Next Article in Journal
Controlled Growth of WO3 Photoanode under Various pH Conditions for Efficient Photoelectrochemical Performance
Previous Article in Journal
Personalised Medicine and the Potential Role of Electrospinning for Targeted Immunotherapeutics in Head and Neck Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

M-Encapsulated Be12O12 Nano-Cage (M = K, Mn, or Cu) for CH2O Sensing Applications: A Theoretical Study

by
Hatim Omar Al-Nadary
1,
Khaled Mahmoud Eid
2,
Heba Mohamed Badran
1,* and
Hussein Youssef Ammar
1,*
1
Physics Department, College of Science & Arts, Najran University, Najran 11001, Saudi Arabia
2
Physics Department, Faculty of Education, Ain Shams University, Cairo 11566, Egypt
*
Authors to whom correspondence should be addressed.
Nanomaterials 2024, 14(1), 7; https://doi.org/10.3390/nano14010007
Submission received: 19 November 2023 / Revised: 9 December 2023 / Accepted: 11 December 2023 / Published: 19 December 2023
(This article belongs to the Section Theory and Simulation of Nanostructures)

Abstract

:
DFT and TD-DFT studies of B3LYP/6–31 g(d,p) with the D2 version of Grimme’s dispersion are used to examine the adsorption of a CH2O molecule on Be12O12 and MBe12O12 nano-cages (M = K, Mn, or Cu atom). The energy gap for Be12O12 was 8.210 eV, while the M encapsulation decreased its value to 0.685–1.568 eV, whereas the adsorption of the CH2O gas decreased the Eg values for Be12O12 and CuBe12O12 to 4.983 and 0.876 eV and increased its values for KBe12O12 and MnBe12O12 to 1.286 and 1.516 eV, respectively. The M encapsulation enhanced the chemical adsorption of CH2O gas with the surface of Be12O12. The UV-vis spectrum of the Be12O12 nano-cage was dramatically affected by the M encapsulation as well as the adsorption of the CH2O gas. In addition, the adsorption energies and the electrical sensitivity of the Be12O12 as well as the MBe12O12 nano-cages to CH2O gas could be manipulated with an external electric field. Our results may be fruitful for utilizing Be12O12 as well as MBe12O12 nano-cages as candidate materials for removing and sensing formaldehyde gas.

1. Introduction

Recently, the problem of pollution of the air, soil, and water has attracted the attention of many scientists. There have been several attempts to reduce pollution sources, whether by capturing the pollutants or detecting the contaminated materials. Formaldehyde (CH2O) is considered one of these pollutants. That is because of its diverse household uses in addition to its assorted uses in several industries [1,2,3,4,5,6,7,8]. CH2O gas has a pungent smell and no color [9]. Exposure to CH2O gas can cause voluminous hazards or even death to humans [10,11,12,13]. Accordingly, it is necessary to search for a material capable of removing the CH2O or that can be used as a CH2O sensor. Beryllium oxide (BeO) has some electronic features, making it a candidate material for this purpose. BeO is a semiconductor possessing an energy gap (Eg) value of 8.21 eV [14]. The Be-O bond appears to be a shared behavior of ionic and covalent bonds [15]. Several experimental and theoretical efforts have been made to investigate BeO nano-structures, which have been synthesized in various forms such as nano-fibers [16] and nano-particles [17,18]. The geometrical and electrical characteristics of diverse nano-clusters of BeO have been inspected, and this has proven that a Be12O12 nano-cage possesses adequate stability [19]. Moreover, Be12O12 is a semiconductor with Eg = 7.41–8.29 eV [20,21,22,23,24,25]. Furthermore, metal oxide gas sensors have desirable properties such as a small size, lower cost, and extended lifetime [26]. Accordingly, Be12O12 nano-cages are employed for various uses such as a catalyst to convert methane to organic compounds [27]; a sensor for sulfur mustard [23], tabun, mercaptopyridine, formaldehyde, sulfur hydride, and sulfur dioxide [20,22,28,29]; and a hydrogen storage material [25]. The impact of Li, Na, and K deposition and encapsulation on the electronic and non-linear optical properties of Be12O12 has been studied [24,28]. It was found that the deposition and encapsulation of the alkali metals sharply decreased the Eg of the nano-cage to 1.10–1.56 eV and 0.68–0.69 eV, respectively. Additionally, it proved that the non-linear optical properties of Be12O12 nano-cages could be modified by the alkali metals. The UV-vis spectra for the Be12O12 nano-cage were predicted by Fallahi et al. [22] and Jouypazadeh et al. [23]. They found that the λmax absorbance peak for a Be12O12 nano-cage is located at 140 to 150 nm. Kosar et al. [30] found that the encapsulation of Be12O12 with Be, Mg, and Ca atoms led to a red shift that was located at 505 nm, 781 nm, and 1136 nm, respectively.
Several efforts have been made to sense the CH2O molecules [31,32]. Diverse compounds such as MgO [33], BN [34,35,36,37], N-doped TiO2 [38], ZnO [4,39], transition metal-doped MoS2 [40], SnS [41], In2O3 [42], SnO2 [43], NiO [44], and BeO [20] have been employed as CH2O sensors.
According to previous studies [45,46,47,48,49,50,51,52], the existence of an external electric field (EF) impacts the geometric structure, substrate–adsorbate interaction, and electrical characteristics. Additionally, the electric field has a significant impact on key sensor properties such as the adsorption energy (Eads), sensitivity, and recovery time (τ). Our previous study [20] focused on studying the effect of the EF on the interaction characteristics of CH2O on a pristine Be12O12 nano-cage in different solvents. It was discovered that the magnitude and orientation of the EF can alter the sensing parameters of Be12O12 for the CH2O molecule.
According to our knowledge, there is a deficiency in the investigation of the impact of encapsulation of 3d metal atoms on the electrical, optical, and adsorption features of Be12O12. Moreover, no work has been interested in the effect of the existence of an EF on the adsorption features of CH2O molecules on metal-encapsulated Be12O12. Therefore, in this work, we sought to study the effect of an alkali metal (K) or a transition metal (Mn or Cu) as well as the effect of an EF on the adsorption characteristics of the CH2O molecule on a Be12O12 nano-cage.

2. Methods

The BeO nano-cage was represented by 12 Be-O dimers. Then, as a trial to modify the properties of the beryllium oxide nano-cage (Be12O12), an M atom (M = K, Mn, or Cu) was encapsulated into the nano-cage to form MBe12O12 nano-cages (see Figure 1). DFT-D2 computations were performed to inspect the adsorption features of the CH2O molecule on the Be12O12 as well as MBe12O12 nano-cages. To estimate the optimized energetic geometrical structures and the electronic properties, the DFT-D2 computations were carried out by utilizing the hybrid functional B3LYP and 6–31 g(d,p) basis set [53]. The exchange functional B3 refers to Becke’s three parameters, while the correlation functional LYP refers to the correlation functional of Lee, Yang, and Parr [54]. The investigated structures were fully optimized under the following cutoff conditions for the total force on an atom (0.129 eV/Å), the root mean square of the force (0.086 eV/Å), the displacement ( 5.29 × 10 3 Å ), and the root mean square of the displacement ( 3.53 × 10 3 Å ).
D2 refers to Grimme’s dispersion [55,56], which takes into account the van der Waals interactions. The optimal multiplicity M (M = 2S + 1) was examined for the considered nano-cages, where S is the total spin. The dynamic stability for the considered nano-cages was checked by utilizing the vibrational frequencies, which were estimated by using the FT-IR spectra as well as a molecular dynamic simulation using the ADMP model. The UV-vis spectra were estimated for the considered structures by using the time-dependent DFT (TD-DFT) method. In TD-DFT calculations, an adequate number of excited states (n = 15) is estimated to cover all the probable transitions in the appropriate range (0–2000 nm).
The binding energy per atom (Eb) for the investigated structures was assessed using Equation (1) [57].
E b = 1 n E c a g e i = 1 n E i
Here, E c a g e is the optimized nano-cage total energy, E i is the single atom energy, and n is the number of the cage atoms.
The ionization potentials (IPs) for the considered cages were evaluated as shown below [5,58]:
I P = E c a g e + E c a g e
where E c a g e + is the energy of the cage after removing an electron while keeping the same geometry of the neutral cage. The Fermi level (EF), hardness (η), and electrophilicity (ω) were evaluated as shown in Equations (3)–(5), respectively [59,60]:
E F = 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 E F 2 / 2 η  
where E H O M O   E L U M O are the energies of the highest occupied and the lowest unoccupied molecular orbitals, respectively. The adsorption energy (Eads) for CH2O on a nano-cage was calculated by Equation (6):
E a d s = E C H 2 O / c a g e E c a g e + E C H 2 O
where E C H 2 O / c a g e , E c a g e   and E C H 2 O are the energies for the C H 2 O / c a g e complex, the nano-cage and the CH2O molecule, respectively.
For the adsorption of multiple n molecules of CH2O gas, the adsorption energy per molecule ( E ¯ a d s ) was calculated by Equation (7):
E ¯ a d s = 1 n E n C H 2 O / c a g e E c a g e + n E C H 2 O
where E n C H 2 O / c a g e   is the energy for the n C H 2 O / c a g e complex. A negative sign for the Eads and E ¯ a d s indicates an exothermic reaction and more stable products.
The Gaussian 09 program was used to achieve all the calculations, whereas the Gauss View 5 was used for the visualization of the results [61]. The GaussSum 3.0 program estimated the partial density of states (PDOS) of the adsorbates–substrates [62]. The NBO version 3.1 measured the charges of atoms [63]. The Multiwfn 3.7 software package was employed to achieve the Quantum Theory of Atoms in Molecules (QTAIM) analysis [64].

3. Results and Discussion

3.1. Effect of Metal Encapsulation in Be12O12

In this section, the impact of doping on the geometrical, electrical, optical, and magnetic characteristics of the Be12O12 nano-cage was scrutinized. Therefore, full geometrical optimization was performed for the pristine and M-encapsulated Be12O12 nano-cage (M = K, Mn, or Cu). The optimized structures for Be12O12 as well as MBe12O12 are shown in Figure 1. The Be12O12 nano-cage comprises eight hexagonal and six tetragonal rings. Furthermore, the Be-O bonds are distinguished into two types. The first type, denoted by d1, is shared between hexagonal and tetragonal rings. The second type, denoted by d2, is shared between two hexagonal rings. It is found that d1 and d2 have bond lengths of 1.58 and 1.52 Å, respectively, for the Be12O12 nano-cage in good agreement with previous works [20,21,29,65]. Whereas the metal encapsulation into Be12O12 elongates d1 and d2 bonds to be 1.62 and 1.57 Å for KBe12O12, 1.60 and 1.55 Å for MnBe12O12, and 1.61 and 1.55 Å for CuBe12O12.
To investigate the stability of Be12O12 as well as MBe12O12 nano-cages, molecular dynamic (MD) simulations by the ADMP model as implemented in Gaussian 09 were performed. The MD simulations were performed at room temperature (300 K) for 500 fs. Figure S1, in the Supplementary Materials, illustrates the potential energy (PE) fluctuation against time besides the structure of Be12O12 as well as MBe12O12 at the end of the time. One can notice that the PE irrelevantly fluctuates, and unimportant distortion occurs for the cages. Therefore, the Be12O12 and MBe12O12 nano-cages have stable structures. Additionally, in Figure S2, in the Supplementary Materials, the disappearance of the imaginary frequencies for all investigated nano-cages declares that the optimized nano-cages are true minima on the potential energy surfaces [66,67,68,69]. This is in agreement with our previous work on the Be12O12 nano-cage [29].
One can judge the relative stability as well as relative reactivity of Be12O12 and MBe12O12 nano-cages in terms of the following: Eb, HOMO–LUMO energy gap (Eg), IP, EF, η, and ω. Their values are listed in Table 1. The Eb is calculated via Equation (1). The Eb value for Be12O12 is −5.630 eV, which is in agreement with Sajid et al. [21], while the M encapsulation decreases its value by 8.77%, 7.63%, and 4.01% for KBe12O12, MnBe12O12, and CuBe12O12, respectively. This indicates that the MBe12O12 nano-cages are lower in stability and consequently higher in reactivity than the Be12O12 nano-cage. Moreover, the Eg value for Be12O12 is 8.210 eV and it decreased to 0.685, 1.057, and 1.568 eV for KBe12O12, MnBe12O12, and CuBe12O12, respectively, which is in good agreement with previous studies [20,21,65].
Furthermore, the IP, EF, η, and ω values for Be12O12 are 10.135, −4.358, 4.105, and 2.314 eV, respectively. While the IP values decrease by 67.46%, 64.10%, and 48.54%, the EF values lower to −1.786, −1.753, and −2.760 eV, and the η values decrease by 91.65%, 87.12%, and 80.90%, whereas the ω values rise by 101.18%, 25.56%, and 110.07% for KBe12O12, MnBe12O12, and CuBe12O12, respectively. It was stated [69,70,71,72] that the clusters distinguished by lower Eg, IP, and η values and larger ω values are chemically less stable and more reactive. Thus, our results assert that the M-encapsulated Be12O12 nano-cages are chemically higher in reactivity than the Be12O12 nano-cage.
For more intuition, the NBO atomic charges were estimated. As listed in Table 1, for MBe12O12 nano-cages, the K, Mn, and Cu atoms own positive charges of 0.096, 0.237, and 0.108 e.
This indicates the occurrence of charge transfer from the M atoms to the rest of the beryllium oxide nano-cage. Despite the charge transfer that occurred, the charge distribution on the beryllium and oxygen atoms is still symmetric. Therefore, no obvious change in the magnitudes of the dipole moment values of the MBe12O12 nano-cages was observed. Figure 2 represents the charge density difference ( Δ ρ ) for MBe12O12 nano-cages.
Obviously, the M atom is enclosed by positive and negative values of Δ ρ , which indicates the occurrence of a donation-back donation of charges between the metal atom and the rest of the nano-cage atoms. In other words, there is a charge transfer from the M atom to the nano-cage and vice versa. Additionally, the electronic configuration was calculated for M atoms in a free state as well as in the MBe12O12 nano-cages and listed in Table 2. It is obvious that because of the interaction of the K, Mn, and Cu atoms with the rest of the KBe12O12, MnBe12O12, and CuBe12O12 nano-cages atoms, the 4s sublevel loses a charge of 0.54, 0.37, and 0.24 e, and the 4p sublevel gains 0.41, 0.41, and 0.48 e, respectively. Meanwhile, the 3d sublevel gains 0.23 and 0.76 e for KBe12O12 and MnBe12O12 nano-cages, respectively, and it loses 0.33 e for CuBe12O12 nano-cage. This emphasizes the donation-back donation mechanism of charges between the M atom and the rest of the nano-cage atoms.
In Figure 3, the molecular electrostatic potential (MESP) for Be12O12 as well as MBe12O12 nano-cages was displayed. Figure 3a reveals that positive and negative electrostatic potentials surround the Be sites and the O sites, respectively, which is in good agreement with our previous work [20]. Figure 3b–d show that the K, Mn, or Cu encapsulation in the nano-cage has no obvious effect on the MESP distribution around the nano-cage. Therefore, the Be and O sites for Be12O12 as well as MBe12O12 nano-cages are expected to perform as electrophilic and nucleophilic sites, respectively.
Furthermore, the PDOS as well as the surfaces of the frontier orbitals (HOMO and LUMO) for Be12O12 and MBe12O12 nano-cages are depicted in Figure 4. As seen in Figure 4a the HOMO and LUMO of the Be12O12 nano-cage are located at −8.463 and −0.253 eV, respectively. Therefore, Be12O12 nano-cage is a wide-band gap semiconductor possessing an Eg value of 8.21 eV [14].
Additionally, the HOMO is commonly attributed to the 2p orbitals of the O atoms, whereas the LUMO is commonly attributed to the 2p orbitals of the Be atoms. Thus, the HOMO and LUMO are chiefly localized on the O and the Be sites, respectively. Figure 4b–d show that the presence of the K, Mn, and Cu atoms, respectively, causes noticeable changes in the states attributed to the Be12O12 nano-cage. Moreover, for KBe12O12, new donor and acceptor localized states appear at −2.14 and −1.45 eV, respectively. While for MnBe12O12, new donor localized states appear at −6.74, −6.62, −5.88, −3.84, and −2.28 eV, and new acceptor localized states appear at −0.98 and −0.38 eV. Whereas for CuBe12O12, new donor localized states appear at −7.58 and −3.54 and new acceptor localized states appear at −1.97, −0.36, and −0.02 eV. Consequently, the Eg values for KBe12O12, MnBe12O12, and CuBe12O12 are narrowed to 0.69, 1.30, and 1.57 eV, respectively. Furthermore, for MBe12O12 nano-cages, an overlap is observed between the occupied states of the K, Mn, and Cu atoms and the occupied states of the rest of the nano-cage atoms, which indicates an interaction among them has occurred. The electrical conductivity (σ) is governed by the Eg value as follows [73,74,75,76,77,78].
σ T 3 / 2 e x p ( E g 2 k T )
where k is Boltzmann’s constant, and T is the temperature. Subsequently, the metal doping increases the σ value of the Be12O12 nano-cage.
To scrutinize the influence of metal encapsulation on the optical properties of the Be12O12 nano-cage, the UV-vis spectra for the Be12O12, KBe12O12, MnBe12O12, and CuBe12O12 nano-cages were estimated via TD-DFT calculations and graphed in Figure 5. It is obvious the Be12O12 nano-cage has a λmax absorbance peak in the UV region at 169 nm, which is consistent with previous studies [22,23]. Moreover, the KBe12O12 nano-cage has two absorbance peaks; the first peak is located in the visible region at 424 nm, and the second peak is located in the IR region at 1527 nm, while the MnBe12O12 nano-cage has a λmax absorbance peak in the IR region at 889 nm. However, the CuBe12O12 nano-cage has a λmax absorbance peak in the visible region at 431 nm. In other words, the metal encapsulation obviously affects the optical activity of the Be12O12 nano-cage. The Be12O12 nano-cage is an ultraviolet active compound, while the MBe12O12 nano-cages are active compounds in the visible and IR regions.

3.2. Adsorption of CH2O on M-Doped Be12O12

To examine the interaction characteristics of the CH2O molecule with the Be12O12 and MBe12O12 nano-cages, an optimization was carried out for the free CH2O, the bare nano-cages (Be12O12 and MBe12O12), and the CH2O/nano-cages (CH2O/Be12O12 and CH2O/MBe12O12) complexes. Our calculations show that the bond lengths of the C-O and C-H bonds for the CH2O molecule are 1.21 and 1.11 Å, respectively. The geometrical parameters for the bare Be12O12 and MBe12O12 nano-cages were discussed in the previous section. Additionally, in our previous work [20], it was found that negative and positive electrostatic potentials surround the O atom and CH2 group, respectively, of the CH2O molecule. While Figure 3 shows that positive and negative electrostatic potentials surround the Be and O sites, respectively, of the Be12O12 and MBe12O12 nano-cages. Therefore, the interaction of CH2O with the nano-cage is investigated in different orientations, as shown in Figure S3 in the Supplementary Materials. The optimization process for the suggested orientations shows that the CH2O molecule always interacts via its O head with the Be site of the nano-cage, as depicted in Figure 6.
One can see that the distance between the CH2O and the Be12O12 nano-cage (dBe1-O1 = 1.76 Å) is longer than that between the CH2O and the MBe12O12 nano-cage (dBe1-O1 in the range 1.49–1.50 Å). Moreover, the C-O bond length of the CH2O molecule is elongated by 0.83%, 10.74%, 11.57%, and 11.57% for the CH2O/Be12O12, CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12, respectively. Furthermore, the CH2O adsorption elongates the bond lengths between the adsorbing site (Be1) and the neighboring oxygen sites (O2, O3, and O4); this elongation is higher for the CH2O/MBe12O12 complexes than that for the CH2O/Be12O12 complex. Table 3 lists the adsorption features of CH2O on MBe12O12 nano-cages.
One can notice that the CH2O adsorption on the Be12O12 as well as the MBe12O12 nano-cages is a chemisorption. Meanwhile, the Eads values for the CH2O/MBe12O12 complexes are more negative than the Eads value for the CH2O/Be12O12 complex. In other words, the presence of the M atom enhances the CH2O adsorption on the nano-cage. Thus, the Be12O12 as well as MBe12O12 nano-cages can be utilized as removal materials for formaldehyde gas, whereas the MBe12O12 nano-cages are more efficient than the pristine Be12O12 nano-cage for this purpose. Furthermore, it was found in previous studies that the Eads of CH2O on the B3O3 monolayer [79], Ti-functionalized porphyrin-like C70 fullerenes [80], carbon nano-tube (CNT) [81], Pd-loaded CNT [81], BeO nano-tube [82], Zn12O12 nano-cage [4], and Al-deposited Zn12O12 nano-cage [4] are −0.402, −1.862, −0.106, −1.299, −1.088, −1.27, and −2.23 eV, respectively. Therefore, the present work shows that KBe12O12 and MnBe12O12 nano-cages as CH2O sorbent materials are more efficient than those mentioned in the previous studies. In addition, the CH2O adsorption decreases the Eg value for Be12O12 and CuBe12O12 by 39.31% and 44.13%, respectively, and increases its value for KBe12O12 and MnBe12O12 by 87.74% and 43.42%, respectively. For a more detailed explanation of the results, NBO atomic charge analysis, charge density difference ( Δ ρ ) analysis, QTAIM analysis, and PDOS analysis were performed. The NBO atomic charge analysis shows that for the CH2O/Be12O12 complex, the CH2O molecule earns a positive charge of 0.148 e due to a charge transfer from the CH2O molecule to the Be12O12 nano-cage. On the other side, for CH2O/MBe12O12 complexes, the CH2O molecule acquires negative charges of 0.703, 0.960, and 0.960 e. Furthermore, the positive charges for K, Mn, and Cu rose by 0.347, 0.333, and 0.470 e for CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12, respectively; i.e., a charge transfer has occurred from the MBe12O12 nano-cages, mainly from the metal atom to the CH2O molecule. This may be owing to the lower IP value and consequently higher ability to donate electrons (higher basicity) for the MBe12O12 nano-cages than the Be12O12 nano-cages. Additionally, the electronic configurations of 2s and 2p orbitals for the O and C atoms of CH2O were estimated for the free CH2O, CH2O/Be12O12, and CH2O/MBe12O12 complexes (see Table 4).
It is clear that the interaction of the CH2O molecule with the Be12O12 as well as the MBe12O12 nano-cages is accompanied by losing electrons from the 2s and gaining electrons to the 2p orbitals of the oxygen atom. Whereas for the carbon atom, the 2s orbital has an insignificant change in its electronic configuration, while the 2p orbital loses electrons for the CH2O/Be12O12 complex and gains electrons for the CH2O/MBe12O12 complexes. This means a donation-back donation mechanism has happened between the CH2O molecule and the nano-cages. Figure 7 demonstrates the charge density difference ( Δ ρ ) for CH2O/Be12O12 as well as the CH2O/MBe12O12 complexes. One can see that the CH2O molecule for all the investigated complexes is surrounded by positive (blue color) and negative (red color) Δ ρ values. This means the CH2O molecule loses and gains charges, emphasizing the donation-back donation mechanism between the CH2O molecule and the nano-cages. It is worth noticing that the high dipole moment values for the CH2O/Be12O12 and the CH2O/MBe12O12 complexes in Table 3 refer to the charge redistribution between the CH2O molecule and the nano-cages.
For more insight into the nature of the CH2O adsorption on the Be12O12 as well as the MBe12O12 nano-cage, the QTAIM analysis was accomplished. Figure S4, in the Supplementary Materials, shows the bond critical points (BCP) of type (3,−1) for the CH2O/Be12O12 and the CH2O/MBe12O12 complexes. The topological parameters are stated in Table 5.
The kind of bond can be distinguished by the BCP parameters [83,84,85]. As reported, the ionic bond, weak hydrogen bond, and van der Waals interaction are characterized by 2 ρ > 0, H(r) > 0, −G(r)/V(r) > 1. Furthermore, the strong interaction is categorized by 2 ρ >10−1 au, and the weak interaction is categorized by 2 ρ < 10−1 au. Additionally, the partly covalent interaction is categorized by 2 ρ > 0 and H(r) < 0. Our results show that for the CH2O/Be12O12 complex, there are two BCPs; the first BCP is Be1-O1, which has a 2 ρ of 0.327 au, H(r) of 0.005 au, and −G(r)/V(r) ratio of 1.069, while the other BCP is O2-H1, which has a 2 ρ of 0.054 au, H(r) of 0.002 au, and −G(r)/V(r) ratio of 1.200. This indicates the formation of a pure ionic bond between the O1 atom of the CH2O molecule and the Be1 site of the Be12O12 nano-cage. Meanwhile, a weak hydrogen bond is found between the H1 atom of the CH2O molecule and the O2 site of the Be12O12 nano-cage. For CH2O/MBe12O12 complexes, only one BCP exists between the O1 atom of the CH2O molecule and the Be1 site of the MBe12O12 nano-cage. This BCP has 2 ρ values of 0.858, 0.879, and 0.867, −G(r)/V(r) ratios of 1.005, 1.000, and 1.005, and H(r) values of −0.002, −0.001, and −0.001 for CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12 complexes, respectively. These results categorize the interaction as an ionic interaction with a partially covalent character. Therefore, one can say that the presence of the metal atom has an obvious role in altering the character of the interaction between the CH2O molecule and the nano-cages. Furthermore, the higher ρ values [80] for the CH2O/MBe12O12 complexes explain the higher adsorption of the CH2O molecule than the CH2O/Be12O12 complex.
Figure 8 depicts the HOMO and LUMO surfaces and the PDOS for the CH2O molecule, CH2O/Be12O12, and CH2O/MBe12O12 complexes. Figure 8a shows four occupied states of the CH2O molecule located at −7.30, −10.86, −12.24, and −13.44 eV. Looking at Figure 8b–e, obvious changes in the occupied states of the CH2O molecule indicate a strong reaction between the CH2O and the nano-cages. Regarding Figure 8b, one can see the HOMO of Be12O12 rises to −7.87 eV and an acceptor state is created at −2.89 eV; thus, the Eg value decreases from 8.21 to 4.98 eV. Looking at Figure 4b–d, it is obvious that there are occupied states located at −2.13, −2.28, and −3.54 eV, which are attributed to the metal atoms for KBe12O12, MnBe12O12, and CuBe12O12, respectively. As seen in Figure 8c–e, these states disappear for the complexes CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12.
This confirms that the adsorption of the CH2O leads to a charge loss from the metal atom. Additionally, new occupied states assigned to CH2O appear at −2.67, −3.01, and −2.90 eV for CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12, respectively. This confirms that the adsorption of the CH2O leads to a charge transfer from the MBe12O12 to the CH2O molecule. Furthermore, the adsorption of the CH2O molecule causes a shift for LUMO states from −1.44, −1.22, and −1.98 eV to −1.38, −1.49, and −2.02 eV for KBe12O12, MnBe12O12, and CuBe12O12, respectively. Accordingly, the Eg values increase from 0.69 and 1.06 eV to 1.29 and 1.52 eV for KBe12O12 and MnBe12O12 nano-cages, respectively, while Eg value decreases from 1.57 to 0.88 eV for CuBe12O12 nano-cage. Consequently, due to the CH2O adsorption and Equation (8), the electrical conductivity is increased for the Be12O12 and CuBe12O12, whereas it is decreased for the KBe12O12 and MnBe12O12. Therefore, the Be12O12 and CuBe12O12 nano-cages can be utilized as electrical sensors for the CH2O molecule. In addition, the essential sensing factor, the recovery time (τ), depends on the Eads as in the following equation [20,86,87].
τ = ν ο 1 exp ( E a d s k T )
where ν ο is the attempt frequency, k is Boltzmann’s constant, and T is the temperature. Therefore, the τ value is in the following trend: MnBe12O12 > KBe12O12 > CuBe12O12 >Be12O12.

3.3. Effect of EF

In the present section, the influence of EF on the electronic properties of the CH2O molecule, Be12O12 as well as MBe12O12 nano-cages, and the CH2O/Be12O12 as well as CH2O/MBe12O12 complexes was examined. The EF was considered in the range of −514 to +514 kV/mm with a step of 102.8 kV/mm. All the investigated structures are fully optimized for each EF value. The direction of the EF relative to the investigated structures is sketched in Figure 9. The influence of the EF on the dipole moment for the CH2O molecule and the Be12O12 nano-cage, as well as the MBe12O12 nano-cages, was investigated. Because the EF was applied in the X-axis direction, the change in the dipole moment was only evident in its x-component.
Therefore, the x-component only of the dipole moment was considered. Figure 10a shows the dipole moment for the CH2O molecule versus EF. It is clear that with the EF varying from −514 to +514 kV/mm, the dipole moment of the CH2O molecule decreases. The dipole moments versus EF for the Be12O12 and MBe12O12 nano-cages are illustrated in Figure 10b.
It is seen that with the varying of the EF from −514 to +514 kV/mm, the dipole moment of the investigated nano-cages decreases from 0.311, 2.041, 0.901, and 0.420 Debye to −0.311, −2.032, −0.903, and −0.427 Debye for Be12O12, KBe12O12, MnBe12O12, and CuBe12O12, respectively. It is worth noticing that the polarity of the dipole moment is inverted as the EF direction is inverted. Additionally, the rate of change in the dipole moment with the EF is in this trend: KBe12O12 > MnBe12O12> CuBe12O12> Be12O12. These results manifest that the EF affects the charge distribution for the CH2O molecule, Be12O12, and MBe12O12 nano-cages. This is emphasized by Figure 10b–f, which demonstrates the charge density difference ( Δ ρ ) for the CH2O molecule, Be12O12, and MBe12O12 nano-cages for EF values of −514 and +514 kV/mm. The dipole moment of a molecule plays an important role in its reactivity with the surrounding medium [88]. Therefore, it is expected that the EF would have an obvious impact on the CH2O adsorption on the Be12O12 as well as MBe12O12nano-cages.
The manipulation of the EF on the CH2O adsorption on the Be12O12 and MBe12O12 nano-cages has been completed under the same criteria mentioned above. The Eads values for the CH2O/Be12O12 and CH2O/MBe12O12 complexes are estimated and plotted against the EF in Figure 11.
It is seen that by increasing the negative EF, the value of Eads is gradually enhanced for the CH2O/Be12O12 complex up to 3.4% at EF = −514 kV/mm with respect to their value at zero EF. Meanwhile, the values of Eads are gradually inhibited for CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12 complexes up to 4.8%, 3.6%, and 6.2%, respectively, at EF = −514 kV/mm with respect to its value at zero EF. On the other side, by increasing the positive EF, the value of Eads is gradually inhibited for the CH2O/Be12O12 complex up to 3.1% at EF = +514 kV/mm with respect to its value at zero EF. Meanwhile, the values of Eads are gradually enhanced for CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12 complexes up to 4.2%, 3.5%, and 6.4%, respectively, at EF = +514 kV/mm with respect to their value at zero EF. Therefore, the Eads values for CH2O/Be12O12 and CH2O/MBe12O12 complexes are dominated by either the value or direction of the EF. These results could be explained in terms of the mechanism of CH2O interaction with the nano-cage. For the CH2O/Be12O12 complex, as mentioned before, there is a charge transfer from the CH2O molecule to the Be12O12 nano-cage. Looking at Figure 9b,c, one can see that the negative EF induces a negative Δ ρ value on the oxygen head of the CH2O molecule and a positive Δ ρ value on the side of the Be12O12 nano-cage facing the CH2O molecule. This in turn encourages the charge transfer and, consequently, enhances the Eads value. In contrast, the positive electric field induces a positive Δ ρ value on the oxygen head of the CH2O molecule and a negative Δ ρ value on the side of the Be12O12 nano-cage facing the CH2O molecule. This discourages the charge transfer and, consequently, inhibits the Eads value. This could be confirmed by Figure 12, which shows the NBO charges of the CH2O molecule ( Q C H 2 O ) vs. electric field. It is clear that the positive Q C H 2 O for the CH2O/Be12O12 complex increases as the negative EF increases and decreases as the positive EF increases.
Thus, the Eads value is enhanced by increasing the negative EF and inhibited by increasing the positive EF (see Figure 11). On the other side, for the CH2O/MBe12O12 complexes, the interaction has occurred due to the charge transfer from the MBe12O12 nano-cage to the CH2O molecule. Looking at Figure 9d–f, it is clear that for negative EF values, the induced positive Δ ρ value on the side of the MBe12O12 nano-cage facing the CH2O molecule inhibits this charge transfer and consequently inhibits the Eads values. Whereas for the positive EF values, the induced negative Δ ρ value on the side of the MBe12O12 nano-cage facing the CH2O molecule encourages the charge transfer and thus enhances the Eads values. This can be confirmed by Figure 12, where the negative Q C H 2 O for CH2O/MBe12O12 complexes decreases as the negative EF increases and decreases as the positive EF increases. Thus, the Eads value is inhibited as the negative EF increases and enhanced as the positive EF increases (see Figure 11).
Figure 13 represents the impact of the EF on the Eg for the Be12O12 and MBe12O12 nano-cages as well as the CH2O/Be12O12 and CH2O/MBe12O12 complexes. It is clear that the EF has a negligible impact on the Eg values of the Be12O12 and MBe12O12 nano-cages. Figure 13a declares that the Eg value for the CH2O/Be12O12 complex increases as the negative EF increases and decreases as the positive EF increases. On the other side, Figure 13b–d show that the Eg values for the CH2O/KBe12O12, CH2O/MnBe12O12, CH2O/CuBe12O12 complexes increase as the negative EF decreases and the positive EF increases. Therefore, according to Equation (8), the electric conductivity (σ) for the CH2O/Be12O12 and CH2O/MBe12O12 complexes can be controlled by the EF. Here, σ for the CH2O/Be12O12 complexes will increase by increasing the positive EF, while for the CH2O/MBe12O12 complexes, σ will increase by increasing the negative EF.
To investigate the electrical sensitivity dependence on the EF, the percentage of the variance of Eg (ΔEg) by the adsorption for CH2O/Be12O12 and CH2O/MBe12O12 complexes versus the EF is estimated by Equation (10) and represented in Figure 14.
Δ E g = E g ( C H 2 O / n a n o c a g e ) E g ( n a n o c a g e ) E g ( n a n o c a g e ) × 100
ΔEg values at EF = 0 were −41.2%, 87.8%, 38.9%, and −44.1% for CH2O/Be12O12, CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12 complexes, respectively. It is clear that the variance in value and the direction of the EF has a negligible effect on ΔEg for the CH2O/Be12O12 complex. Whereas for the CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12 complexes, as the negative EF increases, the ΔEg values are lowered, reaching 58.2%, 24.4%, and −55.0%, while as the positive EF increases, the ΔEg values are rising, reaching 114.2%, 53.4%, and −34.4%, respectively. Therefore, the electrical sensitivity of Be12O12 to the adsorption of CH2O molecules is independent of the EF, while the electrical sensitivity of MBe12O12 depends on the EF.

3.4. UV-vis Spectra Investigation

Herein, the effect of the adsorption of the CH2O molecule as well as the EF on the UV-vis spectra of the examined nano-cages was considered. The UV-vis spectra for the CH2O/Be12O12 and CH2O/MBe12O12 complexes are predicted for EF values of −514, 0, and +514 kV/mm, and these are displayed in Figure 15a–c, respectively. It is worth noting that the effect of the EF on the UV-vis spectra of bare nano-cages is studied, and it is found that the EF has no noticeable effect. It is clear that at EF values of −514, 0, and +514 kV/mm, the CH2O/Be12O12 complex has a λmax absorbance peak in the UV region at 224, 252, and 260 nm, respectively. Comparing these results with Figure 5, one can observe that the adsorption of CH2O causes a red shift for the UV-vis spectrum of the Be12O12. The negative EF value decreases the red shift value of λmax, while the positive EF increases it. Furthermore, at zero EF, the CH2O/KBe12O12 exhibits three peaks located at 324 nm, 509 nm, and 810 nm in the UV, visible, and IR regions of the spectra, respectively. The presence of the EF shifts these peaks to 328, 536, and 896 nm at negative EF and to 316, 484, and 736 nm at positive EF. When compared with Figure 5, the Kbe12O12 nano-cage has a peak in the visible region at 424 nm, whereas the adsorption of CH2O induces a red shift in this peak to 536, 509, and 484 for EF values of −514, 0, and +514 kV/mm, respectively. Moreover, at EF values of −514, 0, and +514 kV/mm, the CH2O/MnBe12O12 exhibits one prominent peak in the visible region of the spectra, which is located at 460, 463, and 448 nm, respectively. In comparison with Figure 5, it is evident that the MnBe12O12 nano-cage has one peak in the IR region (889 nm), while the adsorption of CH2O causes a blue shift for λmax to the visible region. Additionally, Figure 5 shows that the CuBe12O12 nano-cage is optically active in the visible region with a λmax value of 431 nm. Meanwhile, the CH2O/CuBe12O12 has peaks in the visible region at 452 and 596 nm at an EF value of −514 kV/mm, 332, 429, and 563 nm at an EF value of 0 kV/mm, and 328, 408 and 528 nm at an EF value of +514 kV/mm. From the above discussion, one can summarize the following: (i) the adsorption of a CH2O molecule affects the UV-vis spectra of Be12O12 as well as the MBe12O12. (ii) The EF has no effect on the UV-vis spectra of Be12O12 as well as the MBe12O12. (iii) The EF has an obvious effect on the UV-vis spectra of CH2O/Be12O12 as well as CH2O/MBe12O12. In other words, the adsorption of the CH2O molecule causes changes in the colors of the MBe12O12 nano-cages; consequently, they could be employed as bare-eye sensors for the CH2O molecule.

3.5. Effect of Concentration

This section concerns the impact of CH2O concentration on the adsorption characteristics. The adsorption of n molecules (n = 1–6) on the surface of the MBe12O12 was investigated to form nCH2O/MBe12O12 complexes. The nCH2O/MBe12O12 complexes were fully optimized. The adsorption energies per molecule ( E ¯ a d s ) were estimated by Equation (7) and graphed in Figure 16a. It is seen that as n increases, the negative value of the E ¯ a d s decreases. Nevertheless, the E ¯ a d s values remain within the bounds of chemical adsorption for all n values.
In Figure 16b, the relationship between Eg and n is illustrated. It is observed that for KBe12O12 and MnBe12O12, as n increases up to n = 3 and 2, respectively, the Eg decreases after which there is no significant change. Conversely, for CuBe12O12, the Eg decreases as n increases, reaching a minimum at n = 4, and then it rises again. According to Equations (8) and (10), it is noted that for all the investigated nano-cages, the electrical conductivity (σ) and the electrical sensitivity at all values of n > 1 are higher than their values at n = 1. Furthermore, the highest electrical conductivity (σ) and electrical sensitivity are recorded for CuBe12O12 at n = 4.

4. Conclusions

This work is a DFT and TD-DFT study that investigates the capability of the M atom-encapsulated Be12O12 nano-cage to capture or sense CH2O gas; M = K, Mn, or Cu. The molecular dynamic simulations and the frequency calculations assert that the Be12O12 and MBe12O12 nano-cages are stable structures. In contrast, the values of the binding energies per atom (Eb), the ionization potential (IP), the hardness (η), and the electrophilicity (ω) prove that the M-encapsulated Be12O12 is chemically more reactive than the Be12O12 cage. In addition, the encapsulation of the M atom into the Be12O12 nano-cage increases its basicity, narrows its energy gap, and alerts its optical activity from an ultraviolet active compound into active compounds in the visible and IR regions.
Moreover, the calculated adsorption energies confirm that the CH2O adsorption on the Be12O12 as well as the MBe12O12 is chemisorption, while the presence of the M atom improves the adsorption of the CH2O molecule on the nano-cage. Moreover, the QTAIM analysis confirms that the presence of the metal atom plays an obvious role in altering the character of CH2O interaction with the nano-cages from pure ionic interaction into an ionic interaction with a partially covalent character. Additionally, the CH2O adsorption decreases the Eg value for Be12O12 and CuBe12O12 by 39.31% and 44.13%, respectively, and increases its value for KBe12O12 and MnBe12O12 by 87.74% and 43.42%, respectively.
It is found that the existence of an external static electric field (EF) can enhance or inhibit the adsorption energies of CH2O molecules on the Be12O12 as well as the MBe12O12 nano-cages, depending on the value and the orientation of the EF. Furthermore, the EF has an obvious influence on the Eg of CH2O/MBe12O12 complexes and, consequently, on their electrical conductivity. Thus, the electrical sensitivity of MBe12O12 nano-cages to CH2O gas can be controlled via EF. Moreover, the CH2O adsorption makes a red shift for the UV-vis spectrum of the Be12O12 and causes obvious changes in the absorption peaks of the MBe12O12 nano-cages in the visible region. Additionally, the EF has an obvious effect on the UV-vis spectra of CH2O/Be12O12 as well as the CH2O/MBe12O12 complexes.
Based on these results, the Be12O12 as well as the MBe12O12 nano-cages are candidate materials for removing and sensing the formaldehyde gas. The MBe12O12 nano-cages may be utilized as electrochemical and naked-eye sensors for CH2O gas.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14010007/s1, Figure S1. Potential energy fluctuations during MD simulation as well as the atomic configuration after 500 fs at 300 K for (a) Be12O12, (b) KBe12O12, (c) MnBe12O12, and (d) CuBe12O12 nano-cages, Figure S2. FT-IR spectra for (a) Be12O12, (b) KBe12O12, (c) MnBe12O12, and (d) CuBe12O12 nano-cages, Figure S3. Non-optimized adsorption modes for CH2O on MBe12O12 nano-cage (M = K, Mn, and Cu). (a, b, and c) by O, H, and C atom of CH2O molecule on Be site of MBe12O12 nano-cage and (d, e, and f) by O, H, and C atom of CH2O molecule on O site of MBe12O12 nano-cage, respectively, Figure S4. Bond critical points of type (3,−1) for (a) CH2O/Be12O12, (b) CH2O/KBe12O12, (c) CH2O/MnBe12O12, and (d) CH2O/CuBe12O12 complexes.

Author Contributions

Conceptualization, K.M.E., H.O.A.-N. and H.Y.A.; methodology, H.O.A.-N., H.M.B., K.M.E. and H.Y.A.; software, 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.O.A.-N., H.M.B., K.M.E. 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.Y.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Najran University grant number NU/DRP/SERC/12/11.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors are thankful to the Deanship of Scientific Research at Najran University for funding this work, under the General Research Funding program grant code (NU/DRP/SERC/12/11).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bunev, V.A.; Shvartsberg, V.M.; Babkin, V.S. Distribution of O atoms from CH2O molecules in the combustion products of formaldehyde. Mendeleev Commun. 2015, 25, 157–158. [Google Scholar] [CrossRef]
  2. Jin, W.; Chen, G.; Duan, X.; Yin, Y.; Ye, H.; Wang, D.; Yu, J.; Mei, X.; Wu, Y. Adsorption behavior of formaldehyde on ZnO(101¯0)surface: A first principles study. Appl. Surf. Sci. 2017, 423, 451–456. [Google Scholar] [CrossRef]
  3. Henda, M.; Sadon, S.H.; Abdelmalek, Z.; Li, Z.; Le, Q.H. Removal of formaldehyde pollutant from petroleum industry wastewaters by polymers: A molecular dynamics simulation. Eng. Anal. Bound. Elem. 2023, 151, 400–405. [Google Scholar] [CrossRef]
  4. Ammar, H.Y. CH2O Adsorption on M (M = Li, Mg and Al) Atom Deposited ZnO Nano-Cage: DFT Study. Key Eng. Mater. 2018, 786, 384–392. [Google Scholar] [CrossRef]
  5. Ammar, H.; Eid, K.; Badran, H. Interaction and detection of formaldehyde on pristine and doped boron nitride nano-cage: DFT calculations. Mater. Today Commun. 2020, 25, 101408. [Google Scholar] [CrossRef]
  6. Salthammer, T. Formaldehyde in the Ambient Atmosphere: From an Indoor Pollutant to an Outdoor Pollutant? Angew. Chem. Int. Ed. 2013, 52, 3320–3327. [Google Scholar] [CrossRef]
  7. Zhu, L.; Jacob, D.J.; Keutsch, F.N.; Mickley, L.J.; Scheffe, R.; Strum, M.; Abad, G.G.; Chance, K.; Yang, K.; Rappenglück, B.; et al. Formaldehyde (HCHO) As a Hazardous Air Pollutant: Mapping Surface Air Concentrations from Satellite and Inferring Cancer Risks in the United States. Environ. Sci. Technol. 2017, 51, 5650–5657. [Google Scholar] [CrossRef]
  8. Salthammer, T.; Mentese, S.; Marutzky, R. Formaldehyde in the Indoor Environment. Chem. Rev. 2010, 110, 2536–2572. [Google Scholar] [CrossRef]
  9. Wang, Z.; Huang, Z.; Yu, J.; Shao, X.; Peng, W.; Yu, J.; Jiang, Y. Growth of Ag/g-C3N4 nanocomposites on nickel foam to enhance photocatalytic degradation of formaldehyde under visible light. J. Environ. Sci. 2024, 137, 432–442. [Google Scholar] [CrossRef]
  10. Cockcroft, D.; Hoeppner, V.; Dolovtch, J. Occupational Asthma Caused by Cedar Urea Formaldehyde Particle Board. Chest 1982, 82, 49–53. [Google Scholar] [CrossRef]
  11. Ye, J.; Zhu, X.; Cheng, B.; Yu, J.; Jiang, C. Few-Layered Graphene-like Boron Nitride: A Highly Efficient Adsorbent for Indoor Formaldehyde Removal. Environ. Sci. Technol. Lett. 2016, 4, 20–25. [Google Scholar] [CrossRef]
  12. Cogliano, V.J.; Grosse, Y.; Baan, R.A.; Straif, K.; Secretan, M.B.; El Ghissassi, F. Meeting Report: Summary of IARC Monographs on Formaldehyde, 2-Butoxyethanol, and 1-tert-Butoxy-2-Propanol. Environ. Health Perspect. 2005, 113, 1205–1208. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, L.; Steinmaus, C.; Eastmond, D.A.; Xin, X.K.; Smith, M.T. Formaldehyde exposure and leukemia: A new meta-analysis and potential mechanisms. Mutat. Res. Mol. Mech. Mutagen. 2009, 681, 150–168. [Google Scholar] [CrossRef] [PubMed]
  14. Toosi, A.R.; Shamlouei, H.R. Transition Metals Doping Effects on Non-Linear Optical Properties of Be12O12 Nano-Cluster: A DFT Study. Phys. Chem. Indian J. 2017, 1, 103. [Google Scholar]
  15. Lenglet, M. Iono-Covalent Character of the Metal–Oxygen Bonds in Oxides: A Comparison of Experimental and Theoretical Data. Act. Passiv. Electron. Compon. 2004, 27, 1–60. [Google Scholar] [CrossRef]
  16. Hossain, S.; Nakane, K. Formation of beryllium oxide nanofibers by polyvinyl alcohol/beryllium sulfate/polyethyleneimine composite precursors. SN Appl. Sci. 2022, 4, 325. [Google Scholar] [CrossRef]
  17. Wang, X.-F.; Wang, R.-C.; Peng, C.-Q.; Li, T.-T.; Liu, B. Synthesis and sintering of beryllium oxide nanoparticles. Prog. Nat. Sci. 2010, 20, 81–86. [Google Scholar] [CrossRef]
  18. Selvaraj, V.; Morri, B.; Nair, L.M.; Krishnan, H. Experimental investigation on the thermophysical properties of beryllium oxide-based nanofluid and nano-enhanced phase change material. J. Therm. Anal. Calorim. 2019, 137, 1527–1536. [Google Scholar] [CrossRef]
  19. Ren, L.; Cheng, L.; Feng, Y.; Wang, X. Geometric and electronic structures of (BeO)N (N = 2–12, 16, 20, and 24): Rings, double rings, and cages. J. Chem. Phys. 2012, 137, 014309. [Google Scholar] [CrossRef]
  20. Badran, H.; Eid, K.; Al-Nadary, H.; Ammar, H. Beryllium oxide nano-cage as sorbent and sensor for formaldehyde gas: DFT-D3 calculations. J. Mol. Liq. 2023, 385, 122430. [Google Scholar] [CrossRef]
  21. Sajid, H.; Siddique, S.A.; Ahmed, E.; Arshad, M.; Gilani, M.A.; Rauf, A.; Imran, M.; Mahmood, T. DFT outcome for comparative analysis of Be12O12, Mg12O12 and Ca12O12 nanocages toward sensing of N2O, NO2, NO, H2S, SO2 and SO3 gases. Comput. Theor. Chem. 2022, 1211, 113694. [Google Scholar] [CrossRef]
  22. Fallahi, P.; Jouypazadeh, H.; Farrokhpour, H. Theoretical studies on the potentials of some nanocages (Al12N12, Al12P12, B12N12, Be12O12, C12Si12, Mg12O12 and C24) on the detection and adsorption of Tabun molecule: DFT and TD-DFT study. J. Mol. Liq. 2018, 260, 138–148. [Google Scholar] [CrossRef]
  23. Jouypazadeh, H.; Farrokhpour, H. DFT and TD-DFT study of the adsorption and detection of sulfur mustard chemical warfare agent by the C24, C12Si12, Al12N12, Al12P12, Be12O12, B12N12 and Mg12O12 nanocages. J. Mol. Struct. 2018, 1164, 227–238. [Google Scholar] [CrossRef]
  24. Duraisamy, P.D.; Paul, S.P.M.; Gopalan, P.; Angamuthu, A. Feasibility of halide (F−, Cl− and Br−) encapsulated Be12O12 nanocages as potential anode for metal-ion batteries—A DFT-D3 approach. Mater. Sci. Semicond. Process. 2022, 147, 106719. [Google Scholar] [CrossRef]
  25. Ahangari, M.G.; Mashhadzadeh, A.H. Dynamics study on hydrogen storage capacity of C24, B12N12, Al12N12, Be12O12, Mg12O12, and Zn12O12 nanocages. Int. J. Hydrog. Energy 2020, 45, 6745–6756. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Qin, C.; Wang, Y.; Cao, J. Highly sensitive sensors based on g-C3N4/In2O3 heterostructure for rapid detection of formaldehyde at low temperature. Diam. Relat. Mater. 2023, 135, 109828. [Google Scholar] [CrossRef]
  27. Hwang, D.-Y.; Mebel, A.M. Theoretical study of the reaction of beryllium oxide with methane. Chem. Phys. Lett. 2001, 348, 303–310. [Google Scholar] [CrossRef]
  28. Rezaei-Sameti, M.; Abdoli, S. The capability of the pristine and (Sc, Ti) doped Be12O12 nanocluster to detect and adsorb of Mercaptopyridine molecule: A first principle study. J. Mol. Struct. 2020, 1205, 127593. [Google Scholar] [CrossRef]
  29. Badran, H.M.; Eid, K.M.; Baskoutas, S.; Ammar, H.Y. Mg12O12 and Be12O12 Nanocages as Sorbents and Sensors for H2S and SO2 Gases: A Theoretical Approach. Nanomaterials 2022, 12, 1757. [Google Scholar] [CrossRef]
  30. Kosar, N.; Gul, S.; Ayub, K.; Bahader, A.; Gilani, M.A.; Arshad, M.; Mahmood, T. Significant nonlinear optical response of alkaline earth metals doped beryllium and magnesium oxide nanocages. Mater. Chem. Phys. 2019, 242, 122507. [Google Scholar] [CrossRef]
  31. Lorrain, J.M.; Fortune, C.R.; Dellinger, B. Sampling and ion chromatographic determination of formaldehyde and acetaldehyde. Anal. Chem. 1981, 53, 1302–1305. [Google Scholar] [CrossRef]
  32. Mann, B.; Grayeski, M. New chemiluminescent derivatizing agent for the analysis of aldehydes and ketones by high-performance liquid chromatography with peroxyoxalate chemiluminescence. J. Chromatogr. A 1986, 386, 149–158. [Google Scholar] [CrossRef]
  33. Xu, Y.-J.; Zhang, Y.-F.; Lu, N.-X.; Li, J.-Q. Cluster models study of CH2O adsorption and dissociation at defect sites of MgO (001) surface. Phys. B 2004, 348, 190–197. [Google Scholar] [CrossRef]
  34. Noorizadeh, S.; Shakerzadeh, E. Formaldehyde adsorption on pristine, Al-doped and mono-vacancy defected boron nitride nanosheets: A first principles study. Comput. Mater. Sci. 2012, 56, 122–130. [Google Scholar] [CrossRef]
  35. Shakerzadeh, E. A DFT study on the formaldehyde (H2CO and (H2CO)2) monitoring using pristine B12N12 nanocluster. Phys. E 2016, 78, 1–9. [Google Scholar] [CrossRef]
  36. Cuba-Supanta, G.; Guerrero-Sánchez, J.; Rojas-Tapia, J.; Landauro, C.; Takeuchi, N. Formaldehyde trapping by radical initiated reaction on hydrogenated boron nitride. Appl. Surf. Sci. 2019, 484, 470–478. [Google Scholar] [CrossRef]
  37. Wang, R.; Zhu, R.; Zhang, D. Adsorption of formaldehyde molecule on the pristine and silicon-doped boron nitride nanotubes. Chem. Phys. Lett. 2008, 467, 131–135. [Google Scholar] [CrossRef]
  38. Abbasi, A.; Sardroodi, J.J.; Ebrahimzadeh, A.R. Chemisorption of CH2O on N-doped TiO2 anatase nanoparticle as modified nanostructure mediums: A DFT study. Surf. Sci. 2016, 654, 20–32. [Google Scholar] [CrossRef]
  39. Li, X.; Zhang, N.; Liu, C.; Adimi, S.; Zhou, J.; Liu, D.; Ruan, S. Enhanced gas sensing properties for formaldehyde based on ZnO/Zn2SnO4 composites from one-step hydrothermal synthesis. J. Alloys Compd. 2020, 850, 156606. [Google Scholar] [CrossRef]
  40. Deng, X.; Liang, X.; Ng, S.-P.; Wu, C.-M.L. Adsorption of formaldehyde on transition metal doped monolayer MoS2: A DFT study. Appl. Surf. Sci. 2019, 484, 1244–1252. [Google Scholar] [CrossRef]
  41. Shukla, A.; Gaur, N. Adsorption of O3, SO3 and CH2O on two dimensional SnS monolayer: A first principles study. Phys. B Condens. Matter 2019, 572, 12–17. [Google Scholar] [CrossRef]
  42. Wang, X.; Zhang, J.; Wang, L.; Li, S.; Liu, L.; Su, C.; Liu, L. High Response Gas Sensors for Formaldehyde Based on Er-doped In2O3 Nanotubes. J. Mater. Sci. Technol. 2015, 31, 1175–1180. [Google Scholar] [CrossRef]
  43. Hu, J.; Wang, T.; Wang, Y.; Huang, D.; He, G.; Han, Y.; Hu, N.; Su, Y.; Zhou, Z.; Zhang, Y.; et al. Enhanced formaldehyde detection based on Ni doping of SnO 2 nanoparticles by one-step synthesis. Sens. Actuators B Chem. 2018, 263, 120–128. [Google Scholar] [CrossRef]
  44. Dey, S.; Santra, S.; Sen, S.; Burman, D.; Ray, S.K.; Guha, P.K. Photon-Assisted Ultra-Selective Formaldehyde Sensing by Defect Induced NiO-Based Resistive Sensor. IEEE Sens. J. 2018, 18, 5656–5661. [Google Scholar] [CrossRef]
  45. Mohsennia, M.; Rakhshi, M.; Rasa, H. A computational study on interactions of Ni- and Pt-doped boron nitride nano tubes with NH3 in presence and absence of electric fields. Comput. Theor. Chem. 2018, 1136–1137, 1–9. [Google Scholar] [CrossRef]
  46. Rakhshi, M.; Mohsennia, M.; Rasa, H.; Sameti, M.R. First-principle study of ammonia molecules adsorption on boron nitride nanotubes in presence and absence of static electric field and ion field. Vacuum 2018, 155, 456–464. [Google Scholar] [CrossRef]
  47. Farmanzadeh, D.; Ghazanfary, S. Single-wall Boron Nitride Nanotube with Various Diameters under the Influence of Electric Field, A Computational Investigation. J. Chin. Chem. Soc. 2012, 59, 961–966. [Google Scholar] [CrossRef]
  48. Behzad, S.; Chegel, R. Thermal conductivity, heat capacity and magnetic susceptibility of graphene and boron nitride nanoribbons. Diam. Relat. Mater. 2018, 88, 101–109. [Google Scholar] [CrossRef]
  49. Zheng, F.; Liu, Z.; Wu, J.; Duan, W.; Gu, B.-L. Scaling law of the giant Stark effect in boron nitride nanoribbons and nanotubes. Phys. Rev. B 2008, 78, 085423. [Google Scholar] [CrossRef]
  50. Park, C.-H.; Louie, S.G. Energy Gaps and Stark Effect in Boron Nitride Nanoribbons. Nano Lett. 2008, 8, 2200–2203. [Google Scholar] [CrossRef]
  51. Chegel, R. Effects of carbon doping on the electronic properties of boron nitride nanotubes: Tight binding calculation. Phys. E 2016, 84, 223–234. [Google Scholar] [CrossRef]
  52. Chegel, R.; Behzad, S. Electro-optical properties of zigzag and armchair boron nitride nanotubes under a transverse electric field: Tight binding calculations. J. Phys. Chem. Solids 2012, 73, 154–161. [Google Scholar] [CrossRef]
  53. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef]
  54. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  55. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787. [Google Scholar] [CrossRef] [PubMed]
  56. 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] [PubMed]
  57. Ammar, H.; Eid, K.; Badran, H. TM-doped Mg12O12 nano-cages for hydrogen storage applications: Theoretical study. Results Phys. 2022, 35, 105349. [Google Scholar] [CrossRef]
  58. 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]
  59. Pearson, R.G. Absolute electronegativity and hardness correlated with molecular orbital theory. Proc. Natl. Acad. Sci. USA 1986, 2683, 8440–8441. [Google Scholar] [CrossRef]
  60. Parr, R.G.; Szentpály, L.V.; Liu, S. Electrophilicity Index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  61. 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]
  62. O’Boyle, N.M.; Tenderholt, A.L.; Langner, K.M. cclib: A library for package-independent computational chemistry algorithms. J. Comput. Chem. 2008, 29, 839–845. [Google Scholar] [CrossRef] [PubMed]
  63. 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]
  64. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  65. Shakerzdeh, E.; Tahmasebi, E.; Shamlouei, H.R. The influence of alkali metals (Li, Na and K) interaction with Be12O12 and Mg12O12 nanoclusters on their structural, electronic and nonlinear optical properties: A theoretical study. Synth. Met. 2015, 204, 17–24. [Google Scholar] [CrossRef]
  66. 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]
  67. Balan, C.; Pop, L.-C.; Baia, M. IR, Raman and SERS analysis of amikacin combined with DFT-based calculations. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 214, 79–85. [Google Scholar] [CrossRef]
  68. Yüksel, B.Ş. Spectroscopic characterization (IR and NMR), structural investigation, DFT study, and Hirshfeld surface analysis of two zinc(II) 2-acetylthiophenyl-thiosemicarbazone complexes. J. Mol. Struct. 2020, 1229, 129617. [Google Scholar] [CrossRef]
  69. Badran, H.; Eid, K.; Ammar, H. 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]
  70. 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]
  71. Yang, M.; Zhang, Y.; Huang, S.; Liu, H.; Wang, P.; Tian, H. Theoretical investigation of CO adsorption on TM-doped (MgO)12 (TM = Ni, Pd, Pt) nanotubes. Appl. Surf. Sci. 2011, 258, 1429–1436. [Google Scholar] [CrossRef]
  72. 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]
  73. Kittel, C. Introduction to Solid State Physics; John Wiley and Sons: Hoboken, NJ, USA, 2005; pp. 216–226. [Google Scholar]
  74. Kasap, S.O. Principles of Electronic Materials and Devices; McGraw-Hill: Boston, MA, USA, 2006; pp. 378–405. [Google Scholar]
  75. 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]
  76. 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]
  77. 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] [PubMed]
  78. 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]
  79. Kadhim, M.M.; Sadoon, N.; Gheni, H.A.; Hachim, S.K.; Majdi, A.; Abdullaha, S.A.; Rheima, A.M. Application of B3O3 monolayer as an electrical sensor for detection of formaldehyde gas: A DFT study. Comput. Theor. Chem. 2023, 1219, 113941. [Google Scholar] [CrossRef]
  80. Kumar, A.; Al-Bahrani, M.; Hossain, A.; Mehedi, I.M.; Iskanderani, A.I.; Gavilán, J.C.O.; Ramaiah, G.B. Transition metal-functionalized porphyrin-like C70 fullerenes as sensors and adsorbents of formaldehyde- DFT, NBO, and QTAIM study. Inorg. Chem. Commun. 2023, 153, 110883. [Google Scholar] [CrossRef]
  81. Yoosefian, M.; Raissi, H.; Mola, A. The hybrid of Pd and SWCNT (Pd loaded on SWCNT) as an efficient sensor for the formaldehyde molecule detection: A DFT study. Sens. Actuators B Chem. 2015, 212, 55–62. [Google Scholar] [CrossRef]
  82. Rastegar, S.F.; Peyghan, A.A.; Soleymanabadi, H. Ab initio studies of the interaction of formaldehyde with beryllium oxide nanotube. Phys. E 2015, 68, 22–27. [Google Scholar] [CrossRef]
  83. 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]
  84. Celik, S.; Tanıs, 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]
  85. Srivastava, A.K. DFT and QTAIM studies on the reduction of carbon monoxide by superalkalis. J. Mol. Graph. Model. 2020, 102, 107765. [Google Scholar] [CrossRef] [PubMed]
  86. Demir, S.; Fellah, M.F. A DFT study on Pt doped (4,0) SWCNT: CO adsorption and sensing. Appl. Surf. Sci. 2020, 504, 144141. [Google Scholar] [CrossRef]
  87. Cui, H.; Zhang, G.; Zhang, X.; Tang, J. Rh-doped MoSe2 as a toxic gas scavenger: A first-principles study. Nanoscale Adv. 2019, 1, 772–780. [Google Scholar] [CrossRef] [PubMed]
  88. Soliman, I.; El-Nahass, M.; Eid, K.; Ammar, H. Vibrational spectroscopic analysis of aluminum phthalocyanine chloride. experimental and DFT study. Phys. B 2016, 491, 98–103. [Google Scholar] [CrossRef]
Figure 1. Optimized structures for (a) Be12O12, (b) KBe12O12, (c) MnBe12O12, and (d) CuBe12O12 nano-cages.
Figure 1. Optimized structures for (a) Be12O12, (b) KBe12O12, (c) MnBe12O12, and (d) CuBe12O12 nano-cages.
Nanomaterials 14 00007 g001
Figure 2. Charge density difference ( Δ ρ ) at 0.001 au isovalue for (a) KBe12O12 (b) MnBe12O12 (c) CuBe12O12 nano-cages. Red and blue colors refer to negative and positive Δ ρ values; Δ ρ = ρ M B e 12 O 12 ( ρ B e 12 O 12 + ρ M ) .
Figure 2. Charge density difference ( Δ ρ ) at 0.001 au isovalue for (a) KBe12O12 (b) MnBe12O12 (c) CuBe12O12 nano-cages. Red and blue colors refer to negative and positive Δ ρ values; Δ ρ = ρ M B e 12 O 12 ( ρ B e 12 O 12 + ρ M ) .
Nanomaterials 14 00007 g002
Figure 3. The molecular electrostatic potential maps (MESP) for (a) Be12O12, (b) KBe12O12, (c) MnBe12O12, and (d) CuBe12O12 nano-cages.
Figure 3. The molecular electrostatic potential maps (MESP) for (a) Be12O12, (b) KBe12O12, (c) MnBe12O12, and (d) CuBe12O12 nano-cages.
Nanomaterials 14 00007 g003
Figure 4. HOMO, PDOS, and LUMO for (a) Be12O12, (b) KBe12O12, (c) MnBe12O12, and (d) CuBe12O12 nano-cages. HOMO and LUMO surfaces are plotted at ± 0.02 isovalue. Brown and green colors refer to positive and negative isovalues, respectively.
Figure 4. HOMO, PDOS, and LUMO for (a) Be12O12, (b) KBe12O12, (c) MnBe12O12, and (d) CuBe12O12 nano-cages. HOMO and LUMO surfaces are plotted at ± 0.02 isovalue. Brown and green colors refer to positive and negative isovalues, respectively.
Nanomaterials 14 00007 g004
Figure 5. UV-vis spectra for Be12O12, KBe12O12, MnBe12O12, and CuBe12O12 nano-cages.
Figure 5. UV-vis spectra for Be12O12, KBe12O12, MnBe12O12, and CuBe12O12 nano-cages.
Nanomaterials 14 00007 g005
Figure 6. Optimized structures for (a) CH2O/Be12O12, (b) CH2O/KBe12O12, (c) CH2O/MnBe12O12, and (d) CH2O/CuBe12O12.
Figure 6. Optimized structures for (a) CH2O/Be12O12, (b) CH2O/KBe12O12, (c) CH2O/MnBe12O12, and (d) CH2O/CuBe12O12.
Nanomaterials 14 00007 g006
Figure 7. Charge density difference ( Δ ρ ) at 0.001 au isovalue for (a) CH2O/Be12O12, (b) CH2O/KBe12O12, (c) CH2O/MnBe12O12, and (d) CH2O/CuBe12O12 complexes. Red and blue colors refer to negative and positive Δ ρ values; Δ ρ = ρ C H 2 O / M B e 12 O 12 ( ρ M B e 12 O 12 + ρ C H 2 O ) .
Figure 7. Charge density difference ( Δ ρ ) at 0.001 au isovalue for (a) CH2O/Be12O12, (b) CH2O/KBe12O12, (c) CH2O/MnBe12O12, and (d) CH2O/CuBe12O12 complexes. Red and blue colors refer to negative and positive Δ ρ values; Δ ρ = ρ C H 2 O / M B e 12 O 12 ( ρ M B e 12 O 12 + ρ C H 2 O ) .
Nanomaterials 14 00007 g007
Figure 8. HOMO, PDOS, and LUMO for (a) CH2O, (b) CH2O/Be12O12, (c) CH2O/KBe12O12, (d) CH2O/MnBe12O12, and (e) CH2O/CuBe12O12 complexes.
Figure 8. HOMO, PDOS, and LUMO for (a) CH2O, (b) CH2O/Be12O12, (c) CH2O/KBe12O12, (d) CH2O/MnBe12O12, and (e) CH2O/CuBe12O12 complexes.
Nanomaterials 14 00007 g008
Figure 9. (a) Electric field direction relative to the nano-cage, and the charge density difference ( Δ ρ ) isovalue surfaces at 0.00005 au for EF values of −514 and +514 kV/mm for (b) CH2O, (c) Be12O12, (d) KBe12O12, (e) MnBe12O12, and (f) KBe12O12. Δ ρ = ρ   E = ± 514 ρ   E = 0 , X-component of dipole moment (Dx) in Debye. Red and blue colors represent negative and positive Δ ρ values, respectively.
Figure 9. (a) Electric field direction relative to the nano-cage, and the charge density difference ( Δ ρ ) isovalue surfaces at 0.00005 au for EF values of −514 and +514 kV/mm for (b) CH2O, (c) Be12O12, (d) KBe12O12, (e) MnBe12O12, and (f) KBe12O12. Δ ρ = ρ   E = ± 514 ρ   E = 0 , X-component of dipole moment (Dx) in Debye. Red and blue colors represent negative and positive Δ ρ values, respectively.
Nanomaterials 14 00007 g009
Figure 10. Dipole moment vs. electric field at different mediums for (a) CH2O, (b) Be12O12 and MBe12O12 substrates, and (c) CH2O/Be12O12 and CH2O/MBe12O12 complexes.
Figure 10. Dipole moment vs. electric field at different mediums for (a) CH2O, (b) Be12O12 and MBe12O12 substrates, and (c) CH2O/Be12O12 and CH2O/MBe12O12 complexes.
Nanomaterials 14 00007 g010
Figure 11. Adsorption energies (Eads) for CH2O/Be12O12 and CH2O/MBe12O12 complexes.
Figure 11. Adsorption energies (Eads) for CH2O/Be12O12 and CH2O/MBe12O12 complexes.
Nanomaterials 14 00007 g011
Figure 12. NBO charges of CH2O molecule ( Q C H 2 O ) vs. electric field for CH2O/Be12O12 and CH2O/MBe12O12 complexes.
Figure 12. NBO charges of CH2O molecule ( Q C H 2 O ) vs. electric field for CH2O/Be12O12 and CH2O/MBe12O12 complexes.
Nanomaterials 14 00007 g012
Figure 13. HOMO–LUMO energy gap (Eg) vs. electric field for (a) Be12O12 and CH2O/Be12O12, (b) KBe12O12 and KCH2O/Be12O12, (c) MnBe12O12 and CH2O/MnBe12O12, and (d) CuBe12O12 and CH2O/CuBe12O12.
Figure 13. HOMO–LUMO energy gap (Eg) vs. electric field for (a) Be12O12 and CH2O/Be12O12, (b) KBe12O12 and KCH2O/Be12O12, (c) MnBe12O12 and CH2O/MnBe12O12, and (d) CuBe12O12 and CH2O/CuBe12O12.
Nanomaterials 14 00007 g013
Figure 14. The change percentage for HOMO–LUMO energy gap (ΔEg) vs. electric field.
Figure 14. The change percentage for HOMO–LUMO energy gap (ΔEg) vs. electric field.
Nanomaterials 14 00007 g014
Figure 15. UV-vis spectra for CH2O/Be12O12, CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12 complexes for EF values of (a) −514 kV/mm, (b) 0 kV/mm, and (c) +514 kV/mm, respectively.
Figure 15. UV-vis spectra for CH2O/Be12O12, CH2O/KBe12O12, CH2O/MnBe12O12, and CH2O/CuBe12O12 complexes for EF values of (a) −514 kV/mm, (b) 0 kV/mm, and (c) +514 kV/mm, respectively.
Nanomaterials 14 00007 g015
Figure 16. (a) E ¯ ads and (b) Eg against the number of adsorbed CH2O molecules.
Figure 16. (a) E ¯ ads and (b) Eg against the number of adsorbed CH2O molecules.
Nanomaterials 14 00007 g016
Table 1. Electronic properties of Be12O12 and MBe12O12 nano-cages (M = K, Mn, or Cu). HOMO and LUMO energy levels (eV), HOMO–LUMO gap (Eg, eV), average binding energy per atom (Eb, eV), NBO charges (Q, e), ionization potential (IP, eV), Fermi level (EF, eV), hardness (η, eV), electrophilicity (ω, eV), and dipole moment (D, Debye).
Table 1. Electronic properties of Be12O12 and MBe12O12 nano-cages (M = K, Mn, or Cu). HOMO and LUMO energy levels (eV), HOMO–LUMO gap (Eg, eV), average binding energy per atom (Eb, eV), NBO charges (Q, e), ionization potential (IP, eV), Fermi level (EF, eV), hardness (η, eV), electrophilicity (ω, eV), and dipole moment (D, Debye).
Be12O12KBe12O12MnBe12O12CuBe12O12
HOMO (α)−8.463−2.128−2.281−3.544
LUMO (α)−0.253−0.973−0.547−0.373
HOMO (β)−8.463−8.607−3.803−7.446
LUMO (β)−0.253−1.443−1.224−1.977
Eg8.2100.6851.0571.568
Eb−5.630−5.137−5.201−5.405
QBe1.2101.2411.1991.185
QO−1.210−1.249−1.211−1.194
QM-0.0960.2370.108
IP10.1353.2983.6395.216
EF−4.358−1.786−1.753−2.760
η4.1050.3430.5290.784
ω2.3144.6552.9054.860
D0.0010.0030.0010.105
Table 2. The electronic configurations for M atom in a free state and MBe12O12 nano-cages.
Table 2. The electronic configurations for M atom in a free state and MBe12O12 nano-cages.
Structure4s3d4p
K1.00--
KBe12O120.460.230.41
Mn2.005.00-
MnBe12O120.635.760.41
Cu1.0010.0-
CuBe12O120.769.670.48
Table 3. Adsorption properties of CH2O on MBe12O12 nano-cages. Adsorption energies (Eads, eV), HOMO and LUMO energy levels (eV), HOMO–LUMO gap (Eg, eV), NBO charges (Q, e), and dipole moment (D, Debye).
Table 3. Adsorption properties of CH2O on MBe12O12 nano-cages. Adsorption energies (Eads, eV), HOMO and LUMO energy levels (eV), HOMO–LUMO gap (Eg, eV), NBO charges (Q, e), and dipole moment (D, Debye).
CH2O/Be12O12CH2O/KBe12O12CH2O/MnBe12O12CH2O/CuBe12O12
Eads−0.963−2.518−2.551−1.541
HOMO−7.869−2.665−3.007−2.900
LUMO−2.886−1.380−1.491−2.024
Eg4.9831.2861.5160.876
QBe11.1341.2301.2081.191
QO2−1.221−1.233−1.181−1.177
QO3−1.210−1.232−1.191−1.182
QO4−1.210−1.229−1.174−1.181
QM-0.4430.5700.578
QO1−0.516−0.858−0.853−0.849
QH10.1940.1340.1340.134
QH20.1800.1480.1530.151
QC0.290−0.128−0.124−0.126
Q C H 2 O 0.148−0.703−0.690−0.690
D5.3878.4946.8046.962
Table 4. The electronic configurations for 2s and 2p orbitals for oxygen and carbon atoms of CH2O for free CH2O, CH2O/Be12O12, and CH2O/MBe12O12.
Table 4. The electronic configurations for 2s and 2p orbitals for oxygen and carbon atoms of CH2O for free CH2O, CH2O/Be12O12, and CH2O/MBe12O12.
StructureO C
2s2p2s2p
CH2O1.724.761.052.70
CH2O/Be12O121.664.841.072.62
CH2O/KBe12O121.665.181.073.04
CH2O/MnBe12O121.665.181.073.03
CH2O/CuBe12O121.665.181.073.03
Table 5. The estimated topological parameters. Electron densities ( ρ ), Laplacian of charge density ( 2 ρ ), kinetic electron density (G(r)), potential energy density (V(r)), and energy density (H(r)). All units are in au.
Table 5. The estimated topological parameters. Electron densities ( ρ ), Laplacian of charge density ( 2 ρ ), kinetic electron density (G(r)), potential energy density (V(r)), and energy density (H(r)). All units are in au.
ComplexBCP * ρ 2 ρ G(r)V(r)H(r)−G(r)/V(r)
CH2O/Be12O12Be1-O10.0480.3270.077−0.0720.0051.069
O2-H10.0140.0540.012−0.0100.0021.200
CH2O/KBe12O12Be1-O10.1100.8580.212−0.211−0.0021.005
CH2O/MnBe12O12Be1-O10.1130.8790.219−0.219−0.0011.000
CH2O/CuBe12O12Be1-O10.1120.8670.216−0.215−0.0011.005
* The atom numbering shown in the Figure 6 is used.
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

Al-Nadary, H.O.; Eid, K.M.; Badran, H.M.; Ammar, H.Y. M-Encapsulated Be12O12 Nano-Cage (M = K, Mn, or Cu) for CH2O Sensing Applications: A Theoretical Study. Nanomaterials 2024, 14, 7. https://doi.org/10.3390/nano14010007

AMA Style

Al-Nadary HO, Eid KM, Badran HM, Ammar HY. M-Encapsulated Be12O12 Nano-Cage (M = K, Mn, or Cu) for CH2O Sensing Applications: A Theoretical Study. Nanomaterials. 2024; 14(1):7. https://doi.org/10.3390/nano14010007

Chicago/Turabian Style

Al-Nadary, Hatim Omar, Khaled Mahmoud Eid, Heba Mohamed Badran, and Hussein Youssef Ammar. 2024. "M-Encapsulated Be12O12 Nano-Cage (M = K, Mn, or Cu) for CH2O Sensing Applications: A Theoretical Study" Nanomaterials 14, no. 1: 7. https://doi.org/10.3390/nano14010007

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop