edisp component exhibited notable contributions to the overall attractive forces beyond the occurrence of the adsorptions process. While the Eexch component was found with positive values, ensuring its unfavorable contribution to the adsorption process of the VOCs onto the surface of the Al12P12 nanocage.

2.4. QTAIM and NCI analyses

QTAIM and NCI index analyses were employed to unveil an in-depth elucidation of the nature and origin of the interactions within the investigated complexes (Fig. 4). As demonstrated in Fig. 4, the occurrence of the adsorption process within the optimized MeOH⋯ and EtOH⋯Al12P12 complexes was assured by the presence of bond paths (BPs) and bond critical points (BCPs) between the interacted species.
image file: d4ra01828a-f4.tif
Fig. 4 QTAIM and 3D NCI plots of the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within all plausible configurations.

To better comprehend the interaction of VOCs with the Al12P12 nanocage, total energy density (Hb), electron density (ρb), Laplacian (∇2ρb), kinetic electron density (Gb), local potential electron energy density (Vb), and the negative ratio of kinetic and potential electron energy density (−Gb/Vb) were computed at bond critical points and tabulated in Table 2. From the summarized data in Table 2, the positive ∇2ρb and Hb values asserted the closed-shell nature of the interactions within the complexes under investigation. For instance, the Hb/∇2ρb values of the optimized MeOH⋯ and EtOH⋯Al12P12 complex within configuration A1 and B1 were 0.0040/0.3198 and 0.0051/0.3370 au, respectively.

Table 2 Topological features at BCPs of the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within all the plausible configurations. All parameters are provided in au
Complexes Configuration Bond ρb Hb 2ρb Gb Vb Gb/Vb
MeOH⋯Al12P12 A1 O⋯Al 0.0505 0.0040 0.3198 0.0759 −0.0719 1.0559
A2 O⋯Al 0.0499 0.0042 0.3180 0.0753 −0.0711 1.0590
A3 H⋯Al 0.0086 0.0001 0.0158 0.0040 −0.0040 1.0005
H⋯P 0.0119 0.0008 0.0348 0.0079 −0.0071 1.1113
A4 H⋯Al 0.0610 0.0001 0.1233 0.0037 −0.0036 1.0310
H⋯P 0.0610 0.0008 0.1234 0.0078 −0.0069 1.1184
A5 Al⋯H 0.0074 0.0002 0.0148 0.0035 −0.0032 1.0743
EtOH⋯Al12P12 B1 O⋯Al 0.0513 0.0051 0.3370 0.0792 −0.0741 1.0682
H⋯P 0.0058 0.0010 0.0172 0.0033 −0.0024 1.4054
B2 O⋯Al 0.0503 0.0039 0.3171 0.0754 −0.0716 1.0539
H⋯P 0.0089 0.0013 0.0280 0.0057 −0.0043 1.3059
B3 O⋯Al 0.0514 0.0047 0.3346 0.0789 −0.0742 1.0633
B4 O⋯Al 0.0509 0.0051 0.3345 0.0785 −0.0734 1.0696
H⋯P 0.0057 0.0009 0.0170 0.0033 −0.0024 1.4001
B5 O⋯Al 0.0510 0.0045 0.3295 0.0779 −0.0733 1.0618
B6 H⋯Al 0.0082 0.0002 0.0162 0.0038 −0.0036 1.0635
H⋯P 0.0095 0.0010 0.0298 0.0064 −0.0054 1.1933
B7 H1⋯P1 0.0115 0.0008 0.0340 0.0077 −0.0069 1.1193
H2⋯P2 0.0050 0.0007 0.0144 0.0029 −0.0021 0.7427
B8 H⋯Al 0.0067 0.0001 0.0137 0.0033 −0.0031 1.0446
H⋯P 0.0052 0.0009 0.0153 0.0029 −0.0020 1.4400
B9 C⋯Al 0.0064 0.0001 0.0137 0.0033 −0.0032 0.9681
B10 C⋯Al 0.0065 0.0001 0.0138 0.0033 −0.0032 0.9684


Looking at 3D NCI isosurfaces displayed in Fig. 4, various types of interactions are highlighted by different colored isosurfaces; blue demonstrates a stronger hydrogen bond, green represents van der Waals interactions, and red confirms steric effects. The existence of blue-green colored isosurfaces between VOCs and the surface of the Al12P12 nanocage within the investigated complexes shed light on the propensity of the Al12P12 nanocage toward sensing the inspected VOCs.

2.5. Electronic parameters

With the incorporation of frontier molecular orbital (FMO) theory, the electronic parameters and the distribution of the molecular orbitals were outlined for the VOCs and Al12P12 before and after the adsorption process. In this regard, the energies of the highest occupied molecular orbitals (EHOMO), the lowest unoccupied molecular orbitals (ELUMO), Fermi level (EFL), and energy gap (Egap) were determined to unveil the capability of the inspected systems within monomeric and complex forms to donate and accept electrons (Table 3). Fig. 5 and 6 illustrate the distribution patterns of the molecular orbitals of the studied systems within the monomeric and complex forms, respectively.
Table 3 Computed electronic parameters (in eV) of the optimized VOCs and Al12P12 nanocage within the monomeric and complex forms
System Configuration EHOMO EFL ELUMO Egap
MeOH   −9.577 −4.523 0.531 10.108
EtOH   −9.447 −4.472 0.502 9.949
Al12P12   −7.755 −5.389 −3.024 4.731
MeOH⋯Al12P12 A1 −7.430 −5.063 −2.695 4.735
A2 −7.424 −5.054 −2.683 4.742
A3 −7.842 −5.473 −3.104 4.737
A4 −7.843 −5.476 −3.109 4.735
A5 −7.766 −5.402 −3.038 4.727
EtOH⋯Al12P12 B1 −7.398 −5.029 −2.660 4.738
B2 −7.400 −5.036 −2.672 4.728
B3 −7.393 −5.024 −2.654 4.739
B4 −7.384 −5.020 −2.656 4.729
B5 −7.389 −5.020 −2.651 4.738
B6 −7.836 −5.469 −3.101 4.736
B7 −7.851 −5.484 −3.118 4.733
B8 −7.723 −5.359 −2.994 4.729
B9 −7.757 −5.395 −3.034 4.724
B10 −7.757 −5.393 −3.030 4.726



image file: d4ra01828a-f5.tif
Fig. 5 Plots of the distribution patterns of HOMO and LUMO of the MeOH, EtOH, and Al12P12 monomers.

image file: d4ra01828a-f6.tif
Fig. 6 Plots of the distributions patterns of HOMO and LUMO of the optimized MeOH⋯ and EtOH⋯Al12P12 complexes in all the plausible configurations.

As illustrated in Fig. 5, the HOMO and LUMO distribution patterns were observed over the electronegative and electropositive regions of the studied VOCs. Considering the Al12P12 nanocage, the P and Al atoms were generally noticed with distributions of HOMO and LUMO orbitals, respectively. Following the interactions of the adopted VOCs with Al12P12, redistribution of HOMO and LUMO orbitals was denoted, highlighting the occurrence of the adsorption process (Fig. 6). On the investigated complexes, the HOMO and LUMO levels were found over the Al12P12 nanocage.

Upon the listed data in Table 3, notable changes in the EHOMO, EFL, ELUMO, and Egap values following the adsorption of the VOCs on the Al12P12 nanocage were denoted, confirming the occurrence of the adsorption process. For example, the EHOMO of the pure Al12P12 nanocage showed a value of −7.755 eV and changed to −7.430 and −7.398 eV after adsorbing MeOH and EtOH within the configurations A1 and B1, respectively. An apparent alteration in the Egap values of the studied systems was also detected, outlining the prominent effect of the adsorption process of the MeOH and EtOH on the surface of the Al12P12 nanocage. As an illustration, the Egap value of pure Al12P12 nanocage was 4.731 eV, which changed to 4.735 and 4.738 eV following the adsorption process within the configurations A1 and B1 of the MeOH⋯ and EtOH⋯Al12P12 complexes, respectively.

2.6. Global reactivity descriptors

In an attempt to clarify the effect of the adsorption process of the MeOH and EtOH molecules on the Al12P12 nanocage, global reactivity descriptors of the monomeric and complex forms of the studied systems were evaluated. Numerous parameters, comprising ionization potential (IP), electron affinity (EA), chemical potential (μ), global hardness (η), global softness (S), electrophilicity index (ω), and work function (Φ), were calculated and are compiled in Table 4.
Table 4 Global indices descriptors of the monomeric and complex forms of the investigated VOCs and Al12P12 nanocage
System Configuration IP (eV) EA (eV) μ (eV) η (eV) S (eV−1) ω (eV) Φ (eV)
MeOH   9.577 −0.531 −4.523 5.054 0.198 2.024 4.523
EtOH   9.447 −0.502 −4.472 4.975 0.201 2.010 4.472
Al12P12   7.755 3.024 −5.389 2.366 0.423 6.139 5.389
MeOH⋯Al12P12 A1 7.430 2.695 −5.063 2.368 0.422 5.413 5.063
A2 7.424 2.683 −5.054 2.371 0.422 5.386 5.054
A3 7.842 3.104 −5.473 2.369 0.422 6.323 5.473
A4 7.843 3.109 −5.476 2.367 0.422 6.334 5.476
A5 7.766 3.038 −5.402 2.364 0.423 6.173 5.402
EtOH⋯Al12P12 B1 7.398 2.660 −5.029 2.369 0.422 5.338 5.029
B2 7.400 2.672 −5.036 2.364 0.423 5.364 5.036
B3 7.393 2.654 −5.024 2.370 0.422 5.326 5.024
B4 7.384 2.656 −5.020 2.364 0.423 5.329 5.020
B5 7.389 2.651 −5.020 2.369 0.422 5.319 5.020
B6 7.836 3.101 −5.469 2.368 0.422 6.315 5.469
B7 7.851 3.118 −5.484 2.367 0.423 6.355 5.484
B8 7.723 2.994 −5.359 2.364 0.423 6.072 5.359
B9 7.757 3.034 −5.395 2.362 0.423 6.163 5.395
B10 7.757 3.030 −5.393 2.363 0.423 6.155 5.393


Relying on the summarized data in Table 4, substantial alterations in the values of global reactivity descriptors of the Al12P12 before and following the adsorption process were observed, outlining the influence of the adsorption process on the reactivity character of the utilized nanocage. For instance, the IP value of pure Al12P12 nanocage was 7.755 eV and altered to 7.430 and 7.398 eV following interaction with VOCs within configurations A1 and B1, respectively. Apparently, upward and downward shifts in the η and S values were observed following the adsorption process. As numerical evidence, η of pure Al12P12 nanocage was 2.366 eV and boosted to 2.368 and 2.369 eV for the configurations A1 and B1, respectively. Remarkably, the alterations in work function affirmed the potency of Al12P12 nanocage as a promising sensing material for MeOH and EtOH molecules.

2.7. DOS analysis

DOS analysis was executed to unveil the change in the electronic characteristics of the A12P12 nanocage after the adsorption of MeOH and EtOH molecules. Fig. 7 and 8 depict the DOS plots of the Al12P12 nanocage before and following the adsorption process within all plausible configurations of the VOC⋯Al12P12 complexes, respectively.
image file: d4ra01828a-f7.tif
Fig. 7 The DOS plot for the pure Al12P12 nanocage before the adsorption process.

image file: d4ra01828a-f8.tif
Fig. 8 The DOS plots for Al12P12 nanocage following the adsorption process of the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within all plausible configurations.

Notably, new peaks were detected by comparing the DOS plots of the Al12P12 nanocage before and following the adsorption process (Fig. 7 and 8, respectively). This result outlined the influential effect of adsorbing VOCs on the electrical characteristics of the Al12P12 nanocage. For instance, additional peaks in the valence region were denoted from −13.50 to −18.00 eV in the DOS plots of almost all studied configurations. The variations in energy gap values were scrutinized for all studied complexes, confirming the ability of the studied Al12P12 nanocage to sense VOCs with disparate efficiencies.

2.8. Thermodynamic parameters

To gain a thorough understanding of the adsorption process within the MeOH⋯ and EtOH⋯Al12P12 complexes, thermodynamic parameters (i.e., changes in enthalpy (ΔH), Gibbs free energy (ΔG), and entropy (ΔS)) were calculated for all the investigated complexes, and the outcomes are presented in Table 5.
Table 5 Thermodynamic parameters of the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within all plausible configurations are in kcal mol−1
Complex Configuration ΔG ΔH ΔS
MeOH⋯Al12P12 A1 −13.00 −24.32 −0.038
A2 −12.75 −23.74 −0.037
A3 6.03 −2.64 −0.029
A4 6.00 −2.65 −0.029
A5 5.98 −0.86 −0.023
EtOH⋯Al12P12 B1 −14.33 −25.71 −0.038
B2 −13.43 −24.99 −0.039
B3 −14.37 −25.47 −0.037
B4 −13.99 −24.97 −0.037
B5 −13.76 −24.64 −0.036
B6 5.99 −3.47 −0.032
B7 5.96 −3.69 −0.032
B8 6.61 −1.79 −0.028
B9 5.68 −1.32 −0.023
B10 6.42 −1.27 −0.026


According to the data listed in Table 5, the negative values of ΔG confirm the spontaneity of the adsorption process within the most preferable configurations of the VOC⋯Al12P12 complexes. Significantly, the exothermic nature was noticed and confirmed by negative ΔH values for the optimized VOC⋯Al12P12 complexes within all inspected configurations. Remarkably, small negative ΔS values were obtained, unveiling the randomness in all studied complexes. In alignment with the Eads results, configurations A1 and B1 of the VOC⋯Al12P12 complexes showed the highest negative values of thermodynamic energetic quantities. For instance, configuration B1 was thermodynamically stable with negative ΔG, ΔH, and ΔS values of −14.33, −25.71, and −0.038 kcal mol−1, respectively. The abovementioned observations outlined the proficiency of Al12P12 nanocage toward sensing the studied VOCs.

2.9. IR and Raman spectra

To ensure the occurrence of the adsorption process of the MeOH and EtOH molecules on the Al12P12 nanocage, IR and Raman spectra were extracted for pure Al12P12 nanocage (Fig. 9) and the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within all plausible configurations (Fig. S1 and S2). Fig. 10 represents plots of IR and Raman spectra of the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within configurations A1 and B1 as an illustration.
image file: d4ra01828a-f9.tif
Fig. 9 Plots of (a) IR and (b) Raman spectra of pure Al12P12 nanocage.

image file: d4ra01828a-f10.tif
Fig. 10 Plots of (a) IR and (b) Raman spectra of the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within configurations A1 and B1.

As depicted in Fig. 9(a), the most noticeable IR band in pure Al12P12 nanocage was ascribed to Al⋯P stretching that appeared at 550 cm−1. Following the adsorption of VOCs on the Al12P12 nanocage, the Al⋯P stretching vibrations were denoted with distinct changes in the intensities within the studied complexes (Fig. 10(a)). Obviously, new additional bands appeared in all studied complexes, affirming the substantial adsorption of MeOH and EtOH on Al12P12 nanocage (Fig. S1).

Similarly, significant alterations in the Raman spectra were noticed between the pure and complex forms of the Al12P12 nanocage (Fig. 10(b)). Overall, the notable difference in the IR and Raman spectra (Fig. S1 and S2, respectively) announced the potential efficacy of the Al12P12 nanocage in detecting MeOH and EtOH molecules.

2.10. Recovery time

Recovery time (τ) values were computed for the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within all plausible configurations toward a better comprehension of the required time for the VOC to separate from the surface of the Al12P12 nanocage (Table 6).
Table 6 The calculated τ values of the optimized MeOH⋯ and EtOH⋯Al12P12 complexes within all plausible configurations
Complex Configuration τ (ms)
MeOH⋯Al12P12 A1 1.09 × 1010
A2 3.72 × 109
A3 4.78 × 10−7
A4 4.78 × 10−7
A5 2.46 × 10−8
EtOH⋯Al12P12 B1 8.54 × 1010
B2 2.91 × 1010
B3 6.10 × 1010
B4 2.26 × 1010
B5 1.53 × 1010
B6 1.87 × 10−6
B7 2.66 × 10−6
B8 1.14 × 10−7
B9 6.11 × 10−8
B10 6.01 × 10−8


Relying on the recorded data in Table 6, the τ findings were directly proportional to the Eads values, revealing that the time required for VOC to dissociate from the adsorbent surface increased with augmenting Eads value. For the sake of clarification, the highest negative Eads values were ascribed to the EtOH⋯Al12P12 complexes, which were denoted with longer τ values compared to MeOH⋯Al12P12 complexes. As numerical evidence, the configurations A1 and B1 of the MeOH⋯ and EtOH⋯Al12P12 complexes possessed the most pronounced negative Eads values of −26.01 and −27.23 kcal mol−1 accompanied with τ values of 1.09 × 1010 and 8.54 × 1010 ms, respectively. Consequently, the Al12P12 nanocage was considered an appropriate sensor for MeOH and EtOH molecules.

3. Computational methods

The adsorption amplitude of VOCs (i.e., MeOH and EtOH) over the Al12P12 nanocage was fully investigated using a plethora of DFT computations with the aid of the Gaussian 09 package.43 For the investigated systems, the geometrical optimization accompanied by frequency computations was carried out at the M06-2X44 method simultaneously with a 6-31+G* basis set.

To illustrate the nucleophilic and electrophilic characters of the MeOH, EtOH, and Al12P12, the electrostatic potential (ESP) analysis was conducted. Using an electron density envelope of 0.002 au,45 surface electrostatic potential extrema (Vs,min/Vs,max) and molecular electrostatic potential (MEP) maps were evaluated and extracted to provide numerical and graphical explanations for the investigated systems, respectively. The Vs,min/Vs,max were obtained by adopting the Multiwfn 3.7 software.46

The efficacy of Al12P12 nanocage toward adsorbing VOCs was thoroughly determined in terms of adsorption (Eads) and interaction (Eint) energies. For the VOC⋯Al12P12 complexes, Eads and Eint were computed utilizing the counterpoise corrected (CC) method to eliminate the basis set superposition error (BSSE),47 relying on eqn (1) and (2), respectively.

 
Eads = EVOC⋯Al12P12 − (EVOC + EAl12P12) + EBSSE (1)
 
Eint = EVOC⋯Al12P12 − (EVOC in complex + EAl12P12 in complex) + EBSSE (2)
where EVOC⋯Al12P12, EVOC, and EAl12P12 represent the energies of investigated complexes, isolated VOCs, and Al12P12 nanocage, respectively. Whereas the EVOC in complex and EAl12P12 in complex identify the energies of the MeOH/EtOH molecules and Al12P12 nanocage based on their coordinates in the complex form.

Moreover, SAPT analysis was performed employing the SAPT0 level of truncation using the PSI4 code.48 In the context of SAPT, the total energy (ESAPT0) was divided into Eind, Eelst, Eexch, and Edisp. ESAPT0 was evaluated utilizing eqn (3).49–51

 
ESAPT0 = Eelst + Eexch + Eind + Edisp (3)

Wavefunction analyses, including NCI index and QTAIM, were executed for the VOC⋯Al12P12 complexes using the Multiwfn 3.7 software46 and visualized by the Visual Molecular Dynamics program.52 With the inclusion of QTAIM, the BPs and BCPs between the interacted species were extracted. The topological parameters were evaluated for all the studied complexes. Considering the NCI index, the 3D colored isosurfaces were extracted depending on the sign(λ2)ρ varying between blue (−0.035 au) to red (0.020 au).

Toward obtaining an adequate illustration of the electronic properties before and following the adsorption process, the FMO theory was implemented. In this regard, the HOMO/LUMO distribution patterns were plotted for the monomeric and complex forms. Similarly, HOMO/LUMO energies (EHOMO/ELUMO) were determined. Upon the obtained EHOMO and ELUMO values, the Egap and EFL values were determined as follows:

 
image file: d4ra01828a-t1.tif(4)
 
Egap = ELUMOEHOMO (5)

Based on the data obtained from FMO, the IP and EA were predestined based on eqn (6) and (7).

 
IP ≈ −EHOMO (6)
 
EA ≈ −ELUMO (7)

By applying Koopman's theorem,53 the chemical reactivity descriptors of molecules could be predicted based on quantum mechanical descriptors. Accordingly, η, ω, S, and μ were calculated utilizing eqn (8)–(11).

 
image file: d4ra01828a-t2.tif(8)
 
image file: d4ra01828a-t3.tif(9)
 
image file: d4ra01828a-t4.tif(10)
 
image file: d4ra01828a-t5.tif(11)

Afterwards, the Φ was calculated to determine the sensing ability of the studied nanocages using eqn (12),54 where Vel(+∞) identifies the electrostatic potential far from the nanocage surface that was postulated to be ≈0.

 
Φ = Vel(+∞)EFL (12)

To elucidate the influence of the adsorption process on the electronic properties of the utilized nanocage, density of states (DOS) plots were extracted within an energy range of −20 to +5 eV before and following the adsorption process based on eqn (13) employing the GaussSum software.55

 
image file: d4ra01828a-t6.tif(13)
where ε and δ represent the eigenvalue set of single-particle Hamilton and Dirac delta function, respectively.

To assess the thermodynamic parameters of the inspected complexes, ΔH, ΔG, and ΔS were evaluated based on frequency calculations as follows:

 
ΔM = MVOC⋯Al12P12 − (MVOC + MAl12P12) + EBSSE (14)
 
ΔS = −(ΔG − ΔH)/T (15)
whereas M indicates the quantity of G and H. The M of investigated complexes, VOCs, and nanocage were represented by MVOC⋯Al12P12, MVOC, and MAl12P12, respectively. T refers to temperature with a value of 298.15 K. Upon frequency computations, plots of IR and Raman spectra were extracted with the aid of GaussSum software.55 Recovery time (τ) was subsequently calculated to evaluate the feasibility of the desorption process within the complexes understudy using formula (16),56,57 where v0 and K represent the attempt frequency with a value of 1012 s−1 and Boltzmann's constant, respectively.
 
τ = v−10[thin space (1/6-em)]exp(−ΔEads/KT) (16)

4. Conclusions

The sensitivity of Al12P12 nanocage toward sensing MeOH and EtOH molecules was investigated in all plausible configurations utilizing numerous DFT calculations. The ESP outcomes unveiled the existence of evident nucleophilic and electrophilic regions over the surface of the MeOH and EtOH molecules, in particular around the O and C/H atoms, respectively. In comparison, the Al12P12 nanocage was observed with nucleophilic and electrophilic regions surrounding P and Al atoms, respectively. According to the energetic findings, the adsorption process showed higher preferability in the case of the EtOH⋯Al12P12 complexes compared to the MeOH⋯Al12P12 complexes with Eint/Eads values up to −32.84/−27.23 and −30.76/−26.01 kcal mol−1, respectively. SAPT affirmations revealed the Eelst forces with immense contributions to the attractive forces within the most preferable configurations of the VOC⋯Al12P12 complexes. QTAIM and NCI index results assured the noncovalent nature of the interaction within the studied complexes. The noticeable changes in molecular orbitals distribution patterns of MeOH/EtOH/Al12P12 nanocage, the electronic parameters, and the global reactivity descriptors highlighted the occurrence of the adsorption of VOCs on Al12P12 nanocage. Remarkably, thermodynamic parameters substantiated the exothermic character of the VOC⋯Al12P12 complexes within all plausible configurations. Thermodynamic parameters were denoted with negative values, demonstrating the spontaneous exothermic nature of the most investigated complexes. The appearance of new peaks in DOS plots confirmed the occurrence of the adsorption process between the studied VOCs and Al12P12 nanocage. Based on IR and Raman spectra findings, the occurrence of the adsorption process was ensured by the appearance of new bands in IR and Raman spectra. Recovery time results addressed the Al12P12 nanocage as an appropriate sensor for MeOH and EtOH molecules with τ values ranging from 6.11 × 10−8 to 8.54 × 1010 ms. The emerging findings would provide a comprehensive insight into the efficiency of the Al12P12 nanocage in detecting VOCs, especially for MeOH and EtOH molecules.

Author contributions

Conceptualization, Mahmoud A. A. Ibrahim and Tamer Shoeib; methodology, Mahmoud A. A. Ibrahim, Nayra A. M. Moussa, and Ahmed M. Awad; software, Mahmoud A. A. Ibrahim; formal analysis, Manar H. A. Hamad; investigation, Manar H. A. Hamad and Nayra A. M. Moussa; resources, Mahmoud A. A. Ibrahim, Shaban R. M. Sayed, Omar H. Abd-Elkader and Tamer Shoeib; data curation, Manar H. A. Hamad; writing—original draft preparation, Manar H. A. Hamad; writing—review and editing, Mahmoud A. A. Ibrahim, Nayra A. M. Moussa, Shaban R. M. Sayed, Omar H. Abd-Elkader, Muhammad Naeem Ahmed, Ahmed M. Awad, and Tamer Shoeib; visualization, Manar H. A. Hamad and Muhammad Naeem Ahmed; supervision, Mahmoud A. A. Ibrahim; project administration, Mahmoud A. A. Ibrahim, Nayra A. M. Moussa and Tamer Shoeib. All authors have read and agreed to the published version of the manuscript.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors are grateful to the Researchers Supporting Project number (RSP2024R468), King Saud University, Riyadh, Saudi Arabia. The computational work was completed with resources provided by the CompChem Lab (Minia University, Egypt, http://hpc.compchem.net/), Center for High-Performance Computing (Cape Town, South Africa, http://www.chpc.ac.za), and Bibliotheca Alexandrina (http://hpc.bibalex.org).

References

  1. S. Haghgoo and A. R. Nekoei, RSC Adv., 2021, 11, 17377–17390 RSC .
  2. M. Yoosefian, M. Zahedi, A. Mola and S. Naserian, Appl. Surf. Sci., 2015, 349, 864–869 CrossRef CAS .
  3. A. Kose, N. Yuksel and M. F. Fellah, Diamond Relat. Mater., 2022, 124, 108921 CrossRef CAS .
  4. S. Sajjad, M. A. Hashmi, T. Mahmood and K. Ayub, J. Mol. Graphics Modell., 2020, 101, 107748 CrossRef CAS PubMed .
  5. A. L. P. Silva, A. C. A. Silva, C. N. Navis and J. de Jesus Gomes Varela Júnior, J. Nanopart. Res., 2021, 23, 108 CrossRef CAS .
  6. A. Allangawi, M. A. Gilani, K. Ayub and T. Mahmood, Int. J. Hydrogen Energy, 2023, 48, 16663–16677 CrossRef CAS .
  7. J. Beheshtian, Z. Bagheri, M. Kamfiroozi and A. Ahmadi, J. Mol. Model., 2012, 18, 2653–2658 CrossRef CAS .
  8. F. Ullah, N. Kosar, A. Ali, Maria, T. Mahmood and K. Ayub, Phys. E, 2020, 118, 113906 CrossRef CAS .
  9. F. Ullah, S. Irshad, S. Khan, M. A. Hashmi, R. Ludwig, T. Mahmood and K. Ayub, J. Phys. Chem. Solids, 2021, 151, 109914 CrossRef CAS .
  10. F. Ullah, N. Kosar, K. Ayub, M. A. Gilani and T. Mahmood, New J. Chem., 2019, 43, 5727–5736 RSC .
  11. R. Padash, M. R. Esfahani and A. S. Rad, J. Biomol. Struct. Dyn., 2021, 39, 5427–5437 CrossRef CAS PubMed .
  12. A. S. M. Rady, N. A. M. Moussa, L. A. Mohamed, P. A. Sidhom, S. R. M. Sayed, M. K. Abd El-Rahman, E. Dabbish, T. Shoeib and M. A. A. Ibrahim, Heliyon, 2023, 9, e18690 CrossRef CAS PubMed .
  13. F. Younas, M. Y. Mehboob, K. Ayub, R. Hussain, A. Umar, M. U. Khan, Z. Irshad and M. Adnan, J. Comput. Biophys. Chem., 2020, 20, 85–97 CrossRef .
  14. S. Hussain, S. A. Shahid Chatha, A. I. Hussain, R. Hussain, M. Y. Mehboob, T. Gulzar, A. Mansha, N. Shahzad and K. Ayub, ACS Omega, 2020, 5, 15547–15556 CrossRef CAS PubMed .
  15. H. Farrokhpour, H. Jouypazadeh and S. Vakili Sohroforouzani, Mol. Phys., 2020, 118, 1626506 CrossRef .
  16. R. Padash, M. Rahimi-Nasrabadi, A. Shokuhi Rad, A. Sobhani-Nasab, T. Jesionowski and H. Ehrlich, J. Cluster Sci., 2018, 30, 203–218 CrossRef .
  17. N. Thongsai, N. Tanawannapong, J. Praneerad, S. Kladsomboon, P. Jaiyong and P. Paoprasert, Colloids Surf., A, 2019, 560, 278–287 CrossRef CAS .
  18. S. M. Aghaei, A. Aasi, S. Farhangdoust and B. Panchapakesan, Appl. Surf. Sci., 2021, 536, 147756 CrossRef CAS .
  19. Y. J. Huang, D. R. Ma, J. Cao, Z. Y. Tang, L. L. Hu, Y. X. Zhang, H. N. Zhao, D. H. Xia, C. He and P. K. Wong, Appl. Catal., 2023, 652, 119031 CrossRef CAS .
  20. R. Si, Y. Li, J. Tian, C. Tan, S. Chen, M. Lei, F. Xie, X. Guo and S. Zhang, Nanomater., 2023, 13, 908 CrossRef CAS PubMed .
  21. D. Li, X. Li, J. Wang, T. Wang and Y. Wen, Theor. Chem. Acc., 2023, 142, 112 Search PubMed .
  22. S. Kaviani, I. I. Piyanzina, O. V. Nedopekin and D. A. Tayurskii, Mater. Today Commun., 2022, 33, 104851 Search PubMed .
  23. C. Wen, T. Y. Liu, D. P. Wang, Y. Q. Wang, H. P. Chen, G. Q. Luo, Z. J. Zhou, C. K. Li and M. H. Xu, Prog. Energy Combust. Sci., 2023, 99, 101098 CrossRef .
  24. R. Koppmann, in Hydrocarbons, Oils and Lipids: Diversity, Origin, Chemistry and Fate, ed. H. Wilkes, Springer International Publishing, Cham, 2020, ch. 24, pp. 811–822,  DOI:10.1007/978-3-319-90569-3_24 .
  25. C. Pang, R. Han, Y. Su, Y. Zheng, M. Peng and Q. Liu, Chem. Eng. J., 2023, 454, 140125 CrossRef CAS .
  26. X. Zhang, B. Gao, A. E. Creamer, C. Cao and Y. Li, J. Hazard. Mater., 2017, 338, 102–123 CrossRef CAS PubMed .
  27. F. I. Khan and A. K. Ghoshal, J. Loss Prev. Process Ind., 2000, 13, 527–545 CrossRef .
  28. M. Weng, L. Zhu, K. Yang and S. Chen, Environ. Monit. Assess., 2010, 163, 573–581 CrossRef CAS PubMed .
  29. J. Li, S. Deng, G. Li, Z. Lu, H. Song, J. Gao, Z. Sun and K. Xu, Environ. Res., 2022, 203, 111821 CrossRef CAS PubMed .
  30. Z. Zhong, Q. Sha, J. Zheng, Z. Yuan, Z. Gao, J. Ou, Z. Zheng, C. Li and Z. Huang, Sci. Total Environ., 2017, 583, 19–28 CrossRef CAS PubMed .
  31. Y. Wang, C. Wang, K. Zeng, S. Wang, H. Zhang, X. Li, Z. Wang and C. Zhang, Appl. Surf. Sci., 2022, 576, 151797 CrossRef CAS .
  32. Y. Huang, S. S. Ho, Y. Lu, R. Niu, L. Xu, J. Cao and S. Lee, Molecules, 2016, 21, 56 CrossRef PubMed .
  33. J. Wang, Y. Zhang, Z. Wu, S. Luo, W. Song and X. Wang, J. Environ. Sci., 2022, 114, 322–333 CrossRef CAS PubMed .
  34. L. Zhu, D. Shen and K. H. Luo, J. Hazard. Mater., 2020, 389, 122102 CrossRef CAS .
  35. Q. Zhao, Z. Zhao, R. Rao, Y. Yang, S. Ling, F. Bi, X. Shi, J. Xu, G. Lu and X. Zhang, J. Colloid Interface Sci., 2022, 627, 385–397 CrossRef CAS .
  36. B. Guieysse, C. Hort, V. Platel, R. Munoz, M. Ondarts and S. Revah, Biotechnol. Adv., 2008, 26, 398–410 CrossRef CAS PubMed .
  37. C. Zhang, X. Gao, J. Qin, Q. Guo, H. Zhou and W. Jin, J. Hazard. Mater., 2021, 402, 123817 CrossRef CAS PubMed .
  38. R. A. Alsaigh, S. Rahman, F. S. Alfaifi, M. A. Al-Gawati, R. Shallaa, F. Alzaid, A. F. Alanazi, H. Albrithen, K. E. Alzahrani, A. K. Assaifan, A. N. Alodhayb and P. E. Georghiou, Chemosensors, 2022, 10, 452 CrossRef CAS .
  39. V. Nagarajan and R. Chandiramouli, J. Mol. Graphics Modell., 2018, 81, 97–105 CrossRef CAS PubMed .
  40. A. M. Tolmachev, D. A. Firsov, T. A. Kuznetsova and K. M. Anuchin, Prot. Met. Phys. Chem. Surf., 2009, 45, 163–168 CrossRef CAS .
  41. P. K. Adak, S. K. Singh, J. Singh, S. Mahesh, M. K. Jain, P. A. K. Sagar, S. Banerjee and M. A. S. Khan, J. Mol. Model., 2022, 28, 400 CrossRef CAS PubMed .
  42. F. Khaliq, K. Ayub, T. Mahmood, S. Muhammad, S. Tabassum and M. A. Gilani, Mater. Sci. Semicond. Process., 2021, 135, 106122 CrossRef CAS .
  43. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09 Revision E01, Gaussian Inc., Wallingford CT, USA Search PubMed .
  44. Y. Zhao and D. G. Truhlar, J. Chem. Theory Comput., 2008, 4, 1849–1868 CrossRef CAS PubMed .
  45. M. A. A. Ibrahim, J. Mol. Model., 2012, 18, 4625–4638 CrossRef CAS PubMed .
  46. T. Lu and F. Chen, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed .
  47. S. F. Boys and F. Bernardi, Mol. Phys., 1970, 19, 553–566 CrossRef CAS .
  48. R. M. Parrish, L. A. Burns, D. G. A. Smith, A. C. Simmonett, A. E. DePrince 3rd, E. G. Hohenstein, U. Bozkaya, A. Y. Sokolov, R. Di Remigio, R. M. Richard, J. F. Gonthier, A. M. James, H. R. McAlexander, A. Kumar, M. Saitow, X. Wang, B. P. Pritchard, P. Verma, H. F. Schaefer 3rd, K. Patkowski, R. A. King, E. F. Valeev, F. A. Evangelista, J. M. Turney, T. D. Crawford and C. D. Sherrill, J. Chem. Theory Comput., 2017, 13, 3185–3197 CrossRef CAS PubMed .
  49. T. M. Parker, L. A. Burns, R. M. Parrish, A. G. Ryno and C. D. Sherrill, J. Chem. Phys., 2014, 140, 094106 CrossRef PubMed .
  50. E. G. Hohenstein and C. D. Sherrill, J. Chem. Phys., 2010, 132, 184111–184120 CrossRef .
  51. E. G. Hohenstein, R. M. Parrish, C. D. Sherrill, J. M. Turney and H. F. Schaefer 3rd, J. Chem. Phys., 2011, 135, 174107–174119 CrossRef PubMed .
  52. W. Humphrey, A. Dalke and K. Schulten, J. Mol. Graphics, 1996, 14, 33–38 CrossRef CAS PubMed .
  53. M. Ernzerhof, J. Chem. Theory Comput., 2009, 5, 793–797 CrossRef CAS PubMed .
  54. M. R. Hossain, M. M. Hasan, M. Nishat, N.-E. Ashrafi, F. Ahmed, T. Ferdous and M. Abul Hossain, J. Mol. Liq., 2021, 323, 114627 CrossRef CAS .
  55. N. M. O'Boyle, A. L. Tenderholt and K. M. Langner, J. Comput. Chem., 2008, 29, 839–845 CrossRef PubMed .
  56. H. M. Badran, K. M. Eid and H. Y. Ammar, Results Phys., 2021, 23, 103964 CrossRef .
  57. S. Demir and M. F. Fellah, Appl. Surf. Sci., 2020, 504, 144141 CrossRef CAS .

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra01828a

This journal is © The Royal Society of Chemistry 2024