Next Article in Journal
A Comprehensive Study to Determine the Residual Elimination Pattern of Major Metabolites of Amoxicillin–Sulbactam Hybrid Molecules in Rats by UPLC–MS/MS
Previous Article in Journal
Fluorescence Sensors for the Detection of L-Histidine Based on Silver Nanoclusters Modulated by Copper Ions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DFT-D3 and TD-DFT Studies of the Adsorption and Sensing Behavior of Mn-Phthalocyanine toward NH3, PH3, and AsH3 Molecules

by
Heba Mohamed Badran
1,
Khaled Mahmoud Eid
2,
Hatim Omar Al-Nadary
1,* and
Hussein Youssef Ammar
1,3,*
1
Physics Department, College of Science & Arts, Najran University, Najran 11001, Saudi Arabia
2
Physics Department, Faculty of Education, Ain Shams University, Roxy, Cairo 11566, Egypt
3
STEM Pioneers Training Lab, Najran University, Najran 11001, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(10), 2168; https://doi.org/10.3390/molecules29102168
Submission received: 15 April 2024 / Revised: 1 May 2024 / Accepted: 4 May 2024 / Published: 7 May 2024
(This article belongs to the Section Computational and Theoretical Chemistry)

Abstract

:
This study employs density functional theory (DFT) calculations at the B3LYP/6-311+g(d,p) level to investigate the interaction of XH3 gases (X = N, P, As) with the Mn-phthalocyanine molecule (MnPc). Grimme’s D3 dispersion correction is applied to consider long-range interactions. The adsorption behavior is explored under the influence of an external static electric field (EF) ranging from −0.514 to 0.514 V/Å. Chemical adsorption of XH3 molecules onto the MnPc molecule is confirmed. The adsorption results in a significant decrease in the energy gap (Eg) of MnPc, indicating the potential alteration of its optical properties. Quantum theory of atoms in molecules (QTAIM) analysis reveals partially covalent bonds between XH3 and MnPc, and the charge density differenc (Δρ) calculations suggest a charge donation-back donation mechanism. The UV-vis spectrum of MnPc experiences a blue shift upon XH3 adsorption, highlighting MnPc’s potential as a naked-eye sensor for XH3 molecules. Thermodynamic calculations indicate exothermic interactions, with NH3/MnPc being the most stable complex. The stability of NH3/MnPc decreases with increasing temperature. The direction and magnitude of the applied electric field (EF) play a crucial role in determining the adsorption energy (Eads) for XH3/MnPc complexes. The Eg values decrease with an increasing negative EF, which suggests that the electrical conductivity (σ) and the electrical sensitivity (ΔEg) of the XH3/MnPc complexes are influenced by the magnitude and direction of the applied EF. Overall, this study provides valuable insights into the suggested promising prospects for the utilization of MnPc in sensing applications of XH3 gases.

1. Introduction

Preserving the environment stands as one of the utmost priorities for scientists today. With technological advancements permeating every facet of life, the prevalence of pollutants has escalated to a level that poses a threat to living organisms. Consequently, there arose a necessity to develop and produce nanosensors specifically designed for detecting environmentally harmful gases. Among the hydrides of the fifth group, ammonia (NH3), phosphine (PH3), and arsine (AsH3) merit special attention due to the hazards they pose. NH3 is a colorless gas with a pungent odor. NH3 is used in several industries, such as the production of fertilizers, pesticides, hair dyes, plastics, and the textile industry. The inhalation of ammonia can result in severe irritation to the respiratory system [1,2,3]. PH3 is a scentless toxic gas characterized by its spontaneous flammability in the air or when interacting with oxygen. PH3 has several uses; for instance, it is used to exterminate insects and rodents. It is also used to preserve agricultural crops in warehouses during storage. Its risks are due to inhalation or absorption through the skin. Inhaling PH3 gas causes pneumonia and respiratory poisoning and can harm the central nervous system [4,5]. AsH3 is a colorless gas that has a disagreeable fish-like scent. It is used in the steel industry and metal treatment. It is an extremely toxic gas. The risks caused by arsine are abdominal pain, diarrhea, dyspnea, hemolysis, and renal failure [6,7,8].
Earlier research indicates that attempts were made toward the adsorption of these hazardous gases onto the surfaces of various materials. Luo et al. [9] demonstrated that the adsorption of NH3, PH3, and AsH3 on graphene is weak, while the doping of graphene with La, Ce, Nd, Pm, Sm, Eu, and Gd improved the adsorption of NH3 and weakened the adsorption of AsH3. Habibi-Yangjeh et al. [4] suggested that Ta/P heptazine graphitic carbon nitride serves as a suitable sensor for PH3. Conversely, Ranea et al. [10] found that the V atom of V2O5 is an attractive center for NH3, PH3, and AsH3. Additionally, CdSe functions efficiently as a gas sensor for NH3, PH3, and AsH3 gases [11].
Phthalocyanines (Pcs) represent a class of aromatic macrocyclic organic compounds characterized by the molecular formula C32H18N8. Their distinct attributes encompass remarkable thermal and chemical stability, coupled with favorable electrical and optical properties, primarily attributed to their extensive electronic π-conjugated system [12,13].
The enumerated advantages have significantly propelled the utilization of phthalocyanines (Pcs) across diverse domains, including solar cells, semiconductors, catalysis, sensors, and tumor treatment [12,14,15,16]. Notably, research indicates that the incorporation of transition elements into phthalocyanine structures enhances their suitability for a wide range of applications. For instance, the integration of MnPc and FePc into junctions between two single-walled carbon nanotubes results in superior magnetic spin moments compared to CoPc and NiPc, thereby enhancing their performance in spintronic devices [15]. Furthermore, investigations into the magnetic properties of transition metal phthalocyanine sheets (TM = Cr-Zn) reveal that only MnPc exhibits ferromagnetic behavior [17]. Doping Pcs with Co and Mn has been shown to heighten their redox activity, a crucial aspect for electrochemical applications [16]. Given these findings, it can be reasonably anticipated that grafting phthalocyanine with a transition element, particularly Mn, will induce notable changes in its properties as a gas sensor.
To the best of our knowledge, the application of MnPc as an adsorbent or sensor for NH3, PH3, and AsH3 gases has not been explored previously. This study is designed to investigate the adsorption properties of NH3, PH3, and AsH3 gases on the MnPc molecule. Additionally, we will explore the impact of an external static electric field on these adsorption characteristics.

2. Results and Discussions

In this study, we examined the adsorption of XH3 gases (where X = N, P, As) on a MnPc molecule. Initially, we focused on determining the most energetically stable structures of both the individual XH3 and MnPc molecules. This investigation aimed to understand the influence of the Mn atom on the properties of Pc. Subsequently, we investigated the most energetically stable structures of the XH3/Pc complexes.

2.1. Properties of XH3 Adsorbates and the MnPc Adsorbent

Figure 1 displays the geometrical structure, density of states (DOS), highest occupied molecular orbital (HOMO), lowest unoccupied molecular orbital (LUMO), and the molecular electrostatic potential (MESP) for XH3 molecules. The X–H bond lengths and H–X–H angles were determined as follows: NH3 (1.01 Å, 108.3°), PH3 (1.42 Å, 93.61°), and AsH3 (1.52 Å, 92.36°), which are consistent with previous findings [2,3,18,19]. The MESP analysis reveals a negative electrostatic potential around the X atom, indicating the presence of a 2p electron lone pair. This suggests that the X atom functions as a nucleophilic center. Moreover, the calculated electrical dipole moments (D) were found to be 1.673, 0.823, and 0.345 Debye for NH3, PH3, and AsH3, respectively, which aligns with the findings of Zhang et al. [18]. Consequently, the chemical reactivity of a molecule increases with its dipole moment [14]. Therefore, one can anticipate the following trend in chemical reactivity: NH3 > PH3 > AsH3.
Figure 2 illustrates the optimized structures and MESP for Pc and MnPc, while their electronic properties are summarized in Table 1. It was observed that in MnPc, the Mn–N bond length was determined to be 1.958 Å and the N–Mn–N angle was found to be 90°, which is consistent with previous theoretical and experimental studies [15,20,21]. The calculated band gap (Eg) values indicate that both Pc and MnPc are semiconductors, with the presence of the Mn atom leading to a 34.5% reduction in the Eg value of the Pc molecule. Furthermore, the binding energy (Eb) for MnPc was more negative compared to that of Pc, implying that the Mn atom enhances the stability of the MnPc molecule. This can be attributed to intramolecular partial charge transfer (PCT) occurring from the Mn atom to the rest of the molecule, resulting in a positive charge of 0.954 e accumulating on the Mn atom. Moreover, the MnPc molecule exhibits lower values of Eg, IP, and η, while displaying higher values of Ef and ω compared to the Pc molecule. These characteristics indicate that the MnPc molecule is more reactive than the Pc molecule. Additionally, the positive charge on the Mn atom results in a positive electrostatic potential surrounding it, as depicted in Figure 2d. Consequently, the Mn atom can be considered an electrophilic site. It is worth noting that the Pc and MnPc molecules do not exhibit noticeable values of the electrical dipole moment (D) due to the symmetrical distribution of electrical charges on the atoms within these molecules. Figure 3b,e depict the DOS and PDOS for Pc and MnPc molecules, respectively. It is evident that the presence of the Mn atom causes the HOMO to shift upwards by 0.507 eV, while the LUMO is shifted downwards by 0.225 eV. Consequently, the Eg value of the Pc molecule decreases from 2.120 eV to 1.388 eV in the MnPc molecule, which is consistent with previous research [21,22].
Furthermore, the UV-vis spectra of Pc and MnPc molecules are displayed in Figure 3g. The Pc molecule exhibits two absorption peaks, known as the Soret band and Q band, at wavelengths of 340 nm and 608 nm, respectively. These findings align with previous studies [22,23,24]. On the other hand, the MnPc molecule demonstrates a broad absorption peak at 604 nm.

2.2. Adsorption of XH3 on MnPc

The adsorption of XH3 molecules on the Mn site of the MnPc molecule was investigated in three different adsorption modes, as shown in Figure 4. All modes undergo full optimization, and it was found that modes 1 and 2 resulted in energetically stable complexes. Interestingly, mode 3 reoriented to yield the same complexes as mode 2. Top and side views of the optimized complexes for adsorption modes 1 and 2 are presented in Figure 5. The adsorption properties for modes 1 and 2 are summarized in Table 2.
In accordance with previous studies [25,26,27], it is well-established that chemisorption is characterized by high adsorption energy ( E a d s  ≥ 0.2 eV). Therefore, based on the obtained XH3/MnPc complexes in adsorption modes 1 and 2, it can be concluded that the XH3 molecule undergoes chemical adsorption on the MnPc molecule. It is important to note that a more negative Eads value indicates a higher degree of adsorption. For adsorption mode 1, the trend of adsorption strength was observed as NH3/MnPc > PH3/MnPc > AsH3/MnPc complexes, indicating that NH3 exhibited the highest adsorption strength, followed by PH3 and AsH3. On the other hand, for adsorption mode 2, the trend was NH3/MnPc < PH3/MnPc < AsH3/MnPc, suggesting that AsH3 exhibited the highest adsorption strength, followed by PH3 and NH3. Furthermore, it can be observed that the adsorption strength in mode 1 was higher compared to mode 2. The distances between the X atom and the Mn adsorbing site (dX-Mn) followed the trend of NH3/MnPc < PH3/MnPc < AsH3/MnPc for both adsorption modes. This indicates that NH3 exhibited the closest proximity to the Mn adsorbing site, followed by PH3 and AsH3, in both modes.
Furthermore, the adsorption of XH3 molecules resulted in a decrease in the Eg value of the MnPc molecule. Specifically, for adsorption mode 1, the Eg value decreased to 86.2%, 91.6%, and 92.8% for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively. For adsorption mode 2, the Eg value decreased to 96.2%, 98.2%, and 97.9% for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively. In other words, adsorption mode 1 resulted in a greater reduction in the Eg value compared to adsorption mode 2. Notably, the largest decrease in the Eg value was observed for the NH3/MnPc complex in adsorption mode 1. Additionally, the adsorption of XH3 molecules led to an increase in the dipole moment. Specifically, for adsorption mode 1, the dipole moment increased to 3.130, 2.083, and 1.576 Debye for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively. For adsorption mode 2, the dipole moment increased to 0.949, 0.228, and 0.213 Debye for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively.
The results obtained can be further analyzed using several methods, including QTAIM, electrostatic potential (ESP), NBO atomic charge, charge density difference (∆ρ), and PDOS. The QTAIM analysis is particularly useful for understanding the nature of interactions [28,29,30]. Previous studies have established certain characteristics for different types of interactions, such as van der Waals interactions, weak hydrogen bonds, and ionic bonds. These characteristics include −G(r)/V(r) > 1, H(r) > 0, and ∇2ρ > 0. Strong interactions are classified by ∇2ρ > 10–1 au, while weak interactions have ∇2ρ < 10–1 au. Partially covalent bonds are characterized by ∇2ρ > 0 and 0.5 < –G(r)/V(r) < 1. Therefore, the QTAIM theory can be employed to analyze the topological parameters of the bond critical points (BCP) of type (3, –1) formed between the XH3 and MnPc molecules. The BCPs, as depicted in Figure 6, and their corresponding parameters are summarized in Table 3. The analysis reveals that in adsorption mode 1, only one BCP exists between the X atom of the XH3 molecule and the Mn atom of the MnPc molecule. This BCP exhibits a ∇2ρ value greater than zero, specifically 0.171, 0.064, and 0.046 au for NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes, respectively. Furthermore, the −G(r)/V(r) ratio for these BCPs is greater than 0.5 but less than 1, specifically 0.952, 0.813, and 0.833 for NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes, respectively. As a result, these BCPs are classified as partially covalent bonds. Moreover, the ∇2ρ values indicate that the interaction in the NH3/MnPc complex is characterized as strong, while the interactions in the PH3/MnPc and AsH3/MnPc complexes are classified as weak.
In adsorption mode 2, the NH3/MnPc complex exhibits one BCP between the N atom of the NH3 molecule and the Mn atom of the MnPc molecule, with a ∇2ρ value of 0.030 au and a −G(r)/V(r) ratio of 0.897. This BCP is classified as a weak, partially covalent bond. On the other hand, for the PH3/MnPc and AsH3/MnPc complexes, two BCPs are observed between the XH3 molecule and the MnPc molecule. The first BCP (X–Mn) is formed between the X atom of the XH3 molecule and the Mn atom of the MnPc molecule, while the second BCP (H–N) is formed between an H atom of the XH3 molecule and an N atom of the MnPc molecule. These BCPs have positive ∇2ρ values of less than 10–1 and −G(r)/V(r) ratios greater than 1. Consequently, the first BCPs (X–Mn) are classified as van der Waals interactions, while the second BCPs (H–N) are classified as weak hydrogen bonds.
Figure 7 illustrates the ESP for the MnPc and XH3 molecules. The ESP was calculated along the Z-axis, with the X and Mn atoms positioned at the origin point. The ESP for the MnPc molecule appears positive and symmetric around the Mn atom along the Z-axis. In contrast, the ESP for the XH3 molecule exhibits an asymmetric pattern along the Z-axis.
In the −Z direction, the ESP curves for NH3, PH3, and AsH3 display minimum negative values of −0.110, −0.037, and −0.021 au, respectively, occurring at distances of –1.303, −1.949, and −2.152 Å from the origin. On the other hand, in the +Z direction, the ESP curves have minimum values of 0.003, 0.002, and −0.003 au at distances of 1.887 Å, 3.373 Å, and 2.828 Å, respectively. Notably, the minima of the curves in the –Z direction are lower than those in the +Z direction. This discrepancy arises from the presence of the X atom’s electron lone pair in the –Z direction. These observations may help explain the stronger interaction observed for adsorption in mode 1 compared to mode 2, as well as the trend observed in the adsorption behavior. To analyze partial charge transfer (PCT), NBO atomic charges were computed. In all the examined adsorption structures, the XH3 molecule undergoes a positive charge, suggesting a transfer of charge from the XH3 molecule to the MnPc molecule. This PCT is more significant in adsorption mode 1 than in adsorption mode 2. Consequently, it is reasonable to anticipate that the PCT contributes to the reinforcement of adsorption mode 1. In addition, for adsorption mode 1, the XH3 molecule experiences a loss of charges ( Q X H 3 ) of 0.177, 0.324, and 0.226 e, while the Mn adsorbing site gains charges of 0.127, 0.125, and 0.187 e for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively. This implies that the XH3 molecule not only transfers charge to the adsorbing site but also to the other atoms of the MnPc molecule. Specifically, the XH3 molecule loses more charge than what is gained by the Mn adsorbing site.
It is important to note that the trend of  Q X H 3  does not necessarily align with the trend of Eads, indicating the presence of another mechanism influencing the adsorption process. To further elucidate this, the charge density differences (Δρ) for the XH3/MnPc complexes were analyzed and depicted in Figure 8. The Δρ values exhibit both positive (blue) and negative (red) regions for both the XH3 adsorbate and MnPc substrate. This suggests that charge transfer occurs in two directions, from the XH3 molecule to the MnPc molecule and vice versa. Consequently, a charge donation-back donation mechanism is proposed for the adsorption process. Furthermore, Figure 8 illustrates that the Δρ values for adsorption mode 1 are higher than those for adsorption mode 2, which corresponds well with the adsorption energies described in Table 2. Lastly, as a result of the adsorption, a redistribution of charges between the XH3 and MnPc molecules occurs, leading to an increase in the electric dipole moment values of the XH3/MnPc complexes.
The partial densities of states (PDOS) for the XH3/MnPc complexes are presented in Figure 9 for adsorption mode 1 and Figure 10 for adsorption mode 2. By comparing the DOS of the XH3 molecule prior to adsorption (Figure 1) with those after adsorption (Figure 9 and Figure 10), significant changes are evident, indicating the interaction between the XH3 molecule and the MnPc molecule. As a result of this interaction, the HOMO of the MnPc molecule undergoes a shift in energy. Specifically, the HOMO is shifted from −4.798 eV to −5.079, −5.136, and −5.163 eV for the NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes, respectively. This shift in energy reflects the influence of the XH3 molecule on the electron density distribution and electronic structure of the MnPc molecule. Similarly, the LUMO of the MnPc molecule experiences an energy shift following the interaction with the XH3 molecule. The LUMO is shifted from −3.410 eV to −3.230, −3.312, and −3.340 eV for the NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes, respectively. This shift in energy suggests changes in the accessibility of the LUMO for electron transfer or participation in chemical reactions.
Overall, the observed shifts in the HOMO and LUMO energies of the MnPc molecule indicate the modification of its electronic structure due to the interaction with the XH3 molecule in the XH3/MnPc complexes. These changes in the DOS further highlight the influence of the XH3 adsorbate on the electronic properties of the MnPc substrate.
Consequently, for adsorption mode 1, the shifts in the HOMO and LUMO energies lead to a narrowing of the bandgap (Eg) for the NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes by 13.80%, 8.43%, and 7.21%, respectively. In contrast, for adsorption mode 2, the shifts in the HOMO and LUMO energies are not as significant, resulting in a smaller reduction in the Eg values. Specifically, the Eg values for the NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes are slightly narrowed by 3.78%, 1.82%, and 2.08%, respectively.
Furthermore, Equation (1) represents the relationship between the electrical conductivity (σ) and the HOMO-LUMO gap (Eg). This equation suggests that a smaller Eg corresponds to a higher electrical conductivity [31,32,33,34,35,36].
σ = A T 3 / 2 e E g / 2 K T
where A is a constant, K is Boltzmann’s constant, and T is the temperature. As a result, it is reasonable to assume that the XH3 molecule’s adsorption will raise the MnPc molecule’s electrical conductivity. Furthermore, the increase in  σ  in the case of the adsorption mode 1 is greater than in the case of the adsorption mode 2, and in the case of NH3, it is greater than in the case of PH3 and AsH3. Consequently, the MnPc molecule exhibits a higher sensitivity for NH3 molecules than PH3 and AsH3 molecules. Our results for mode 1 clarify that MnPc may be useful for XH3 detection, especially NH3 detection.

2.3. UV-Vis Spectra Analysis

The influence of XH3 adsorption on the UV-vis spectrum of the MnPc molecule was investigated for adsorption modes 1 and 2. TD-DFT calculations were performed on the optimized structures of XH3/MnPc complexes to predict the UV-vis absorption spectra. In Figure 11a,b, the UV-vis spectra for XH3/MnPc complexes in adsorption modes 1 and 2 are presented, respectively.
For adsorption mode 1, the NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes exhibit maximum absorption wavelength peaks (λmax) in the visible region at 524, 552, and 572 nm, respectively. In contrast, Figure 3g illustrates that the λmax of the MnPc molecule is situated at 604 nm. This indicates that the adsorption of XH3 molecules induces a blue shift in the UV-vis spectrum of MnPc. Consequently, the color of the MnPc molecule may be altered by the adsorption of XH3 molecules, suggesting the potential utility of MnPc as a naked-eye sensor for the investigated XH3 molecules.
Conversely, for adsorption mode 2, the NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes display λmax in the visible region at 604, 600, and 600 nm, respectively. In this case, XH3 molecules exhibit no significant impact on the UV-vis spectrum of MnPc.

2.4. Thermodynamic Analysis

Given that the majority of gas sensors operate at temperatures below 800 K [37,38], thermodynamic calculations were conducted for the XH3 molecules, MnPc substrate, and XH3/MnPc complexes within the temperature range of 300 to 800 K. As adsorption mode 1 exhibited the most pronounced impact on the electrical and optical properties of the MnPc molecule, subsequent discussions will focus on this particular adsorption mode.
The thermodynamic parameters, including enthalpy difference (ΔH) and free energy difference (ΔG), play a crucial role in characterizing the strength and spontaneity of gas adsorption. In the case of XH3/MnPc complexes, ΔH was computed using Equation (10) and is illustrated as a function of temperature (T) in Figure 12a. Negative ΔH values are indicative of an exothermic reaction, and a more negative ΔH signifies greater stability of the products. Figure 12a demonstrates that for XH3/MnPc complexes, the ΔH values exhibit negativity across the entire temperature range under consideration. Furthermore, as temperature increases, the ΔH values become less negative. Notably, at a specific temperature, the ΔH value for the NH3/MnPc complex is more negative compared to the PH3/MnPc and AsH3/MnPc complexes. This observation suggests that the interaction between the XH3 molecule and the MnPc molecule is exothermic. Specifically, the NH3/MnPc complex is more stable than the PH3/MnPc and AsH3/MnPc complexes. Additionally, the stability of NH3/MnPc complexes decreases with an increase in temperature.
The ΔG values for the XH3/MnPc complexes were determined using Equation (11) and are depicted as a function of temperature in Figure 12b. Negative and positive ΔG values signify spontaneous and non-spontaneous reactions, respectively, with low negative ΔG values suggesting a potential for reversing the reaction [39,40,41]. Figure 12b shows a linear increase in ΔG values with temperature. At room temperature (300 K), all investigated complexes display negative ΔG values. However, beyond 300, 400, and 600 K for AsH3/MnPc, PH3/MnPc, and NH3/MnPc complexes, respectively, ΔG values turn positive, indicating a non-spontaneous adsorption process. Additionally, at T = 300 K, the ΔG value for the NH3/MnPc complex is more negative than that for the PH3/MnPc and AsH3/MnPc complexes. Consequently, the adsorption process can be more easily reversed for the PH3/MnPc and AsH3/MnPc complexes than for the NH3/MnPc complex.
The thermodynamic adsorption equilibrium constant (K) for the adsorption process was calculated using Equation (12), and log K was plotted against temperature, as shown in Figure 12c. K serves as a crucial parameter in assessing the strength and spontaneity of the adsorption process, with higher K values indicating stronger adsorption [42]. Furthermore, K values greater than 1 suggest spontaneous adsorption, while values less than 1 indicate non-spontaneous adsorption. Observing Figure 12c, it is evident that log K for the NH3/MnPc complex surpasses that of the PH3/MnPc and AsH3/MnPc complexes. Consequently, the adsorption of the NH3 molecule is stronger compared to the adsorption of PH3 and AsH3 molecules. Additionally, for the examined XH3/MnPc complexes, log K decreases as temperature increases, signifying that elevated temperatures diminish the ability of the MnPc molecule to adsorb XH3 molecules. Moreover, beyond 300, 400, and 600 K for AsH3/MnPc, PH3/MnPc, and NH3/MnPc complexes, respectively, the K value falls below 1 (log K < 0), highlighting non-spontaneous adsorption.

2.5. Effect of the EF on the Adsorption of XH3 on MnPc

This study explored the adsorption properties under the influence of an external static electric field (EF), applied within the range of −0.514 to 0.514 V/Å with increments of 0.125 V/Å along the axis perpendicular to the MnPc molecule’s plane, as depicted in Figure 13. At each EF step, both XH3 and MnPc molecules, as well as XH3/MnPc complexes, undergo full optimization. The dipole moment plotted against the electric field is illustrated in Figure 14a for free XH3 gases and Figure 14b for MnPc and XH3/MnPc complexes.
Figure 14 presents the z-component of the dipole moment for free XH3 molecules, the MnPc molecule, and XH3/MnPc complexes in response to the applied electric field (EF) along the Z-axis. Notably, the z-component of the dipole moment is considered for this analysis. In Figure 14a, it is evident that the dipole moment of XH3 molecules increases with the magnitude of the EF in the negative direction and decreases with increasing the EF in the positive direction. Furthermore, the dipole moment follows the trend NH3 > PH3 > AsH3. Moving to Figure 14b, the dipole moment of the MnPc molecule increases with the magnitude of the EF in the negative direction, while it decreases with increasing the EF in the positive direction. The chemical reactivity of a material with its surrounding environment tends to increase with a higher dipole moment [14]. The dipole moments of both XH3 and MnPc molecules are enhanced by negative EF values, leading to an enhancement of the adsorption energy (refer to Figure 15). Conversely, positive EF values result in a decrease in the dipole moment for both the MnPc and XH3 molecules, leading to inhibition of the adsorption energy (refer to Figure 15). Therefore, the direction and magnitude of the applied EF play a crucial role in determining the adsorption energy (Eads) for XH3/MnPc complexes.
The impact of the EF on the Eg values for the MnPc molecule and XH3/MnPc complexes is illustrated in Figure 16a. It is observed that the Eg value for the MnPc molecule remains relatively constant with varying EF values. In contrast, the Eg values for XH3/MnPc complexes exhibit changes: they increase as the positive EF rises, reaching 1.22, 1.38, and 1.38 eV for NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes, respectively, at EF = 0.514 V/Å. Conversely, the Eg values decrease with an increasing negative EF, reaching 1.18, 1.22, and 1.23 eV for NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes, respectively, at EF = −0.514 V/Å. This suggests that the electrical conductivity (σ) of the XH3/MnPc complexes is influenced by the magnitude and direction of the applied EF. Additionally, the electrical sensitivity (ΔEg) dependency on the EF was investigated using the following equation:
Δ E g = E g ( X H 3 / M n P c ) E g ( M n P c ) E g ( M n P c ) × 100
In Figure 16b, at EF = 0.0 V/Å, the ΔEg values were observed to be −13.68%, −8.43%, and −7.21% for NH3/MnPc, PH3/MnPc, and AsH3/MnPc complexes, respectively. This indicates that the adsorption of XH3 molecules leads to a decrease in the Eg value for the MnPc molecule, accompanied by an increase in σ. Furthermore, the MnPc molecule exhibited greater sensitivity to NH3 compared to PH3 and AsH3 molecules. Additionally, the influence of the electric field is evident in the ΔEg values: increasing the positive EF reduces the magnitude of ΔEg, making it less negative. Conversely, increasing the negative EF intensifies the magnitude of ΔEg, making it more negative. Therefore, a positive electric field diminishes the sensitivity, whereas a negative electric field enhances the sensitivity of the MnPc molecule to XH3 molecules.
It is noteworthy that the effect of the electric field on the UV-vis spectrum of the XH3/MnPc complexes was investigated, but no considerable effect was observed.

3. Methods

The interaction of XH3 gases onto the MnPc molecule is investigated utilizing DFT [43] calculations at the B3LYP/6-311+g(d,p) level of theory. D3, Grimme’s dispersion that considers the long-range interactions, is considered [44,45]. Yang et al. [46] state that B3LYP gave good performances for metal phthalocyanines. Additionally, B3LYP is utilized for transition metal-doped porphyrins [31,47], and the 3D transition metal complexes for hexaazabipy H2 and gave reliable results [48]. Geometrical optimizations to obtain the most energetically stable structures for free XH3 gases, MnPc bare molecule, and XH3/MnPc complexes are accomplished. The UV-vis absorption spectra for Pc and MnPc molecules, as well as the XH3/MnPc complexes, are estimated by Time-Dependent DFT (TD-DFT) calculations. To cover all the expected electronic transitions in the investigated range (0–1000 nm), a satisfactory number of excited states (n = 15) was projected. To estimate the relative stability of the Pc and MnPc molecules, the average binding energy for each atom (Eb) was estimated according to Equation (3) [39]. The more negative the Eb value, the more stable the molecule.
E b = 1 n E m o l e c u l e i = 1 n E i
where n is the total number of atoms of the molecule,  E m o l e c u l e  is the total energy of the molecule-optimized structure, and  E i  is the single atom’s total energy. The relative chemical reactivity of molecules is estimated according to the following parameters: the HOMO-LUMO energy gap (Eg), the ionization potential (IP), the chemical potential (µ), hardness (η), and electrophilicity (ω). The high reactive molecule is characterized by low values of Eg, IP, µ, and η and high value of ω [49,50,51].
The IP is evaluated via Koopman’s approximation [49,50]
I P   E H O M O
µ, η, and ω are evaluated by Equations (5)–(7), respectively [50,51].
µ 1 2   E H O M O + E L U M O
η 1 2   E L U M O E H O M O
ω µ 2 2 η
The strength of the XH3-MnPc interaction is judged in terms of the adsorption energy (Eads), which is assessed by Equation (8). The strong XH3-MnPc interaction is accompanied by more negative Eads values.
E a d s = E X H 3 / M n P c ( E M n P c + E X H 3 )
E X H 3 / M n P c E M n P c , and  E X H 3  are the total energies of the XH3/MnPc complex, MnPc bare molecule, and free XH3 gases, respectively. The charge density difference (Δρ) for the XH3/MnPc complex is calculated as follows:
Δ ρ = ρ X H 3 / M n P c ( ρ M n P c + ρ X H 3 )
ρ X H 3 / M n P c ρ M n P c , and  ρ X H 3  are the charge densities for the XH3/MnPc complex, MnPc bare molecule, and free XH3 gases, respectively.
Thermodynamic calculations are performed in the range of T = 300–800 K to evaluate the enthalpy difference (ΔH) and Gibbs free energy difference (ΔG) for the XH3/MnPc complexes by Equations (10) and (11), respectively [52].
Δ H = H X H 3 / M n P c ( H M n P c + H X H 3 )
Δ G = G X H 3 / M n P c ( G M n P c + G X H 3 )
where  H X H 3 / M n P c H M n P c , and  H X H 3  are the enthalpies while  G X H 3 / M n P c G M n P c , and  G X H 3  are the Gibbs free energies for the XH3/MnPc complex, MnPc bare molecule, and free XH3 gases, respectively. The thermodynamic adsorption equilibrium constant (Kads) is a crucial adsorption factor. A Kads value higher than 100 is essential for the sorbent to be a useful application. K is evaluated by Equation (12) [18].
K a d s = e Δ G R T
where R is 8.314 J.mol−1.K−1.
The adsorption behavior under the influence of an external static electric field (EF) was investigated. The EF was applied from −0.514 to 0.514 V/Å with a step of 0.125 V/Å (0.0025 au) along the axis, which was perpendicular to the plane of the MnPc molecule. Gaussian 09 software [53] was used for all calculations. GaussSum 3.0 was used to show the density of states (DOS) for the structures under investigation [54]. To estimate the atomic charges for the structures under examination, a thorough natural bond orbital (NBO) analysis was carried out using the NBO program version 3.1 [55]. A Quantum Theory of Atoms in Molecules (QTAIMs) examination was performed utilizing the Multiwfn 3.7 software package [56].

4. Conclusions

An investigation into the interaction of XH3 gases (X = N, P, As) with the MnPc molecule, employing DFT calculations and external electric field modulation, has provided valuable insights into the adsorption behavior and its consequences on the electronic and optical properties of MnPc. The confirmed chemical adsorption of XH3 onto MnPc, supported by QTAIM analysis indicating partially covalent bonds, underscores the relevance of this study in elucidating the nature of the XH3/MnPc interaction.
The significant reduction in the energy gap (Eg) of MnPc upon XH3 adsorption, accompanied by the observed blue shift in the UV-vis spectrum, not only reveals the sensitivity of MnPc to XH3 molecules but also suggests its potential application as a visual sensor. The thermodynamic calculations establish the exothermic nature of the interactions, with NH3/MnPc emerging as the most stable complex. The temperature-dependent stability of NH3/MnPc complexes adds a nuanced understanding of the dynamics of the adsorption process.
Moreover, the influence of an external electric field on the adsorption energy (Eads) highlights the tunability of MnPc’s sensitivity to XH3 gases. The magnitude of the applied electric field plays a pivotal role, emphasizing the importance of external factors in tailoring the adsorption characteristics.
In conclusion, this study not only contributes to the fundamental understanding of XH3/MnPc interactions at a molecular level but also opens up avenues for potential applications in sensor technologies. The versatility of MnPc, demonstrated through its responsiveness to external stimuli and the dynamic nature of the adsorption process, suggests promising prospects for its utilization in sensing applications. Further exploration in this direction could lead to the development of innovative materials for gas sensing and related technologies.

Author Contributions

Conceptualization, H.M.B., K.M.E., H.O.A.-N. and H.Y.A.; methodology, H.M.B., K.M.E., H.O.A.-N. and H.Y.A.; software, H.Y.A.; validation, H.M.B., K.M.E., H.O.A.-N. and H.Y.A.; formal analysis, H.M.B., K.M.E., H.O.A.-N. and H.Y.A.; investigation, H.M.B., K.M.E., H.O.A.-N. and H.Y.A.; resources, H.M.B. and H.Y.A.; data curation, H.M.B., K.M.E., H.O.A.-N. and H.Y.A.; writing—original draft preparation, H.M.B., K.M.E., H.O.A.-N. and H.Y.A., writing—review and editing, H.M.B., K.M.E., H.O.A.-N. and H.Y.A.; visualization, H.M.B., K.M.E., H.O.A.-N. and H.Y.A.; supervision, K.M.E. and H.Y.A.; project administration, H.Y.A.; funding acquisition, H.M.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Scientific Research at Najran University grant number NU/DRP/SERC/12/10.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors are thankful to the Scientific Research at Najran University for funding this work under the General Research Funding program grant code NU/DRP/SERC/12/10.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Srivastava, A.; Bhat, C.; Jain, S.K.; Mishra, P.K.; Brajpuriya, R. Electronic transport properties of BN sheet on adsorption of ammonia (NH3) gas. J. Mol. Model. 2015, 21, 39. [Google Scholar] [CrossRef] [PubMed]
  2. Ammar, H.; Badran, H.; Eid, K. TM-doped B12N12 nano-cage (TM = Mn, Fe) as a sensor for CO, NO, and NH3 gases: A DFT and TD-DFT study. Mater. Today Commun. 2020, 25, 101681. [Google Scholar] [CrossRef]
  3. Badran, H.; Eid, K.; Ammar, H. A DFT study on the effect of the external electric field on ammonia interaction with boron nitride nano-cage. J. Phys. Chem. Solids 2020, 141, 109399. [Google Scholar] [CrossRef]
  4. Habibi-Yangjeh, A.; Basharnavaz, H.; Kamali, S.H.; Nematollahzadeh, A. A first-principles investigation of PH3 gas adsorption on the graphitic carbon nitride sheets modified with V/P, Nb/P, and Ta/P elements. Mater. Chem. Phys. 2021, 269, 124282. [Google Scholar] [CrossRef]
  5. Sagisaka, K.; Marz, M.; Fujita, D.; Bowler, D. Adsorption of phosphorus molecules evaporated from an InP solid source on the Si(100) surface. Phys. Rev. B 2013, 87, 155316–155324. [Google Scholar] [CrossRef]
  6. Buasaeng, P.; Rakrai, W.; Wanno, B.; Tabtimsai, C. DFT investigation of NH3, PH3, and AsH3 adsorptions on Sc-, Ti-, V-, and Cr-doped single-walled carbon nanotubes. Appl. Surf. Sci. 2017, 400, 506–514. [Google Scholar] [CrossRef]
  7. Holleman, A.F.; Wiberg, E. Inorganic Chemistry; Academic Press: San Diego, CA, USA, 2001; ISBN 0123526515. [Google Scholar]
  8. Luo, H.; Zhang, L.; Xu, S.; Shi, M.; Wu, W.; Zhang, K. NH3, PH3 and AsH3 adsorption on alkaline earth metal (Be-Sr) doped graphenes: Insights from DFT calculations. Appl. Surf. Sci. 2021, 537, 147542. [Google Scholar] [CrossRef]
  9. Luo, H.; Xu, K.; Gong, Z.; Li, N.; Zhang, K.; Wu, W. NH3, PH3, AsH3 adsorption and sensing on rare earth metal doped graphene: DFT insights. Appl. Surf. Sci. 2021, 566, 150390. [Google Scholar] [CrossRef]
  10. Ranea, V.A.; Quiña, P.L.D.; Yalet, N.M. General adsorption model for H2S, H2Se, H2Te, NH3, PH3, AsH3 and SbH3 on the V2O5(0 0 1) surface including the van der Waals interaction. Chem. Phys. Lett. 2019, 720, 58–63. [Google Scholar] [CrossRef]
  11. Winder, E.J.; Moore, D.E.; Neu, D.R.; Ellis, A.B.; Geisz, J.F.; Kuech, T.F. Detection of ammonia, phosphine, and arsine gases by reversible modulation of cadmium selenide photoluminescence intensity. J. Cryst. Growth 1995, 148, 63–69. [Google Scholar] [CrossRef]
  12. Keleş, T.; Biyiklioglu, Z.; Özel, A. A comparative study on DNA/BSA binding, DNA photocleavage and antioxidant activities of water soluble peripherally and non-peripherally tetra-3-pyridin-3-ylpropoxy-substituted Mn(III), Cu(II) phthalocyanines. Dye. Pigment. 2017, 139, 575–586. [Google Scholar] [CrossRef]
  13. Ndebele, N.; Nyokong, T. Electrocatalytic behaviour of chalcone substituted Co, Cu, Mn and Ni phthalocyanines towards the detection of nitrite. J. Electroanal. Chem. 2022, 926, 116951. [Google Scholar] [CrossRef]
  14. Soliman, I.M.; El-Nahass, M.M.; Eid, K.M.; Ammar, H.Y. Vibrational spectroscopic analysis of aluminum phthalocyanine chloride. experimental and DFT study. Phys. B 2016, 491, 98–103. [Google Scholar] [CrossRef]
  15. Caliskan, S.; Laref, A. Manipulation of quantum spin transport of TM (Mn, Fe, Co, and Ni) (II) phthalocyanines coupled to carbon nanotube leads applicable in spintronic devices. Diam. Relat. Mater. 2023, 139, 110274. [Google Scholar] [CrossRef]
  16. Yalazan, H.; Maden, Y.E.; Koca, A.; Kantekin, H. Multi-step syntheses, electrochemistry and spectroelectrochemistry of peripheral CoII, CuII and MnIII Cl phthalocyanines bearing pyrazoline. J. Mol. Struct. 2022, 1269, 133788. [Google Scholar] [CrossRef]
  17. Zhang, X.; Zhao, P. Molecular logic gates based on spin caloritronic transport properties of Mn phthalocyanine nanoribbon. Phys. Lett. A 2020, 384, 126256. [Google Scholar] [CrossRef]
  18. Zhang, Q.; Mao, J.; Peng, W.; Li, H.; Qian, L.; Yang, W.; Liu, J. A DFT study on selective adsorption of NH3 from ammonia synthesis tail gas with typical aromatic boranes. Mater. Today Commun. 2023, 37, 107495. [Google Scholar] [CrossRef]
  19. Larki, S.; Shakerzadeh, E.; Anota, E.C.; Behjatmanesh-Ardakani, R. The Al, Ga and Sc dopants effect on the adsorption performance of B12N12 nanocluster toward pnictogen hydrides. Chem. Phys. 2019, 526, 110424. [Google Scholar] [CrossRef]
  20. Janczak, J.; Kubiak, R.; Sledz, M.; Borrmann, H.; Grin, Y. Synthesis, structural investigations and magnetic properties of dipyridinated manganese phthalocyanine, MnPc(py)2. Polyhedron 2003, 22, 2689–2697. [Google Scholar] [CrossRef]
  21. Zouaghi, M.O.; Arfaoui, Y.; Champagne, B. Density functional theory investigation of the electronic and optical properties of metallo-phthalocyanine derivatives. Opt. Mater. 2021, 120, 111315. [Google Scholar] [CrossRef]
  22. Shalabi, A.S.; Aal, S.A.; Assem, M.M.; Soliman, K.A. Metallophthalocyanine and Metallophthalocyanine–fullerene complexes as potential dye sensitizers for solar cells DFT and TD-DFT calculations. Org. Electron. 2012, 13, 2063–2074. [Google Scholar] [CrossRef]
  23. Caplins, B.W.; Mullenbach, T.K.; Holmes, R.J.; Blank, D.A. Femtosecond to nanosecond excited state dynamics of vapor deposited copper phthalocyanine thin films. Phys. Chem. Chem. Phys. 2016, 18, 11454–11459. [Google Scholar] [CrossRef] [PubMed]
  24. Serra, O.A.; Iamamoto, Y. Chromophores: Porphyrin-Based Materials, Encyclopedia of Materials: Science and Technology, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2001; pp. 1227–1229. [Google Scholar] [CrossRef]
  25. Eid, K.; Ammar, H. Adsorption of SO2 on Li atoms deposited on MgO (100) surface: DFT calculations. Appl. Surf. Sci. 2011, 257, 6049–6058. [Google Scholar] [CrossRef]
  26. AShalabi, S.; Aal, S.A.; Kamel, M.A.; Taha, H.O.; Ammar, H.Y.; Halim, W.S.A. The role of oxidation states in FA1 Tln+ (n = 1, 3) lasers and CO interactions at the (1 0 0) surface of NaCl: An ab initio study. Chem. Phys. 2006, 328, 8–16. [Google Scholar] [CrossRef]
  27. Shalabi, A.S.; Aal, S.A.; Halim, W.S.A.; Ammar, H.Y. Artificial polarization effects on FA1:Sr2+ lasers and NO interactions at NaCl (001) surface: First principles calculations. J. Mol. Struct. Theochem 2007, 823, 47–58. [Google Scholar] [CrossRef]
  28. Ziółkowski, M.; Grabowski, S.J.; Leszczynski, J. Cooperativity in Hydrogen-Bonded Interactions: Ab Initio and “atoms in molecules” analyses. J. Phys. Chem. A 2006, 110, 6514–6521. [Google Scholar] [CrossRef] [PubMed]
  29. Celik, S.; Tanış, E. Toxic potential of poly-hexamethylene biguanide hydrochloride (PHMB): A DFT, AIM and NCI analysis study with solvent effects. Comput. Theor. Chem. 2022, 1212, 113709. [Google Scholar] [CrossRef]
  30. Srivastava, A.K. DFT and QTAIM studies on the reduction of carbon monoxide by superalkalis. J. Mol. Graph. Model. 2021, 102, 107765. [Google Scholar] [CrossRef] [PubMed]
  31. Ammar, H.Y.; Badran, H.M. Effect of CO adsorption on properties of transition metal doped porphyrin: A DFT and TD-DFT study. Heliyon 2019, 5, e02545. [Google Scholar] [CrossRef]
  32. Jigang, W.; Ji, L.; Yong, D.; Yan, Z.; Lihui, M.; Asadi, H. Effect of platinum on the sensing performance of ZnO nanocluster to CO gas. Solid State Commun. 2020, 316–317, 113954. [Google Scholar] [CrossRef]
  33. Sconza, A.; Torzo, G. An undergraduate laboratory experiment for measuring the energy gap in semiconductors. Eur. J. Phys. 1989, 10, 123–126. [Google Scholar] [CrossRef]
  34. Kittel, C. Introduction to Solid State Physics; John Wiley & Sons: Hoboken, NJ, USA, 2005; pp. 216–226. [Google Scholar]
  35. Kasap, S.O. Principles of Electronic Materials and Devices; McGraw-Hill: Boston, UK, 2006; pp. 378–405. [Google Scholar]
  36. Ammar, H.Y.; Badran, H.M.; Umar, A.; Fouad, H.; Alothman, O.Y. ZnO nanocrystal-based chloroform detection: Density Functional Theory (DFT) study. Coatings 2019, 9, 769. [Google Scholar] [CrossRef]
  37. He, C.; Zhang, M.; Li, T.T.; Zhang, W.X. Electric field-modulated high sensitivity and selectivity for NH3 on α-C2N2 nanosheet: Insights from DFT calculations. Appl. Surf. Sci. 2020, 505, 144619. [Google Scholar] [CrossRef]
  38. He, C.; Zhang, M.; Li, T.; Zhang, W. A novel C6N2 monolayer as a potential material for charge-controlled CO2 capture. J. Mater. Chem. C 2020, 8, 6542–6551. [Google Scholar] [CrossRef]
  39. Ammar, H.Y.; Eid, K.M.; Badran, H.M. TM-doped Mg12O12 nano-cages for hydrogen storage applications: Theoretical study. Results Phys. 2022, 35, 105349. [Google Scholar] [CrossRef]
  40. Abbasi, M.; Nemati-Kande, E.; Mohammadi, M.D. Doping of the first row transition metals onto B12N12 nanocage: A DFT study. Comput. Theor. Chem. 2018, 1132, 1–11. [Google Scholar] [CrossRef]
  41. Alam, M.S.; Lee, D. Molecular structure, spectral (FT-IR, FT-Raman, Uv-Vis, and fluorescent) properties and quantum chemical analyses of azomethine derivative of 4-aminoantipyrine. J. Mol. Struct. 2021, 1227, 129512. [Google Scholar] [CrossRef]
  42. Airam, S.V.; Ramesh, S. Engineering Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 2013. [Google Scholar]
  43. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. A 1965, 140, 1133. [Google Scholar] [CrossRef]
  44. Grimme, S. Semiempirical GGA-Type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787. [Google Scholar] [CrossRef]
  45. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104–154119. [Google Scholar] [CrossRef]
  46. Yang, H.; Bashir, B.; Luo, G. Towards superior metal phthalocyanine catalysts for electrochemical oxygen reduction: A comprehensive screening under experimental conditions. Chem. Eng. J. 2023, 473, 145101. [Google Scholar] [CrossRef]
  47. Badran, H.M.; Eid, K.M.; Ammar, H.Y. DFT and TD-DFT studies of halogens adsorption on cobalt-doped porphyrin: Effect of the external electric field. Results Phys. 2021, 23, 103964. [Google Scholar] [CrossRef]
  48. Badran, H.M.; Eid, K.M.; Al-Nadary, H.O.; Ammar, H.Y. DFT and TD-DFT calculations for electronic, magnetic, and optical characteristics of the 3d transition metal complexes for hexaazabipyH2. Comput. Theor. Chem. 2023, 1226, 114215. [Google Scholar] [CrossRef]
  49. Roy, D.R.; Shah, E.V.; Roy, S.M. Optical activity of Co-porphyrin in the light of IR and Raman spectroscopy: A critical DFT investigation. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 190, 121–128. [Google Scholar] [CrossRef] [PubMed]
  50. Pearson, R.G. Absolute electronegativity and hardness correlated with molecular orbital theory. Proc. Natl. Acad. Sci. USA 1986, 2683, 8440–8441. [Google Scholar] [CrossRef] [PubMed]
  51. Parr, R.G.; Szentpaly, L.V.; Liu, S. Electrophilicity index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  52. Guo, C.; Wang, C. A theoretical study on cage-like clusters (C12-Ti6 and C12-Ti62+) for dihydrogen storage. Int. J. Hydrogen Energy 2019, 44, 10763–10769. [Google Scholar] [CrossRef]
  53. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  54. O’Boyle, N.M.; Tenderholt, A.L. cclib: A library for package-independent computational chemistry algorithms. J. Comput. Chem. 2008, 29, 839–845. [Google Scholar] [CrossRef]
  55. Glendening, E.D.; Reed, A.E.; Carpenter, J.E.; Weinhold, F. Natural Bond Orbital; Version 3.1; Theoretical Chemistry Institute and Department of Chemistry, University of Wisconsin: Madison, WI, USA, 1998. [Google Scholar]
  56. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
Figure 1. Geometrical structure, DOS (HOMO and LUMO are inserted), and MESP for (a) NH3, (b) PH3, and (c) AsH3. Red and blue colors represent negative and positive values of MESP, respectively.
Figure 1. Geometrical structure, DOS (HOMO and LUMO are inserted), and MESP for (a) NH3, (b) PH3, and (c) AsH3. Red and blue colors represent negative and positive values of MESP, respectively.
Molecules 29 02168 g001
Figure 2. (a,b) The optimized structures and (c,d) molecular electrostatic potential (MESP) in atomic units for Pc and MnPc, respectively.
Figure 2. (a,b) The optimized structures and (c,d) molecular electrostatic potential (MESP) in atomic units for Pc and MnPc, respectively.
Molecules 29 02168 g002
Figure 3. (ac) HOMO, DOS, and LUMO for Pc, (df) HOMO, PDOS, and LUMO for MnPc, and (g) UV-vis spectra for Pc and MnPc, respectively.
Figure 3. (ac) HOMO, DOS, and LUMO for Pc, (df) HOMO, PDOS, and LUMO for MnPc, and (g) UV-vis spectra for Pc and MnPc, respectively.
Molecules 29 02168 g003
Figure 4. The proposed adsorption modes for the XH3 molecule on the MnPc molecule (a) mode 1, (b) mode 2, and (c) mode 3.
Figure 4. The proposed adsorption modes for the XH3 molecule on the MnPc molecule (a) mode 1, (b) mode 2, and (c) mode 3.
Molecules 29 02168 g004
Figure 5. Top and side views of the optimized structures (ac) for adsorption mode 1 and (df) for adsorption mode 2 for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively.
Figure 5. Top and side views of the optimized structures (ac) for adsorption mode 1 and (df) for adsorption mode 2 for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively.
Molecules 29 02168 g005
Figure 6. Bond critical points of type (3, −1): (ac) for adsorption mode 1 and (df) for adsorption mode 2 for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively.
Figure 6. Bond critical points of type (3, −1): (ac) for adsorption mode 1 and (df) for adsorption mode 2 for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively.
Molecules 29 02168 g006
Figure 7. (a) Orientation and (b) electrostatic potential (ESP) for XH3 and MnPc molecules.
Figure 7. (a) Orientation and (b) electrostatic potential (ESP) for XH3 and MnPc molecules.
Molecules 29 02168 g007
Figure 8. Charge density difference (Δρ) at 0.001 au isovalue (ac) for adsorption mode 1 and (df) for adsorption mode 2 for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively. Red and blue colors refer to negative and positive Δρ values.
Figure 8. Charge density difference (Δρ) at 0.001 au isovalue (ac) for adsorption mode 1 and (df) for adsorption mode 2 for NH3/MnPc, PH3/MnPc, and AsH3/MnPc, respectively. Red and blue colors refer to negative and positive Δρ values.
Molecules 29 02168 g008
Figure 9. HOMO, PDOS, and LUMO for adsorption mode 1 (a) NH3/MnPc, (b) PH3/MnPc, and (c) AsH3/MnPc. The dashed line refers to the Fermi level.
Figure 9. HOMO, PDOS, and LUMO for adsorption mode 1 (a) NH3/MnPc, (b) PH3/MnPc, and (c) AsH3/MnPc. The dashed line refers to the Fermi level.
Molecules 29 02168 g009
Figure 10. HOMO, PDOS, and LUMO for adsorption mode 2 (a) NH3/MnPc, (b) PH3/MnPc, and (c) AsH3/MnPc. The dashed line refers to the Fermi level.
Figure 10. HOMO, PDOS, and LUMO for adsorption mode 2 (a) NH3/MnPc, (b) PH3/MnPc, and (c) AsH3/MnPc. The dashed line refers to the Fermi level.
Molecules 29 02168 g010
Figure 11. UV-vis spectra for XH3/Pc complexes for (a) adsorption mode 1 and (b) adsorption mode 2.
Figure 11. UV-vis spectra for XH3/Pc complexes for (a) adsorption mode 1 and (b) adsorption mode 2.
Molecules 29 02168 g011
Figure 12. Enthalpy difference (ΔH), Gibbs free energy difference (ΔG), and the logarithm of the thermodynamic adsorption equilibrium constant (log K) for adsorption mode 1 (a) NH3/Pc, (b) PH3/Pc, and (c) AsH3/Pc.
Figure 12. Enthalpy difference (ΔH), Gibbs free energy difference (ΔG), and the logarithm of the thermodynamic adsorption equilibrium constant (log K) for adsorption mode 1 (a) NH3/Pc, (b) PH3/Pc, and (c) AsH3/Pc.
Molecules 29 02168 g012
Figure 13. Electric field direction relative to the XH3/MnPc complex (X = N, P, and As).
Figure 13. Electric field direction relative to the XH3/MnPc complex (X = N, P, and As).
Molecules 29 02168 g013
Figure 14. The dipole moment vs. the electric field for (a) free XH3 gases and (b) MnPc and XH3/MnPc complexes.
Figure 14. The dipole moment vs. the electric field for (a) free XH3 gases and (b) MnPc and XH3/MnPc complexes.
Molecules 29 02168 g014
Figure 15. The adsorption energies (Eads) vs. the electric field for XH3/MnPc complexes.
Figure 15. The adsorption energies (Eads) vs. the electric field for XH3/MnPc complexes.
Molecules 29 02168 g015
Figure 16. (a) Eg and (b) ΔEg% versus the electric field for MnPc and XH3/MnPc complexes.
Figure 16. (a) Eg and (b) ΔEg% versus the electric field for MnPc and XH3/MnPc complexes.
Molecules 29 02168 g016
Table 1. Electronic properties of Pc and MnPc. HOMO and LUMO energy levels (eV) for α and β spins, HOMO-LUMO gap (Eg, eV), average binding energy per atom (Eb, eV), NBO charges (Q, e), ionization potential (IP, eV), chemical potential (µ, eV), hardness (η, eV), electrophilicity (ω, eV), and dipole moment (D, Debye).
Table 1. Electronic properties of Pc and MnPc. HOMO and LUMO energy levels (eV) for α and β spins, HOMO-LUMO gap (Eg, eV), average binding energy per atom (Eb, eV), NBO charges (Q, e), ionization potential (IP, eV), chemical potential (µ, eV), hardness (η, eV), electrophilicity (ω, eV), and dipole moment (D, Debye).
PcMnPc
HOMO (α)−5.305−5.242
LUMO (α)−3.185−3.117
HOMO (β)−5.305−4.798
LUMO (β)−3.185−3.410
Eg2.1201.388
Eb−5.402−5.474
QTM-0.954
IP5.3054.798
µ4.2454.104
η1.060.694
ω8.49912.134
D0.0210.010
Table 2. Adsorption properties of XH3 (X = N, P, As) on MnPc. Adsorption energies (Eads, eV), HOMO and LUMO energy levels (eV) for α and β spins, HOMO-LUMO gap (Eg, eV), NBO charges (Q, e), and dipole moment (D, Debye).
Table 2. Adsorption properties of XH3 (X = N, P, As) on MnPc. Adsorption energies (Eads, eV), HOMO and LUMO energy levels (eV) for α and β spins, HOMO-LUMO gap (Eg, eV), NBO charges (Q, e), and dipole moment (D, Debye).
Mode 1Mode 2
NH3/MnPcPH3/MnPcAsH3/MnPcNH3/MnPcPH3/MnPcAsH3/MnPc
Eads−0.777−0.414−0.370−0.307−0.340−0.344
dX-Mn2.3332.7862.9542.9953.5423.612
HOMO (α)−5.079−5.136−5.163−5.285−5.250−5.229
LUMO (α)−2.984−3.035−3.058−3.176−3.133−3.111
HOMO (β)−4.427−4.582−4.628−4.794−4.781−4.755
LUMO (β)−3.230−3.312−3.340−3.459−3.418−3.396
Eg1.1961.2711.2881.3351.3631.359
QM0.8270.8300.7670.9880.8800.898
Q X H 3 0.1770.3240.2260.0840.1000.101
D3.1302.0831.5760.9490.2280.213
Table 3. The estimated topological parameters. Electron densities (ρ), Laplacian of charge density (∇2ρ), kinetic electron density (G(r)), potential energy density (V(r)), and energy density (H(r)). All units are in au.
Table 3. The estimated topological parameters. Electron densities (ρ), Laplacian of charge density (∇2ρ), kinetic electron density (G(r)), potential energy density (V(r)), and energy density (H(r)). All units are in au.
Adsorption ModeComplexBCPρ2ρG(r)V(r)H(r)−G(r)/V(r)
1NH3/MnPcN–Mn0.0450.1710.045−0.047−0.0020.952
PH3/MnPcP–Mn0.0320.0640.021−0.026−0.0050.813
AsH3/MnPcAs–Mn0.0250.0460.015−0.018−0.0030.833
2NH3/MnPcN–Mn0.0130.0300.008−0.009−0.0010.897
PH3/MnPcP–Mn0.0080.0200.005−0.0040.0001.097
H–N0.0070.0190.004−0.0040.0011.155
AsH3/MnPcAs–Mn0.0070.0190.004−0.0040.0001.109
H–N0.0070.0170.004−0.0030.0011.197
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Badran, H.M.; Eid, K.M.; Al-Nadary, H.O.; Ammar, H.Y. DFT-D3 and TD-DFT Studies of the Adsorption and Sensing Behavior of Mn-Phthalocyanine toward NH3, PH3, and AsH3 Molecules. Molecules 2024, 29, 2168. https://doi.org/10.3390/molecules29102168

AMA Style

Badran HM, Eid KM, Al-Nadary HO, Ammar HY. DFT-D3 and TD-DFT Studies of the Adsorption and Sensing Behavior of Mn-Phthalocyanine toward NH3, PH3, and AsH3 Molecules. Molecules. 2024; 29(10):2168. https://doi.org/10.3390/molecules29102168

Chicago/Turabian Style

Badran, Heba Mohamed, Khaled Mahmoud Eid, Hatim Omar Al-Nadary, and Hussein Youssef Ammar. 2024. "DFT-D3 and TD-DFT Studies of the Adsorption and Sensing Behavior of Mn-Phthalocyanine toward NH3, PH3, and AsH3 Molecules" Molecules 29, no. 10: 2168. https://doi.org/10.3390/molecules29102168

Article Metrics

Back to TopTop