Azam Moumivanda,
Fereshteh Naderi*a,
Omid Moradi
a and
Batoul Makiabadib
aDepartment of Chemistry, Shahr-e-Qods Branch, Islamic Azad University, Tehran, Iran. E-mail: fnaderi1@gmail.com; Neginmmm64@gmail.com; moradi.omid@gmail.com; Web: https://scholar.google.com/citations?user=pdCrLOYAAAAJ&hl=en
bDepartment of Chemical Engineering, Sirjan University of Technology, Sirjan, Iran. E-mail: bmakiabadi@yahoo.com; Web: https://scholar.google.com/citations?user=N6z-rHsAAAAJ&hl=en
First published on 17th December 2024
The potential applicability of the C24 nanocage and its boron nitride-doped analogs (C18B3N3 and C12B6N6) as pyrazinamide (PA) carriers was investigated using density functional theory. Geometry optimization and energy calculations were performed using the B3LYP functional and 6-31G(d) basis set. Besides, dispersion-corrected interaction energies were calculated at CAM (Coulomb attenuated method)-B3LYP/6-31G(d,p) and M06-2X/6-31G(d,p) levels of theory. The adsorption energy (Eads), enthalpy (ΔH), and Gibbs free energy (ΔG) values for C24-PA, C18B3N3-PA, and C12B6N6-PA structures were calculated. The molecular descriptors such as electrophilicity (ω), chemical potential (μ), chemical hardness (η) and chemical softness (S) of compounds were investigated. Natural bond orbital (NBO) analysis confirms the charge transfer from the drug molecule to nanocarriers upon adsorption. Based on the quantum theory of atoms in molecules (QTAIM), the nature of interactions in the complexes was determined. These findings suggest that C24 and its doped analogs are promising candidates for smart drug delivery systems and PA sensing applications, offering significant potential for advancements in targeted tuberculosis treatment.
One of the four first-line anti-tuberculosis drugs is pyrazinamide (PA), which, in addition to shortening the treatment period from nine to six months, eradicates the organism by destroying the active and inactive forms of bacillus tuberculosis in the acidic environment inside the macrophages.6 Like other drugs, PA has important side effects, such as fever, anorexia, liver enlargement, jaundice, and liver failure. Targeted drug delivery is a new way to reduce problems caused by direct drug use and drug resistance.7–9 Considering the common challenges of using drugs such as PA, the controlled release of the drug has attracted the attention of researchers. Targeted drug delivery and controlled release has advantages such as the possibility of drug delivery in nanometer dimensions, the ability to maintain the drug concentration in a relatively constant amount for a certain time, the ability to adjust the drug release rate depending on the place of drug delivery, and the ability to transfer several drug substances to a specific tissue or organ.10,11 Nowadays, drug delivery based on nanocarriers is at the center of attention of many companies acting in the pharmaceutical industry since nanomaterials have special properties and potential applications that make them superior to the traditional drug delivery vectors. The use of nanocarriers in medical science helps to increase the availability of the drug at the site of disease, and reduces the dose frequency and adverse side effects.12–15 As new drugs are developed, significant efforts have been directed toward exploring innovative delivery routes. Targeted delivery systems represent a promising solution to overcome the limitations of conventional methods. To realize this potential, a wide range of nanoparticles have been proposed and extensively studied.16–20 These nanoparticles must interact well with biological environments and pass through the cell membrane to deliver therapeutic molecules. Fullerenes (C60) are one of the pioneering classes of carbon-based nanoparticles, which have been widely studied for targeted drug delivery applications.21 Fullerenes have a unique structure and have suitable properties for interaction with drugs. Although fullerene toxicity is somewhat of a concern, several water-soluble fullerene derivatives have shown acceptable cytotoxicity for drug delivery applications.22 Recently, new fullerene-like compounds and hexagonal boron nitrides have been developed, and their potential applications as advanced delivery systems have been investigated.23–27 Studies provide a detailed drug interaction mechanism with boron nitride fullerenes (BNF) and reveal that the BNF can be a smart drug delivery vehicle for the drugs.23,28–30 In this study, the interaction of the pyrazinamide drug with C24, C18B3N3, and C12B6N6 nanocages was investigated using density functional theory (DFT) calculations. To achieve this, we employed a range of analytical tools, including adsorption energy (Eads), thermochemical parameters (ΔH and ΔG), frontier molecular orbitals, ionization energy, electrophilicity, chemical hardness, softness, electronegativity, natural bond orbitals (NBO), and atoms in molecules (AIM), as well as NMR, IR, and UV-Vis spectroscopy. These comprehensive analyses provide a detailed understanding of the interaction mechanisms and potential applications of these nanocages in drug delivery and sensing systems.
Eads = Ecomplex − (Enanocage + EPA) | (1) |
Natural bond orbital analysis provides a chemical viewpoint of van der Waals interactions using a donor–acceptor type interpretation of molecular orbitals.38 Therefore, we used NBO analyses to calculate the charge distribution over the atoms of interacting species and to estimate the stabilization energies (E(2)) of bond–antibond interactions. The energy difference between the highest occupied (HOMO) and the lowest unoccupied molecular orbital (LUMO), known as Eg, for a system is given as:
Eg = ELUMO − EHOMO | (2) |
The electronic sensitivity of the pristine and doped C24 nanocage toward PA is calculated as follows:
![]() | (3) |
Quantum molecular descriptors (also known as the conceptual density functional theory (CDFT) parameters), including the electronegativity (χ), global hardness (η), electrophilicity index (ω), ionization potential (I), electron affinity (A), and electronic chemical potential (μ), were calculated with the help of HOMO and LUMO values. The electrophilicity index (ω) measures the stability of a compound in the presence of an additional electronic charge from the environment. In a charge transfer process, higher values of ω indicate the higher electrophilic power of the structure.39,40
μ = −χ = −(EHOMO + ELUMO)/2 | (4) |
η = (ELUMO − EHOMO)/2 | (5) |
ω = μ2/2η | (6) |
The ionization energy (I) and electron affinity (A) can be expressed as −EHOMO and −ELUMO, respectively. Chemical shielding (CS) calculations were performed using the gauge-including atomic orbital (GIAO) method.41 The chemical shift isotropy (σiso) and anisotropy (Δσ) parameters were calculated applying principal components (σ11 ≤ σ22 ≤ σ33) of the chemical shift tensor, as follows:
σiso (ppm) = (σ11 + σ22 + σ33)/3 | (7) |
Δσ (ppm) = σ33 − (σ11 + σ22)/2 | (8) |
The NBO and AIM analysis was carried out at the M06-2X/6-31G(d) level of theory.38,42
![]() | ||
Fig. 1 (a) Optimized geometries, (b) HOMOs, (c) LUMOs, and (d) MEP plots for C24, C18B3N3, and C12B6N6 nanocages at the M06-2X/6-31G(d,p) level of theory. |
The molecular electrostatic potential (MEP) plot is a valuable tool for determining the electron density distribution over the atoms of a molecule and predicting its most reactive sites. It is widely known that the chemical and physical properties of a molecule are associated with its electrostatic potential. Therefore, the MEP plot can be utilized to characterize electrophilic and nucleophilic sites within an electrostatic interaction.45 In a MEP contour, the surfaces are defined based on electron density and represented by an RGB color model, in which red regions are more negative than −0.010 a.u., yellow shows the regions with electron density between 0 and −0.010 a.u., green regions are between 0.010 and 0.0 a.u., and blue color represents regions more positive than 0.010 a.u. Fig. 2 reveals that O13, N8, and N10 atoms are the most negatively charged centers of the PA molecule. Additionally, the oxygen atom is more prone to contribute to the nucleophilic attack. The most positive atoms are C12 (0.68 |e|), H9 (0.45 |e|), H16 (0.43 |e|), H11 (0.41 |e|), and H15 (0.42 |e|), respectively, as shown by NBO analysis, which is in agreement with the MEP plot of the PA molecule.
![]() | ||
Fig. 2 (a) Optimized geometry, (b) MEP, (c) HOMO, and (d) LUMO profiles of PA at the M06-2X/6-31G(d,p) level of theory. |
The frequency calculations were performed at the M06-2X/6-31G(d,p) level of theory. The vibrational frequency modes of C24, C18B3N3, and C12B6N6 appear in a positive range of 343–1630 cm−1, 345–1603 cm−1, and 282–1473 cm−1, respectively, indicating that the optimized geometries are true local minima on the potential energy surface. Frontier molecular orbitals (HOMO and LUMO) of a molecule play a vital role in determining its reactivity. HOMO and LUMO energy levels are usually considered as the electron-donating and electron-accepting abilities of an organic molecule. A higher EHOMO value indicates that the molecule can donate electrons, while a lower ELUMO implies the molecule's willingness to accept electrons. Considering eqn (2), it can be concluded that molecules with a low HOMO–LUMO energy gap are more polarizable and have high chemical reactivity and low kinetic stability.46 The HOMO levels of C24 are distributed over the C–C bonds throughout the nanocage except some C–C bonds in pentagons, whereas the LUMO levels are located on the C–C bonds of 5-membered rings. In the case of C18B3N3 and C12B6N6, the HOMO levels are mainly located on the N atoms and C–C bonds of pentagons, while the LUMO levels are situated on the B atoms and C–C bonds. The HOMO–LUMO energy gaps of C24, C18B3N3, and C12B6N6 are about 3.92, 4.24, and 4.86 eV, respectively, indicating that these nanocages are semiconductors. In the following sections, we will discuss the role of frontier molecular orbitals in the electronic properties of molecules.
Method | Eads (kJ mol−1) | Ebsseads (kJ mol−1) | ΔH (kJ mol−1) | ΔG (kJ mol−1) |
---|---|---|---|---|
B3LYP/6-31G(d) | ||||
C24-PA | −13.88 | −1.34 | −9.86 | 27.48 |
C18B3N3-PA | −25.72 | −5.75 | −22.16 | 21.98 |
C12B6N6-PA | −125.58 | −98.05 | −120.99 | −70.88 |
CAM-B3LYP/6-31G(d,p) | ||||
C24-PA | −14.81 | −8.34 | −9.57 | −35.40 |
C18B3N3-PA | −124.73 | −97.78 | −119.43 | −68.50 |
C12B6N6-PA | −138.55 | −124.21 | −309.76 | −83.07 |
![]() |
||||
M06-2X/6-31G(d,p) | ||||
C24-PA | −45.95 | −29.98 | −40.08 | −3.62 |
C18B3N3-PA | −163.63 | −56.62 | −155.81 | −95.25 |
C12B6N6-PA | −156.48 | −137.69 | −150.71 | −99.17 |
![]() | ||
Fig. 3 (a) Optimized geometry, (b) MEP, (c) HOMO, and (d) LUMO profiles of C24-PA at the M06-2X/6-31G(d,p) level of theory. |
The values of enthalpy (ΔH) and Gibbs free energy changes (ΔG) of the C24-PA complex are negative (with the exception of the B3LYP/6-31G(d) level), indicating that an exothermic and spontaneous reaction has occurred.
The changes in bond length and bond angles of the drug molecule, upon interaction with C24 at the M06-2X/6-31G(d,p) level of theory are illustrated in Tables 2 and S1,† respectively. The length of the C1–C2, C4–N8, C1–N8, N8–H9, and C12–O13 bonds increases while the length of other bonds decreases upon complexation. The results show that the interaction between the drug molecule and C12B6N6 has no significant effect on the bond lengths of the PA molecule. The most obvious change in the bond angles of the PA molecule is related to the C12–N14–H16 angle, which ranges from 114.5° to 119.1°. Table 3 summarizes the quantum chemistry reactivity parameters of the nonionic surfactants, including EHOMO, ELUMO, and their associated energy gap (Eg), ionization potential (I), electron affinity (A), electronegativity (χ), electronic chemical potential (μ), chemical hardness (η), electrophilicity index (ω), and dipole moments (in Debye). Frontier molecular orbital (FMO) analysis implies that the EHOMO and ELUMO of C24 become more stable by about 1.15 and 0.29 eV, respectively, upon complex formation (Fig. 4). Therefore, the Eg value decreases by 21.9%; thus, the electronic properties of C24 change considerably in the presence of PA molecules. The reduction of parameters ω, χ, A, and I is in agreement with the reduction of Eg. It is also observed that with the decrease of the Eg value, the softness and chemical hardness have increased and decreased, respectively, which indicates the tendency of the nanocarrier to interact with PA. The decrease in electrophilicity by 0.42 eV after interaction with the PA molecule confirms the charge transfer from the drug to the nanocarrier. The dipole moment of the C24-PA complex is significantly higher than that of its components. Therefore, the solubility of the C24-PA complex is expected to be higher than the solubility of C24 and PA.
Bond | Bond length (Å) | Δr | |
---|---|---|---|
After interaction | Before interaction | ||
C1–C2 | 1.343 | 1.341 | 0.002 |
C3–C4 | 1.331 | 1.333 | −0.002 |
C2–H5 | 1.083 | 1.085 | −0.002 |
C3–H6 | 1.081 | 1.082 | −0.001 |
C4–H7 | 1.082 | 1.083 | −0.001 |
C4–N8 | 1.418 | 1.403 | 0.015 |
N8–H9 | 1.017 | 1.011 | 0.006 |
C1–N8 | 1.425 | 1.407 | 0.018 |
C2–N10 | 1.400 | 1.413 | −0.013 |
C3–N10 | 1.423 | 1.434 | −0.011 |
N10–H11 | 1.011 | 1.016 | −0.005 |
C1–C12 | 1.481 | 1.486 | −0.005 |
C12–O13 | 1.223 | 1.221 | 0.002 |
C12–N14 | 1.367 | 1.369 | −0.002 |
N14–H15 | 1.009 | 1.009 | 0.000 |
N14–H16 | 1.007 | 1.007 | 0.000 |
Structure | EHOMO (eV) | ELUMO (eV) | Eg (eV) | EF (eV) | I (eV) | A (eV) | χ (eV) | η (eV) | μ (eV) | S (1/eV) | ω (eV) | μ (Debye) |
---|---|---|---|---|---|---|---|---|---|---|---|---|
PA | −5.68 | 0.31 | 5.99 | 0.15 | 5.68 | −0.31 | 2.69 | 3.00 | −2.69 | 0.17 | 1.20 | 2.73 |
C24 | −6.92 | −3.00 | 3.92 | −1.50 | 6.92 | 3.00 | 4.96 | 1.96 | −4.96 | 0.25 | 6.27 | 0.00 |
C24-PA | −5.77 | −2.70 | 3.06 | −1.35 | 5.77 | 2.70 | 4.24 | 1.53 | −4.24 | 0.33 | 5.86 | 3.79 |
C18B3N3 | −7.04 | −2.80 | 4.24 | −1.40 | 7.04 | 2.80 | 4.92 | 2.12 | −4.92 | 0.24 | 5.70 | 1.10 |
C18B3N3-PA | −5.84 | −1.65 | 4.19 | −0.83 | 5.84 | 1.65 | 3.75 | 2.10 | −3.75 | 0.24 | 3.35 | 11.46 |
C12B6N6 | −7.14 | −2.27 | 4.87 | −1.14 | 7.14 | 2.27 | 4.70 | 2.43 | −4.70 | 0.21 | 4.55 | 2.10 |
C12B6N6-PA | −6.00 | −1.38 | 4.62 | −0.69 | 6.00 | 1.38 | 3.69 | 2.31 | −3.69 | 0.22 | 2.95 | 14.48 |
The mechanism of action of a chemical sensor is directly related to the change in its resistance due to charge exchange with the chemical agent. It is known that the Eg is a dependable indicator to determine the sensitivity of a sensor towards a molecule. The electrical conductivity of a molecule decreases with increasing Eg.47,48 The population of conduction electrons of a nanostructure increases significantly upon the decrease of Eg and vice versa. Changing the population of conducting electrons after the adsorption process causes the generation of an electrical signal.49 The work function (Φ) is another electronic parameter that is affected by the adsorption process and is defined as the minimum amount of energy (i.e., thermodynamic work) required to remove an electron from a solid to a point in the vacuum immediately outside the solid surface (i.e., the Fermi level) and is defined as follows:
Φ = Vel(+∞) − EF | (9) |
EF = EHOMO + (ELUMO − EHOMO)/2 | (10) |
The Φ-type sensors apply a Kelvin oscillator instrument to calculate the values of Φ before and after adsorption.32,50 In these sensors, the adsorption of a molecule changes the gate voltage and produces an electrical signal that leads to chemical agent detection.51 The value of Φ is changed by about −0.147 eV going from C24 to C24-PA; therefore, C24 can be used as a Φ-type sensor for PA detection.
To understand the donor–acceptor (bond–antibond) interactions between the drug molecule and the nanocages, NBO analysis has been performed at the M06-2X/6-31G(d,p) level of theory. In this analysis, all possible interactions between filled (donors) Lewis-type NBOs and empty (acceptors) non-Lewis NBOs are examined. Then, their energetic relevance is evaluated by second order perturbation theory. These interactions are called delocalization corrections to the zeroth-order natural Lewis structure because they result in a donation of electrons from the localized NBOs of the idealized Lewis structure into the empty non-Lewis orbitals. The stabilization energy (E(2)) associated with delocalization i → j is defined as:
![]() | (11) |
![]() | ||
Fig. 5 (a) Optimized geometry, (b) MEP, (c) HOMO, and (d) LUMO profiles of C18B3N3-PA at the M06-2X/6-31G(d,p) level of theory. |
Tables 4 and S3† show the structural parameters (bond lengths and bond angles) of the drug molecule before and after interaction with the C18B3N3 nanostructure. The interaction between the drug molecule and C18B3N3 does not have a significant effect on the length of bond of the PA molecule and is similar to the structure of C24PA. In any case, these changes are very small and can be ignored. The greatest decrease in the bond angle due to PA adsorption on the C18B3N3 surface is related to the C4–N8–H9 angle, which decreased by 8.8°, while C12–N14–H15 and C1–C12–N14 bond angles show increases of 3.8° and 2.6°, respectively. Frontier molecular orbital (FMO) analysis reveals that EHOMO of C18B3N3 changes from −7.04 eV to −5.84 eV after adsorption of PA; also, ELUMO stabilizes by approximately 1.14 eV (Fig. 5). Therefore, the Eg value decreases by 1.2%; thus, the electronic properties of C18B3N3 are sensitive to the presence of PA molecules. Here again the reductions in parameters ω, χ, A, and I are in agreement with the reduction in Eg (Fig. 6).
Bond | Bond length (Å) | Δr | |
---|---|---|---|
After interaction | Before interaction | ||
C1–C2 | 1.352 | 1.341 | 0.012 |
C3–C4 | 1.331 | 1.333 | −0.002 |
C2–H5 | 1.084 | 1.085 | −0.001 |
C3–H6 | 1.080 | 1.082 | −0.002 |
C4–H7 | 1.082 | 1.083 | −0.001 |
C4–N8 | 1.423 | 1.403 | 0.02 |
N8–H9 | 1.019 | 1.011 | 0.008 |
C1–N8 | 1.427 | 1.407 | 0.020 |
C2–N10 | 1.375 | 1.413 | −0.038 |
C3–N10 | 1.427 | 1.434 | −0.0.007 |
N10–H11 | 1.011 | 1.016 | −0.005 |
C1–C12 | 1.450 | 1.486 | −0.036 |
C12–O13 | 1.284 | 1.221 | 0.063 |
C12–N14 | 1.324 | 1.369 | −0.045 |
N14–H15 | 1.008 | 1.007 | 0.001 |
N14–H16 | 1.011 | 1.009 | 0.002 |
The decrease in electrophilicity upon adsorption of PA is calculated to be approximately 0.159 eV, showing that charge transfers from the drug molecule to C18B3N3. Table S4† represents the partial atomic charges of the PA molecule before and after interaction with C18B3N3. The charge of C1, C4, N8, and O13 atoms increased by approximately 0.048, 0.012, 0.001, and 0.022 |e|, respectively. The comparison of the total atomic charges of PA and C18B3N3 shows that due to the formation of a complex between PA and C18B3N3, 0.398 |e| of charge is transferred from the drug to the nanocarrier.
The NBO analysis results showed that the highest values of stabilization energies E(2) of the C18B3N3 nanocarrier are BD*(2) C33–C34 → BD*(2) C23–C24; E(2) = 49.82 kcal mol−1, BD*(2) C29–C30 → BD*(2) C31–C32; E(2) = 50.27 kcal mol−1, BD*(2) C27–C28 → BD*(2) C25–C26; E(2) = 49.69 kcal mol−1, BD(2) C33–C34 → LP*(1) B21; E(2) = 45.12 kcal mol−1, and BD(2) C25–C26 → LP*(1) B21; E(2) = 44.91 kcal mol−1. The antibonding orbitals of boron atoms are the most important non-Lewis acceptors. The greatest values of stabilization energy in the C18B3N3-PA complex are related to transitions within the C18B3N3 molecule: BD*(2) C33–C34 → BD*(2) C23–C24; E(2) = 34.89 kcal mol−1, BD*(2) C29–C30 → BD*(2) C31–C32; E(2) = 42.29 kcal mol−1, BD*(2) C27–C28 → BD*(2) C25–C26; E(2) = 37.10 kcal mol−1, and BD*(2) C37–C36 → BD*(2) C23–C24; E(2) = 57.13 kcal mol−1. The most important stabilization energies related to the interaction between C18B3N3 and PA are BD(2) C25–C26 → LP*(1) B21; E(2) = 46.33 kcal mol−1 and BD(1) C25–C26 → LP*(1) B21; E(2) = 3.43 kcal mol−1, which indicates charge transfer from the drug molecule to C18B3N3.
![]() | ||
Fig. 7 (a) Optimized geometry, (b) MEP, (c) HOMO, and (d) LUMO profiles of C12B6N6-PA at the M06-2X/6-31G(d,p) level pf theory. |
Table 5 shows that despite the relatively strong interaction between the drug molecule and C12B6N6, the geometrical parameters of PA did not change significantly. However, the bond length changes are more evident in the areas involved in the interaction with the nanostructure. For example, after forming the C12B6N6-PA complex, the C–O bond length of PA has increased by approximately 0.057 Å. The process of bond length changes in C12B6N6-PA is similar to that in the C18B3N3-PA and C24-PA complexes. Additionally, the lengths of C1–C12 and C1–N8 bonds increased, which can be attributed to resonance effects. The highest bond angle increase due to PA adsorption on the C12B6N6 surface is related to C1–C12–N14, C2–N10–H11, and C12–N14–H16 angles, which increased by 4.7°, 2.5° and 2.4°, respectively. The C1–C12–O13 bond angle decreased from 120.1° to 115.4° (Table S5†).
Bond | Bond length (Å) | Δr | |
---|---|---|---|
After interaction | Before interaction | ||
C1–C2 | 1.350 | 1.341 | 0.009 |
C3–C4 | 1.329 | 1.333 | −0.004 |
C2–H5 | 1.084 | 1.085 | −0.001 |
C3–H6 | 1.080 | 1.082 | −0.002 |
C4–H7 | 1.081 | 1.083 | −0.002 |
C4–N8 | 1.416 | 1.403 | 0.013 |
N8–H9 | 1.014 | 1.011 | 0.003 |
C1–N8 | 1.425 | 1.407 | 0.018 |
C2–N10 | 1.375 | 1.413 | −0.038 |
C3–N10 | 1.433 | 1.434 | −0.001 |
N10–H11 | 1.009 | 1.016 | −0.007 |
C1–C12 | 1.450 | 1.486 | −0.036 |
C12–O13 | 1.278 | 1.221 | 0.057 |
C12–N14 | 1.326 | 1.369 | −0.043 |
N14–H15 | 1.006 | 1.007 | −0.001 |
N14–H16 | 1.024 | 1.009 | 0.015 |
According to Table 3, the Eg of C12B6N6 has decreased by 0.24 eV after interaction with the drug molecule, which indicates the high reactivity of the nanocarrier (Fig. 8). The significant decrease of Eg shows that the electronic properties of C12B6N6 are very sensitive to the presence of drug molecules in the environment, and therefore, C12B6N6 can be used as a PA sensor. The reduction in parameters ω, χ, A, and I is in agreement with the reduction in Eg. It is also observed that with the decrease in the Eg value, the softness and chemical hardness have increased and decreased, respectively, which indicates the tendency of the nanocarrier to interact with the PA molecule. A considerable decrease in the electrophilicity parameter by 1.60 eV confirms the charge transfer from the drug to the nanocarrier.
Table S6† shows that the partial charge of PA atoms is affected by the presence of the C12B6N6 nanocarrier in the environment. Therefore, the charges of the C1, C3, N8, and O13 atoms increased by 0.029, 0.001, 0.009, and 0.020 |e|. On the other hand, the charges of the C2, N14 and N10 atoms have become more positive by 0.078, 0.067, and 0.053 |e| respectively. Additionally, NBO calculations show that due to the formation of a complex between PA and C12B6N6, a charge amount of 0.324 |e| is transferred from the drug to the nanocarrier. The dipole moment is an important parameter for understanding the symmetry of the complex; the higher the dipole moment, the higher the interaction between the adsorbate and adsorbent molecules.52 The significant increase in the dipole moment of C12B6N6 indicates that the solubility of C12B6N6-PA increases upon the adsorption process.
Examining the stabilization energies E(2) shows that the most important electron transfers of the drug molecule are: LP(1) N14 → BD*(2) C12–O13; E(2) = 56.14 kcal mol−1, LP(1) N10 → BD*(2) C1–C2; E(2) = 34.28 kcal mol−1, LP(1) N8 → BD*(2) C1–C2; E(2) = 33.13 kcal mol−1, LP(2) O13 → BD*(1) C12–N14; E(2) = 29.41 kcal mol−1, and BD(2) C1–C2 → BD*(2) C12–O13; E(2) = 19.41 kcal mol−1. The most important donors are free electron pairs of nitrogen and oxygen that fill the C1–C2 and C12–O13 antibonding orbitals. Additionally, the greatest stabilization energy E(2) values of the nanocarrier are: LP(1) N25 → LP*(1) B17; E(2) = 24.73 kcal mol−1, LP(1) N29 → LP*(1) B19; E(2) = 24.75 kcal mol−1, BD(2) N17–B19 → LP*(1) B19; E(2) = 62.33 kcal mol−1, BD(2) N17–B19 → LP*(1) B121; E(2) = 63.09 kcal mol−1, and BD(2) N33–B21 → LP*(1) B19; E(2) = 62.86 kcal mol−1. As seen, non-Lewis acceptors are antibonding orbitals of boron atoms. Regarding the donor–acceptor transitions in the C12B6N6-PA complex, it should be said that the largest amounts of stabilization energy are related to transitions within the C12B6N6 molecule which are: BD*(2) C24–N25 → LP*(1) B23; E(2) = 99.29 kcal mol−1, LP(1) C2 → LP*(1) C4; E(2) = 90.21 kcal mol−1, LP(1) C26 → BD*(2) C24–N25; E(2) = 74.59 kcal mol−1, BD(2) N27–B18 → LP*(1) B17; E(2) = 70.14 kcal mol−1, and BD*(2) C12–N14 → BD*(2) C1–C2; E(2) = 69.22 kcal mol−1. The most important stabilization energies related to the interaction between C12B6N6 and PA are: BD(2) B17–N18 → BD*(1) B17–O13; E(2) = 13.34 kcal mol−1, LP(2) O13 → BD*(1) B17–N25; E(2) = 6.22 kcal mol−1, and BD(2) C12–N14 → BD*(1) B17–O; E(2) = 1.35 kcal mol−1, which is in agreement with the increase of the C12–O13 bond and also bond formation between the PA oxygen and the boron atom of C12B6N6. As can be seen, the replacement of carbon atoms with boron and nitrogen atoms increased the stabilization energies.
HC = GC + VC | (12) |
On the other hand, the virial theorem states a relation between the Laplacian of the electron density at BCP and its other characteristics.
(1/4)∇2ρC = 2GC + VC | (13) |
A negative value of the Laplacian indicates the concentration of the electron density among the nuclei of interacting atoms, and one may assume a shared interaction such as covalent bonds and lone pairs. A positive Laplacian electron density value at corresponding BCP shows the depletion of electron density as in ionic, van der Waals, and hydrogen bond interactions. Fig. 9 illustrates the molecular graph of C24-PA, C18B3N3-PA, and C12B6N6-PA complexes obtained from AIM calculations. The graphs were obtained at the M06-2X/6-31G(d,p) level of theory. Large circles correspond to attractors and small red and yellow circles are bond and ring critical points, respectively. However, sometimes ∇2ρC > 0 but the total electronic energy density, HC, is negative and such a situation is attributed to the partially covalent interaction. Fig. 9 shows that the adsorption of the PA molecule on the surface of the C24 nanocage led to the appearance of two critical points between the drug and the nanostructure. The electron density properties calculated for the C24-PA complex reveal that for the O13PA⋯C19 interaction ∇2ρC > 0 and HC > 0; hence, this interaction belongs to the category of electrostatic interactions. The C4PA⋯C20 interaction has more electron density (0.006 a.u.), and the values of ∇2ρC (0.019 a.u.) and HC (0.001 a.u.) are positive. Therefore, this bond is also the result of an electrostatic interaction (Table 6).
Complex | Interaction site | ρC | ∇2ρC | GC | VC | HC | Type of interaction |
---|---|---|---|---|---|---|---|
C24-PA | O13(PA)⋯C19 | 0.010 | 0.031 | 0.007 | −0.006 | 0.001 | Electrostatic |
C4(PA)⋯C20 | 0.006 | 0.019 | 0.004 | −0.003 | 0.001 | ||
C18B3N3-PA | O(PA)⋯B17 | 0.117 | 0.512 | 0.194 | −0.259 | −0.065 | Partly covalent |
C12B6N6-PA | O13(PA)⋯B17 | 0.126 | −0.138 | 0.209 | −0.280 | −0.071 | Covalent |
H(PA)⋯N18 | 0.033 | 0.089 | 0.023 | −0.024 | −0.001 | Partly covalent |
AIM calculations showed that only one critical point has appeared between O13 of the PA molecule and B17 of the C18B3N3 nanostructure, for which the values of ρC, ∇2ρC and HC are 0.117, 0.512, and −0.065 a.u. respectively. Therefore, the OPA⋯B17 interaction is partially covalent. The electron density properties calculated for the C12B6N6-PA complex show that the O13PA⋯B17 bond possesses a significant charge density (0.126 a.u.). Moreover, the values of ∇2ρC (−0.138 a.u.) and HC (−0.071 a.u.) are negative. Therefore, as we mentioned before, the O13PA⋯B17 bond is a covalent interaction. AIM calculations show that the HPA⋯N18 interaction has lower values of ρC and ∇2ρC > 0; however, this interaction is categorized as a partly covalent interaction because ∇2ρC > 0 and HC < 0.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4na00560k |
This journal is © The Royal Society of Chemistry 2025 |