Desta Regassa Goljac,
Megersa Olumana Dinkaa,
Umer Sherefedin*b,
Abebe Belayc,
Dereje Gelanuc and
Gadisa Deme Megersac
aDepartment of Civil Engineering Science, Faculty of Engineering and the Built Environment, University of Johannesburg, Johannesburg 2006, South Africa
bDepartment of Physics, College of Natural and Computational Sciences, Madda Walabu University, Bale Robe, P.O. Box 247, Ethiopia. E-mail: umerphysics2005@gmail.com
cDepartment of Applied Physics, School of Applied Natural Sciences, Adama Science and Technology University, Adama, P.O. Box 1888, Ethiopia
First published on 11th August 2025
Metronidazole is widely used as an antimicrobial, particularly effective against anaerobic bacteria and protozoan infections. This study investigates solvent polarity effects on the Fourier transform infrared (FTIR) spectrum, and thermodynamic and electronic properties of metronidazole via semiempirical, Hartree–Fock (HF), and density functional theory (DFT) methods. Its binding with antibacterial drugs was also investigated via molecular docking. The results showed that in water, the dipole moment and polarizability increased, indicating enhanced solubility and reactivity. Solvent-induced changes in bond lengths and angles are important for understanding the behavior of metronidazole in biological systems. FTIR reveals changes in molecular interactions due to solvation effects, especially hydrogen bonding in water. Thermodynamic calculations further revealed that polar solvents increase the energy and dipole moment, enhancing the reactivity of the molecule. Frontier molecular orbital (FMO) analysis indicated that the molecules are more stable in polar environments, while UV-Vis spectral shifts showed that the solvent affects the electronic properties. Molecular docking studies with antibacterial proteins revealed that metronidazole binds strongly to proteins, with the metronidazole-4kov complex showing the highest binding affinity. Molecular docking of metronidazole with secnidazole, tizoxanide, and caffeine enhances the binding affinities, suggesting synergistic effects. In conclusion, this study emphasizes the importance of solvent polarity for optimizing the antibacterial properties of metronidazole and its molecular docking with other drugs.
![]() | ||
Fig. 1 The chemical structures of metronidazole (a), secnidazole (b), tizoxanide (c), and caffeine (d). |
Recently, the effects of solvent polarity on the dipole moment, Fourier transform infrared (FTIR) spectrum, highest occupied molecular orbital (HOMO), lowest unoccupied molecular orbital (LUMO), chemical reactivity, density of states (DOS), electrostatic potentials (ESPs), ultraviolet-visible (UV-Vis) spectra, and fluorescence of various drugs have been studied. These drugs include metformin hydrochloride,9 aspirin,10 zaleplon,11 nifenazone,12 imiquimod,13 sulfisoxazole,14 and hydroxycinnamic acids like sinapic acid,15 chlorogenic acid and caffeic acid,16 and ferulic acid.17 The results of these studies revealed the general solvent effect, which is related to the relative permittivity and refractive index. In addition, a specific effect, driven by hydrogen bonding and intermolecular charge transfer, occurred between the drugs and solvents. As the polarity of solvents changes, shifts in the absorption and emission peaks occur, leading to changes in the dipole moment, FTIR spectrum, HOMO–LUMO gap, chemical reactivity, DOS, and MEP of selected drugs due to solvent–drug interactions. Estimating these changes in drug properties in both the ground and excited states through solvatochromic effects via density functional theory (DFT) and time-dependent DFT (TD-DFT) is essential for understanding the electronic properties and structural modifications of these drugs. The biological activities of molecules depend on their molecular structure. Even small changes in drug properties due to solvent–drug interactions can signal structural modifications. These modifications can, in turn, affect the biological activities of the drug, making properties such as the dipole moment, FTIR spectrum, HOMO–LUMO gap, chemical reactivity, DOS, and MEP important measurable factors in drug analysis.9–17
On the other hand, the binding between ligands and proteins,18 multiple ligand–protein interactions,19,20 or ligand–ligand interactions21 are crucial for biological activity. The pharmacological effectiveness of a drug largely depends on its ability to bind with proteins or its potential for drug–drug interactions.22 Any changes in this binding can directly impact the drug’s activity. Recently, the simultaneous use of multiple drugs has increased, both knowingly and unknowingly.23 This concurrent drug use can lead to interactions that may either enhance or diminish the biological activity of a drug, with such interactions occurring between proteins and ligands or through multiple ligands interacting with proteins.24 Sherefedin et al. (2025) reported the molecular docking of hydroxycinnamic acids such as ferulic, p-coumaric, caffeic, and sinapic acids with anticancer-related proteins such as 3M18, 5EKN, and 6YKY.25 The results revealed strong binding affinities, with favorable root mean square deviation (RMSD) values, indicating stable interactions and potential as anticancer agents. Molecular docking studies were also performed on salicylidene–aniline and their metal mixed-ligand complexes in interaction with caffeine. The results showed that the metal–caffeine complexes had stronger binding affinities than the free ligands, suggesting enhanced biological potential.26 Another study examined the impact of caffeine and flavonoids on tigecycline’s binding to human serum albumin. Docking results revealed that both compounds altered tigecycline’s binding affinity.27 Woldegiorges et al. (2022) reported that the interaction of caffeine with levofloxacin and norfloxacin leads to significant fluorescence quenching, indicating strong molecular interactions between caffeine and these drugs. The quenching effect is attributed to the binding of caffeine with the fluorophores of these drugs, which alters their photophysical properties.28 Furthermore, the interaction between caffeine and aspirin in Kopi Balur 1 was investigated. The results showed that this interaction influences the biological activity of the compound.29
Previously, research has investigated the effects of solvent polarity on drugs such as metformin, aspirin, zaleplon, and hydroxycinnamic acids, including ferulic, p-coumaric, caffeic, and sinapic acids, with a focus on their structure, thermodynamics, and electronic properties via DFT and molecular docking methods. However, the impact of solvent polarity on metronidazole, particularly its antibacterial activity, has not been explored. On the other hand, previously, drug–protein and drug–drug interactions have been investigated for other compounds using techniques like molecular docking and fluorescence quenching; however, there is a notable absence of studies specifically examining metronidazole’s interactions with antibacterial proteins or its behavior in multiple-ligand interactions with agents such as secnidazole (Fig. 1(b)), tizoxanide (Fig. 1(c)), and caffeine (Fig. 1(d)). Therefore, this study addresses these gaps by investigating how solvent polarity affects the structure and properties of metronidazole via semiempirical, Hartree–Fock (HF), and DFT (B3LYP) methods with various basis sets. It also explores drug–drug interactions, particularly with amino acids, through molecular docking (AutoDock Vina 1.1.2, PyRx version 0.8). The goal is to better understand how solvent polarity and drug interactions influence the biological activity and effectiveness of metronidazole.
Ligand preparation: The ligands metronidazole, secnidazole, tizoxanide, and caffeine were prepared via ChemDraw Ultra 8.0.30 Chem3D Ultra43 was used to minimize energy, stabilize their conformations and reduce steric strain. DFT (B3LYP/6-311++G(d,p)) was used to optimize the geometries (Fig. 2(a)–(d)) and was used for docking.
Molecular docking: Molecular docking was performed by importing the cleaned proteins (8fb0, 4kov, 5j62, and 3q5p) into AutoDock Vina 1.1.2.37 Kollman and Gasteiger charges were assigned to optimize the electrostatic properties, and AD4 atom types were applied for compatibility with the docking algorithm. The metronidazole ligand was imported, and a torsion tree was added. The active sites for each protein were identified based on the positions of co-crystallized ligands in the crystal structures available from the Protein Data Bank. These positions were used to set the grid box coordinates (x, y, z) in AutoDock Vina 1.1.2, ensuring docking within the biologically relevant binding pockets. After loading the ligand (ligand.pdbqt) and setting the docking parameters, the results were obtained via the command prompt, which predicts the binding affinities and amino acid interactions. Post-docking analysis was conducted via BIOVIA Discovery Studio, which visualized the binding sites, hydrogen bonds, hydrophobic interactions, and bond distances. The best ligand pose was selected on the basis of hydrogen bond interactions and visualized in both 2D and 3D. In addition, PyRx version 0.8 (ref. 38) was used to dock multiple inhibitors, including metronidazole, secnidazole, and tizoxanide, with the receptor proteins. PyRx version 0.8 (ref. 38) automated the ligand and protein preparation, converting ligands to the PDBQT format for compatibility with AutoDock Vina 1.1.2.37 The docking grid was adjusted to target the receptor’s active site, enabling an efficient search for optimal ligand–receptor interactions. Using the Vina algorithm, PyRx version 0.8 (ref. 38) was used to calculate binding affinities and ranked ligand poses based on docking scores. The results were then analyzed to identify the strongest binding conformations. Finally, Discovery Studio was used to visualize binding interactions, focusing on hydrogen bonding, hydrophobic contacts, and key amino acid interactions.
Calculation method | Vacuum | Water | ||||||||
---|---|---|---|---|---|---|---|---|---|---|
μ | α | E | CV | S | μ | α | E | CV | S | |
PM6 | 4.037 | 83.538 | 100.854 | 42.832 | 112.334 | 5.650 | 110.354 | 100.276 | 42.998 | 111.484 |
PDDG | 4.057 | 76.664 | 106.825 | 41.653 | 109.003 | 5.164 | 96.883 | 106.649 | 41.621 | 107.701 |
AM1 | 3.704 | 84.447 | 111.878 | 40.083 | 108.719 | 4.910 | 110.895 | 111.594 | 40.148 | 108.305 |
PM3 | 3.927 | 76.897 | 108.298 | 42.231 | 110.403 | 5.1195 | 97.412 | 108.113 | 42.270 | 110.229 |
PM3MM | 3.927 | 76.897 | 108.298 | 42.231 | 110.403 | 5.119 | 97.412 | 108.113 | 42.270 | 110.229 |
Table 2 shows dipole moment, polarizability, thermal energy, heat capacity, and entropy of metronidazole calculated using the HF method. The dipole moment increases notably in water compared to in a vacuum, reflecting the solvent’s polarizing effect, with values ranging from 2.707 D (STO-3G*) in vacuum to 5.696 D (3-21+G*) in water. The polarizability also increases consistently in aqueous medium, indicating enhanced electron cloud distortion. While thermal energy, heat capacity, and entropy show minor fluctuations across basis sets, they remain relatively stable between vacuum and solvent conditions, suggesting that solvation has a more pronounced effect on electrostatic properties than on thermal behavior. Among the tested basis sets, larger and more diffuse functions like Aug-CC-pVDZ and 6-311+G(d,p) yield higher accuracy and consistent values, making them more reliable for capturing solvation effects and molecular response properties.
Basis sets | HF (vacuum) | HF (water) | ||||||||
---|---|---|---|---|---|---|---|---|---|---|
μ | α | E | CV | S | μ | α | E | CV | S | |
STO-3G* | 2.707 | 54.102 | 125.006 | 38.764 | 106.853 | 3.324 | 63.322 | 124.963 | 38.767 | 106.898 |
3-21+G* | 3.851 | 95.916 | 115.572 | 39.335 | 104.145 | 5.696 | 130.193 | 115.211 | 39.474 | 103.745 |
6-31+G (d, p) | 3.424 | 94.989 | 117.537 | 38.772 | 104.422 | 4.613 | 125.056 | 117.247 | 38.882 | 104.719 |
6-311+G (d, p) | 3.481 | 95.51 | 117.088 | 38.882 | 104.706 | 4.681 | 125.777 | 116.806 | 38.969 | 104.865 |
Aug-CC-pVDZ | 3.504 | 100.957 | 116.995 | 38.845 | 104.528 | 4.772 | 134.494 | 116.726 | 38.928 | 104.547 |
LANL2DZ | 3.661 | 87.644 | 117.272 | 38.982 | 104.775 | 5.066 | 114.889 | 117.037 | 39.002 | 104.388 |
SDD | 3.658 | 87.679 | 117.243 | 38.985 | 104.773 | 5.063 | 114.93 | 117.009 | 39.007 | 104.398 |
Table 3, shows the calculated dipole moment (μ), polarizability (α), thermal energy (E), heat capacity (CV), and entropy (S) of metronidazole in vacuum and water using various DFT basis sets. The dipole moment increases in water for all basis sets, indicating stronger molecular polarity due to solvent effects. Smaller basis sets like STO-3G* underestimate μ and α values, while larger, more flexible basis sets such as 3-21+G* and 6-31+G(d,p), provide higher and more accurate values. Polarizability also rises significantly in water, reflecting enhanced electron cloud distortion. Thermal energy (E), heat capacity (CV), and entropy (S) exhibit only slight variations across different basis sets and solvation conditions, suggesting that these thermodynamic properties are relatively insensitive to the level of basis set used in DFT calculations. These findings highlight that selecting an appropriate basis set is critical for accurately capturing electronic properties, especially dipole moment and polarizability, in DFT calculations involving solvation effects.
Basis sets | DFT (Vacuum) | DFT (Water) | ||||||||
---|---|---|---|---|---|---|---|---|---|---|
μ | α | E | CV | S | μ | α | E | CV | S | |
STO-3G* | 2.06 | 61.476 | 114.017 | 41.944 | 109.469 | 2.632 | 72.815 | 114.028 | 41.998 | 109.677 |
3-21+G* | 3.944 | 108.798 | 108.405 | 42.132 | 107.241 | 5.711 | 149.643 | 108.391 | 42.079 | 106.526 |
6-31+G (d, p) | 3.611 | 108.664 | 109.701 | 41.655 | 107.509 | 5.065 | 148.19 | 109.597 | 41.697 | 107.497 |
6-311++G (d, p) | 3.439 | 98.298 | 109.284 | 41.66 | 106.975 | 5.035 | 148.392 | 109.236 | 41.781 | 107.913 |
LanL2DZ | 3.737 | 96.83 | 109.443 | 41.808 | 107.612 | 5.185 | 129.154 | 109.383 | 41.863 | 107.836 |
SDD | 3.733 | 96.977 | 109.43 | 41.819 | 107.621 | 5.18 | 129.393 | 109.369 | 41.874 | 107.852 |
Fig. 3 and 4 illustrate the optimized ground-state chemical structures of metronidazole from different computational methods and environments, facilitating a comparative analysis of molecular geometry under vacuum and aqueous conditions. The optimized structures of metronidazole in various environments highlight significant molecular interactions. In Fig. 3, the hydrogen atom H17 interacts with oxygen (O1) and nitrogen atoms, showcasing potential hydrogen bonding interactions. In the absence of solvent, the molecule exhibits a more planar configuration, particularly for the nitro group, with C–N–O angles remaining consistent. However, this stability is sensitive to environmental changes, as evidenced by the planar nature of the nitro group, which suggests an ideal electronic distribution for bonding interactions.
![]() | ||
Fig. 3 Optimized structures of metronidazole via the vacuum semiempirical method (MP6) (a), Hartree–Fock (b), and B3LYP/6-311++G (d, p) (c). |
![]() | ||
Fig. 4 Optimized structures of metronidazole in the water via the semiempirical method (MP6) (a), Hartree–Fock (b), and B3LYP/6-311++G (d, p) (c). |
In Fig. 4, the interaction of H17 with the O1 and N atoms is affected. This results in slight shifts and indications of solvent-induced stabilization through additional hydrogen bonding. The nitro group attached to C9 exhibits a discernible out-of-plane twist. This twist is indicative of solvation effects that alter the electron distribution and bond angles compared to its vacuum state. The rotation of the hydroxyl group connected to C10 and the reorientation of the methyl group on C8 reveal the influence of solvent interactions. These interactions affect torsional strain and dielectric stabilization. Furthermore, slight changes in dihedral angles across the molecule suggest that electrostatic forces promote bent conformations in polar environments. This confirms the dynamic response of metronidazole to solvent polarity.
Table 4 provides the optimized geometric parameters of metronidazole in a vacuum at 298.15 K, highlighting the impact of the solvent on the molecular structure. In a vacuum, the bond lengths show typical covalent characteristics, such as O(1)–C(10) at 1.453 Å, while the bond angles exhibit values like C(10)–O(1)–H(13) at 110.355°, reflecting a stable configuration influenced by electronic repulsions. The dihedral angles reveal substantial flexibility, as seen with H(13)–O(1)–C(10)–C(7) at −83.159°, indicating that steric and torsional dynamics could allow the molecule to adopt multiple conformations.
Bond length | Values (Å) | Bond angles | Values (°) | Dihedral angles | Values (°) |
---|---|---|---|---|---|
O(1)–C(10) | 1.453 | C(10)–O(1)–H(13) | 110.355 | H(13)–O(1)–C(10)–C(7) | −83.159 |
O(1)–H(13) | 0.973 | C(7)–N(4)–C(8) | 125.536 | H(13)–O(1)–C(10)–H(16) | 157.693 |
O(2)–N(6) | 1.271 | C(7)–N(4)-C(9) | 128.731 | H(13)–O(1)–C(10)–H(17) | 38.686 |
O(3)–N(6) | 1.282 | C(8)–N(4)–C(9) | 105.731 | C(8)–N(4)–C(7)–C(10) | 97.765 |
N(4)–C(7) | 1.476 | C(8)–N(5)–C(11) | 106.488 | C(8)–N(4)–C(7)–H(14) | −23.003 |
N(4)–C(8) | 1.378 | O(2)–N(6)–O(3) | 123.24 | C(8)–N(4)–C(7)–H(15) | −140.654 |
N(4)–C(9) | 1.402 | O(2)–N(6)–C(9) | 117.515 | C(9)–N(4)–C(7)–C(10) | −81.667 |
N(5)–C(8) | 1.348 | O(3)–N(6)–C(9) | 119.244 | C(9)–N(4)–C(7)–H(14) | 157.565 |
N(5)–C(11) | 1.37 | N(4)–C(7)–C(10) | 112.756 | C(9)–N(4)–C(7)–H(15) | 39.913 |
N(6)–C(9) | 1.411 | N(4)–C(7)–H(14) | 107.809 | C(7)–N(4)–C(8)–N(5) | −179.647 |
C(7)–C(10) | 1.534 | N(4)–C(7)–H(15) | 108.449 | C(7)–N(4)–C(8)–C(12) | 1.413 |
C(7)–H(14) | 1.087 | C(10)–C(7)–H(14) | 109.32 | C(9)–N(4)–C(8)–N(5) | −0.107 |
C(7)–H(15) | 1.084 | C(10)–C(7)–H(15) | 109.615 | C(9)–N(4)–C(8)–C(12) | −179.047 |
C(8)–C(12) | 1.486 | H(14)–C(7)–H(15) | 108.805 | C(7)–N(4)–C(9)–N(6) | −0.255 |
C(9)–C(11) | 1.384 | N(4)–C(8)–N(5) | 111.32 | C(7)–N(4)–C(9)–C(11) | 179.725 |
C(10)–H(16) | 1.088 | N(4)–C(8)–C(12) | 124.628 | C(8)–N(4)–C(9)–N(6) | −179.775 |
C(10)–H(17) | 1.089 | N(5)–C(8)–C(12) | 124.044 | C(8)–N(4)–C(9)–C(11) | 0.205 |
C(11)–H(18) | 1.072 | N(4)–C(9)–N(6) | 125.362 | C(11)–N(5)–C(8)–N(4) | −0.036 |
C(12)–H(19) | 1.093 | N(4)–C(9)–C(11) | 106.904 | C(11)–N(5)–C(8)–C(12) | 178.912 |
C(12)–H(20) | 1.087 | N(6)–C(9)–C(11) | 127.734 | C(8)–N(5)–C(11)–C(9) | 0.169 |
C(12)–H(21) | 1.093 | O(1)–C(10)–C(7) | 109.19 | C(8)–N(5)–C(11)–H(18) | 179.827 |
O(1)–C(10)–H(16) | 105.954 | O(2)–N(6)–C(9)–N(4) | −176.356 | ||
O(1)–C(10)–H(17) | 111.811 | O(2)–N(6)–C(9)–C(11) | 3.668 | ||
C(7)–C(10)–H(16) | 110.605 | O(3)–N(6)–C(9)–N(4) | 3.906 | ||
C(7)–C(10)–H(17) | 109.908 | O(3)–N(6)–C(9)–C(11) | −176.07 | ||
H(16)–C(10)–H(17) | 109.314 | N(4)–C(7)–C(10)–O(1) | −177.424 | ||
N(5)–C(11)–C(9) | 109.557 | N(4)–C(7)–C(10)–H(16) | −61.207 | ||
N(5)–C(11)–H(18) | 122.773 | N(4)–C(7)–C(10)–H(17) | 59.589 | ||
C(9)–C(11)–H(18) | 127.669 | H(14)–C(7)–C(10)–O(1) | −57.523 | ||
C(8)–C(12)–H(19) | 112.065 | H(14)–C(7)–C(10)–H(16) | 58.694 | ||
C(8)–C(12)–H(20) | 107.873 | H(14)–C(7)–C(10)–H(17) | 179.49 | ||
C(8)–C(12)–H(21) | 112.479 | H(15)–C(7)–C(10)–O(1) | 61.658 | ||
H(19)–C(12)–H(20) | 108.142 | H(15)–C(7)–C(10)–H(16) | 177.876 | ||
H(19)–C(12)–H(21) | 107.708 | H(15)–C(7)–C(10)–H(17) | −61.329 | ||
H(20)–C(12)–H(21) | 108.438 | N(4)–C(8)–C(12)–H(19) | 63.444 | ||
N(4)–C(8)–C(12)–H(20) | −177.627 | ||||
N(4)–C(8)–C(12)–H(21) | −58.084 | ||||
N(5)–C(8)–C(12)–H(19) | −115.365 | ||||
N(5)–C(8)–C(12)–H(20) | 3.565 | ||||
N(5)–C(8)–C(12)–H(21) | 123.107 | ||||
N(4)–C(9)–C(11)–N(5) | −0.235 | ||||
N(4)–C(9)–C(11)–H(18) | −179.872 | ||||
N(6)–C(9)–C(11)–N(5) | 179.744 | ||||
N(6)–C(9)–C(11)–H(18) | 0.107 |
Table 5 provides the optimized geometric parameters of metronidazole in water at 298.15 K. The optimized parameters in water reveal significant alterations; for instance, the O(1)–H(13) bond shortens to 0.964 Å and there are changes in angles, such as C(7)–N(4)–C(9) changing to 129.398°, suggesting that solvent interactions promote changes in molecular geometry, potentially enhancing hydrogen bonding and affecting overall stability. Moreover, the dihedral angle H(15)–C(7)–C(10)–H(16) changes from 177.876° in the gas phase (Table 4) to 179.14° in water (Table 5). This slight increase indicates a solvent-induced conformational adjustment. The polar water environment stabilizes a more extended geometry, reflecting the influence of solvation on molecular structure.
Bond lengths | Values (Å) | Bond angles | Values (°) | Dihedral angles | Values (°) |
---|---|---|---|---|---|
O(1)–C(10) | 1.425 | C(10)–O(1)–H(13) | 109.023 | H(13)–O(1)–C(10)–C(7) | −75.819 |
O(1)–H(13) | 0.964 | C(7)–N(4)–C(8) | 125.241 | H(13)–O(1)–C(10)–H(16) | 164.772 |
O(2)–N(6) | 1.236 | C(7)–N(4)–C(9) | 129.398 | H(13)–O(1)–C(10)–H(17) | 46.848 |
O(3)–N(6) | 1.237 | C(8)–N(4)–C(9) | 105.349 | C(8)–N(4)–C(7)–C(10) | 97.426 |
N(4)–C(7) | 1.472 | C(8)–N(5)–C(11) | 106.317 | C(8)–N(4)–C(7)–H(14) | −23.158 |
N(4)–C(8) | 1.36 | O(2)–N(6)–O(3) | 123.393 | C(8)–N(4)–C(7)–H(15) | −140.198 |
N(4)–C(9) | 1.393 | O(2)–N(6)–C(9) | 117.285 | C(9)–N(4)–C(7)–C(10) | −81.203 |
N(5)–C(8) | 1.338 | O(3)–N(6)–C(9) | 119.322 | C(9)–N(4)–C(7)–H(14) | 158.212 |
N(5)–C(11) | 1.35 | N(4)–C(7)–C(10) | 112.076 | C(9)–N(4)–C(7)–H(15) | 41.173 |
N(6)–C(9) | 1.41 | N(4)–C(7)–H(14) | 107.324 | C(7)–N(4)–C(8)–N(5) | −179.039 |
C(7)–C(10) | 1.535 | N(4)–C(7)–H(15) | 108.547 | C(7)–N(4)–C(8)–C(12) | 1.878 |
C(7)–H(14) | 1.089 | C(10)–C(7)–H(14) | 109.767 | C(9)–N(4)–C(8)–N(5) | −0.137 |
C(7)–H(15) | 1.088 | C(10)–C(7)–H(15) | 110.542 | C(9)–N(4)–C(8)–C(12) | −179.22 |
C(8)–C(12) | 1.489 | H(14)–C(7)–H(15) | 108.465 | C(7)–N(4)–C(9)–N(6) | −1.483 |
C(9)–C(11) | 1.382 | N(4)–C(8)–N(5) | 111.996 | C(7)–N(4)–C(9)–C(11) | 179.043 |
C(10)–H(16) | 1.092 | N(4)–C(8)–C(12) | 124.096 | C(8)–N(4)–C(9)–N(6) | 179.679 |
C(10)–H(17) | 1.093 | N(5)–C(8)–C(12) | 123.901 | C(8)–N(4)–C(9)–C(11) | 0.204 |
C(11)–H(18) | 1.079 | N(4)–C(9)–N(6) | 125.933 | C(11)–N(5)–C(8)–N(4) | 0.012 |
C(12)–H(19) | 1.094 | N(4)–C(9)–C(11) | 106.735 | C(11)–N(5)–C(8)–C(12) | 179.097 |
C(12)–H(20) | 1.089 | N(6)–C(9)–C(11) | 127.33 | C(8)–N(5)–C(11)–C(9) | 0.123 |
C(12)–H(21) | 1.093 | O(1)–C(10)–C(7) | 110.431 | C(8)–N(5)–C(11)–H(18) | 179.997 |
O(1)–C(10)–H(16) | 106.426 | O(2)–N(6)–C(9)–N(4) | −178.382 | ||
O(1)–C(10)–H(17) | 111.221 | O(2)–N(6)–C(9)–C(11) | 0.985 | ||
C(7)–C(10)–H(16) | 110.038 | O(3)–N(6)–C(9)–N(4) | 1.715 | ||
C(7)–C(10)–H(17) | 110.191 | O(3)–N(6)–C(9)–C(11) | −178.918 | ||
H(16)–C(10)–H(17) | 108.445 | N(4)–C(7)–C(10)–O(1) | −176.825 | ||
N(5)–C(11)–C(9) | 109.602 | N(4)–C(7)–C(10)–H(16) | −59.627 | ||
N(5)–C(11)–H(18) | 123.218 | N(4)–C(7)–C(10)–H(17) | 59.908 | ||
C(9)–C(11)–H(18) | 127.18 | H(14)–C(7)–C(10)–O(1) | −57.667 | ||
C(8)–C(12)–H(19) | 111.315 | H(14)–C(7)–C(10)–H(16) | 59.531 | ||
C(8)–C(12)–H(20) | 108.389 | H(14)–C(7)–C(10)–H(17) | 179.066 | ||
C(8)–C(12)–H(21) | 111.962 | H(15)–C(7)–C(10)–O(1) | 61.942 | ||
H(19)–C(12)–H(20) | 108.456 | H(15)–C(7)–C(10)–H(16) | 179.14 | ||
H(19)–C(12)–H(21) | 107.97 | H(15)–C(7)–C(10)–H(17) | −61.325 | ||
H(20)–C(12)–H(21) | 108.661 | N(4)–C(8)–C(12)–H(19) | 64.067 | ||
N(4)–C(8)–C(12)–H(20) | −176.735 | ||||
N(4)–C(8)–C(12)–H(21) | −56.892 | ||||
N(5)–C(8)–C(12)–H(19) | −114.909 | ||||
N(5)–C(8)–C(12)–H(20) | 4.289 | ||||
N(5)–C(8)–C(12)–H(21) | 124.132 | ||||
N(4)–C(9)–C(11)–N(5) | −0.207 | ||||
N(4)–C(9)–C(11)–H(18) | 179.925 | ||||
N(6)–C(9)–C(11)–N(5) | −179.671 | ||||
N(6)–C(9)–C(11)–H(18) | 0.46 |
Fig. 5(b), presenting the FTIR spectrum of metronidazole in water, vividly illustrates the profound impact of solvation on its vibrational modes, contrasting sharply with the gas-phase spectrum (Fig. 5(a)) and providing critical data for validating computational models.
Solvent | Thermodynamic properties | Dipole moment (D) | Total | ||||
---|---|---|---|---|---|---|---|
E (kcal mol−1) | CV (cal mol−1 K−1) | S (cal mol−1 K−1) | X | Y | Z | ||
Gas | 109.353 | 41.739 | 107.881 | −2.841 | 2.019 | 0.853 | 3.588 |
Heptane | 109.334 | 41.738 | 107.814 | −3.193 | 2.213 | 0.948 | 3.999 |
Benzene | 109.331 | 41.733 | 107.747 | −3.272 | 2.251 | 0.971 | 4.088 |
Chloroform | 109.305 | 41.740 | 107.684 | −3.787 | 2.369 | 1.058 | 4.590 |
Dichloromethane | 109.279 | 41.752 | 107.726 | −3.956 | 2.467 | 1.127 | 4.796 |
Acetone | 109.254 | 41.767 | 107.817 | −4.074 | 2.541 | 1.178 | 4.944 |
Ethanol | 109.250 | 41.770 | 107.837 | −4.091 | 2.552 | 1.185 | 4.965 |
Methanol | 109.245 | 41.774 | 107.862 | −4.110 | 2.564 | 1.193 | 4.989 |
Acetonitrile | 109.244 | 41.775 | 107.869 | −4.115 | 2.568 | 1.195 | 4.996 |
Dimethyl sulfoxide | 109.240 | 41.778 | 107.888 | −4.129 | 2.577 | 1.201 | 5.013 |
Water | 109.236 | 41.781 | 107.913 | −4.146 | 2.588 | 1.208 | 5.035 |
Table 6 also provides the dipole moment values (X, Y and Z components and total) for the metronidazole molecule calculated in various solvents and the gas phase. The total dipole moment increases from 3.588 D in the gas phase to 5.035 D in water, indicating that the molecule gains significant polar character in more polar environments, which enhances its ability to interact with other polar molecules. The breakdown of the dipole moments into their X, Y, and Z components reveals that the increases in the total dipole moment are driven primarily by the Y and Z components, which shift from 2.019 D and 0.853 D in the gas phase to 2.588 D and 1.208 D in water, respectively. This suggests that solvent polarity particularly affects the molecular orientation and distribution of charge within metronidazole. Moreover, the negative X component suggests a conventional orientation of dipole moments, possibly reflecting the structural asymmetry of the molecule. Overall, the increase in dipole moment with increasing solvent polarity highlights how solvation can increase the molecular polarity of metronidazole, impacting its solubility and reactivity in biological systems, thereby playing a crucial role in its pharmacological behavior.
Solvent polarity | FMO parameters (eV) | Chemical reactivity parameters | ||||||
---|---|---|---|---|---|---|---|---|
HOMO | LUMO | ΔE | μcp | η | χ | Sg | ω | |
Gas | −2.919 | −7.399 | 4.471 | 5.159 | 2.240 | −5.159 | 0.212 | 5.634 |
Heptane | −2.955 | −7.349 | 4.394 | 5.151 | 2.196 | −5.151 | 0.218 | 5.791 |
Benzene | −2.964 | −7.34 | 4.376 | 5.152 | 2.188 | −5.152 | 0.219 | 5.825 |
Chloroform | −2.993 | −7.314 | 4.321 | 5.154 | 2.1605 | −5.1535 | 0.224 | 5.941 |
Dichloromethane | −3.008 | −7.3 | 4.292 | 5.154 | 2.146 | −5.154 | 0.226 | 6.002 |
Acetone | −3.017 | −7.29 | 4.272 | 5.154 | 2.1365 | −5.1535 | 0.227 | 6.04 |
Ethanol | −3.019 | −7.288 | 4.269 | 5.154 | 2.1345 | −5.1535 | 0.228 | 6.049 |
Methanol | −3.02 | −7.286 | 4.266 | 5.153 | 2.133 | −5.153 | 0.228 | 6.054 |
Acetonitrile | −3.021 | −7.286 | 4.265 | 5.154 | 2.1325 | −5.1535 | 0.228 | 6.057 |
Dimethyl sulfoxide | −3.022 | −7.285 | 4.263 | 5.154 | 2.1315 | −5.1535 | 0.228 | 6.061 |
Water | −3.023 | −7.283 | 4.26 | 5.382 | 2.144 | −5.382 | 0.226 | 6.554 |
Furthermore, Fig. 6 presents the molecular orbital surfaces for metronidazole, illustrating the HOMO, LUMO, and HOMO–LUMO gaps in both the gas phase and water. In the gas phase (Fig. 6(a)), the HOMO–LUMO gap is calculated to be 4.471 eV, indicating a relatively stable electronic configuration with significant electron density localized around the nitro and imidazole groups. This configuration supports the electron-donating ability of metronidazole, which is crucial for its reactivity and interaction with biological targets. In water (Fig. 6(b)), the HOMO–LUMO gap decreases slightly to 4.260 eV, suggesting that the presence of the polar solvent stabilizes the HOMO and increases the energy of the LUMO, facilitating easier electronic transitions. This decrease in the energy gap enhances the reactivity of metronidazole, increasing its susceptibility to electron transfer processes, which can facilitate its interaction with bacterial targets. The ability of metronidazole to dynamically adjust its electronic properties in response to the solvent environment may enhance its antibacterial activity, indicating that the effectiveness of the drug can be optimized in biological systems where polar environments are prevalent.
Chemical reactivity is closely tied to the properties of frontier molecular orbitals, as follows:
IP = −EHOMO | (1) |
EA = −ELUMO | (2) |
Furthermore, the chemical reactivity (eqn (3)–(7)) was calculated according to Koopman’s theory56 as follows:
Chemical potential57
![]() | (3) |
Chemical hardness52
![]() | (4) |
Electronegativity53
![]() | (5) |
Global softness58
![]() | (6) |
Global electrophilicity index55
![]() | (7) |
Table 7 also presents the chemical reactivity of metronidazole in both the gas phase and the solvent environment, as determined via eqn (1)–(7). The dipole moment (μ) values, ranging from 5.152 D in benzene to 5.382 D in the gas phase, suggest that metronidazole has a considerable permanent dipole, indicating a strong polar character that may affect its interactions in biological systems and enhance solubility in polar solvents. The chemical hardness (η) values of metronidazole, ranging from 2.240 eV in gas phase to 2.144 eV in water, suggest increased stability and reduced charge-transfer reactivity in polar solvents. The electronegativity (χ) values range from −5.152 to −5.382, which imply a moderate tendency for electron attraction, enhancing its potential reactivity with electrophiles. The softness (Sg) values, which increase slightly with the reduction in chemical hardness, indicate that in more polar solvents, metronidazole is generally more reactive, suggesting that solvent effects could facilitate various chemical interactions. The global electrophilicity index (ω) increases from 5.634 eV in the gas phase to 6.554 eV in water. This rise indicates that metronidazole becomes a stronger electrophile in polar environments. It suggests enhanced ability to accept electrons, reflecting greater chemical reactivity in solution. Collectively, these parameters reveal how solvent polarity influences the chemical landscape of metronidazole, providing crucial insights into its reactivity and interactions in medicinal chemistry.
Solvent | λmax (nm) | Oscillator strength (f) | Assigned transition |
---|---|---|---|
Benzene | 324 | 0.0003 | π → π∗ (HOMO → LUMO) |
Chloroform | 324.15 | 0.0043 | n → π∗ (HOMO-1 → LUMO) |
DCM | 324.81 | 0.1657 | Intramolecular charge transfer (HOMO → LUMO) |
Acetone | 327.73 | 0.3492 | π → π∗ (HOMO → LUMO) |
Ethanol | 328.22 | 0.3555 | n → π∗ (HOMO-2 → LUMO) |
Methanol | 328.78 | 0.3616 | Intramolecular charge transfer (HOMO → LUMO) |
Acetonitrile | 328.94 | 0.3632 | π → π∗ (HOMO → LUMO+1) |
DMSO | 329.34 | 0.367 | n → π∗ (HOMO-1 → LUMO) |
Water | 329.87 | 0.3716 | Intramolecular charge transfer (HOMO → LUMO) |
Table 8 shows the electronic transitions derived from TD-DFT calculations, which reveal a complex interplay between the molecule and its solvent environment. π → π∗ and n → π∗ transitions are identified, and a dominant feature appears to be intramolecular charge transfer (ICT) in several polar solvents like DCM, methanol, and water. The observed red-shift in λmax and increased oscillator strengths for these transitions in polar solvents are consistent with the stabilization of a more polar excited state relative to the ground state. This highlights the sensitivity of metronidazole’s electronic structure to solvent polarity, significantly influencing its absorption characteristics.
Ligand | Protein | Affinity (kcal mol−1) | RMSD (l.b) | RMSD (u.b) |
---|---|---|---|---|
Metronidazole | 8fb0 | −4.3 | 1.203 | 2.014 |
Metronidazole | 4kov | −5.3 | 1.721 | 2.304 |
Metronidazole | 5j62 | −4.8 | 1.815 | 2.139 |
Metronidazole | 3q5p | −4.5 | 1.929 | 2.206 |
Table 10 shows the interactions between metronidazole and four antibacterial protein targets. These interactions involve amino acid residues such as ARG45, HIS97, TYR68, and PHE89, which contribute to stabilizing the ligand–protein complexes. Notably, conventional hydrogen bonds were observed with bond lengths typically ranging between 2.0–3.0 Å, suggesting strong binding affinities, while interactions like π–π stacking and π–alkyl contacts indicate secondary stabilization. These findings emphasize both common (e.g., hydrogen bonding) and less common (e.g., amide–π stacking, π–σ) interactions, which are important for understanding ligand orientation, selectivity, and overall binding strength. These findings suggest that both hydrogen bonding and hydrophobic interactions are crucial for metronidazole’s effectiveness against bacterial targets.
Ligand | Protein ID | Amino acid | Distance | Category | Types |
---|---|---|---|---|---|
Metronidazole | 8fb0 | ARG45 | 2.209 | Hydrogen bond | Conventional hydrogen bond |
ARG45 | 3.193 | Hydrogen bond | Conventional hydrogen bond | ||
HIS97 | 2.850 | Hydrogen bond | Conventional hydrogen bond | ||
ASP44 | 2.062 | Hydrogen bond | Conventional hydrogen bond | ||
ASP44 | 4.369 | Hydrophobic | Amide-pi stacked | ||
ARG45 | 5.278 | Hydrophobic | Pi-alkyl | ||
Metronidazole | 4kov | TYR68 | 2.882 | Hydrogen bond | Conventional hydrogen bond |
ASN80 | 2.058 | Hydrogen bond | Conventional hydrogen bond | ||
CYS29 | 2.694 | Hydrogen bond | Conventional hydrogen bond | ||
CYS29 | 2.703 | Hydrogen bond | Conventional hydrogen bond | ||
GLY79 | 3.583 | Hydrogen bond | Carbon hydrogen bond | ||
SER116 | 3.691 | Hydrogen bond | Carbon hydrogen bond | ||
TYR68 | 3.290 | Hydrogen bond | Pi-donor hydrogen bond | ||
TYR68 | 5.015 | Hydrophobic | Pi–pi stacked | ||
PRO31 | 4.465 | Hydrophobic | Pi-alkyl | ||
Metronidazole | 5j62 | LEU88 | 2.98767 | Hydrogen bond | Conventional hydrogen bond |
PHE89 | 2.19381 | Hydrogen bond | Conventional hydrogen bond | ||
GLN60 | 2.00908 | Hydrogen bond | Conventional hydrogen bond | ||
LYS87 | 3.55147 | Hydrophobic | Pi-sigma | ||
LYS64 | 4.0364 | Hydrophobic | Pi-alkyl | ||
Metronidazole | 3q5p | TYR152 | 2.57263 | Hydrogen bond | Conventional hydrogen bond |
GLU253 | 2.59958 | Hydrogen bond | Conventional hydrogen bond | ||
TAL148 | 2.13857 | Hydrogen bond | Conventional hydrogen bond | ||
VAL147 | 3.55312 | Hydrophobic | Pi-sigma |
Furthermore, Fig. 10–13 illustrate the nonbonding interactions between metronidazole and 8fb0, 4kov, 6ko5, and 3q5p, respectively.
As shown in Fig. 10, metronidazole interacts with the 8fb0 protein through a network of hydrogen bonds and hydrophobic contacts. Notably, conventional hydrogen bonds are observed with residues such as Ser116 and Cys29, while π-alkyl interactions are formed with Tyr68, Pro31, and Asn80, which likely contribute to the structural stability of the complex. The surface topology in the 3D map highlights regions where donor and acceptor interactions are spatially localized, supporting a favorable binding conformation.
![]() | ||
Fig. 10 3D (a) and 2D (b) map of nonbonding interactions between the 8fb0 protein and metronidazole. |
Fig. 11 displays the interaction profile of metronidazole with the 4kov protein. Strong conventional hydrogen bonds are observed with residues Asp44, His97, and Arg45, suggesting these residues play a vital role in ligand recognition. Additionally, a π–alkyl interaction with Phe43 may reinforce ligand binding via van der Waals forces. The dense surface contact observed in the 3D image reflects a well-fitted binding pocket with high complementarity.
![]() | ||
Fig. 11 3D (a) and 2D (b) map of the nonbonding interactions between the 4kov protein and metronidazole. |
Fig. 12 presents the nonbonding interactions between metronidazole and the 5j62 protein. As shown, the ligand forms conventional hydrogen bonds with Gln60 and Leu88, and π-alkyl interactions with Lys64 and Lys87, which stabilize the complex. These interactions suggest that metronidazole is well-accommodated in the binding pocket, contributing to its binding affinity through hydrogen bonding and hydrophobic contacts.
![]() | ||
Fig. 12 3D (a) and 2D (b) map of nonbonding interactions between the 5j62 protein and metronidazole. |
Fig. 13 shows the critical nonbonding interactions between the 3q5p protein and metronidazole. The 3D representation (a) clearly depicts a strong hydrogen bond with Glutamine 253 (Glu253) and a stabilizing π–σ interaction with Valine 147 (Val147). The 2D diagram (b) reinforces these findings and reveals an additional conventional hydrogen bond with Tyrosine 152 (Tyr152). These precise molecular contacts collectively underscore a robust and specific binding affinity. This detailed mapping of interactions is essential for elucidating the drug’s mechanism and potential implications for its biological activity.
Ligands | Protein | |||
---|---|---|---|---|
8fb0 | 4kov | 5j62 | 3q5p | |
Metronidazole + secnidazole | −4.8 | −6.8 | −5.5 | −5.6 |
Metronidazole + tizoxanide | −8.0 | −8.7 | −8.0 | −7.1 |
Metronidazole + caffeine | −6.0 | −6.7 | −6.5 | −6.1 |
Metronidazole + secnidazole + tizoxanide | −8.1 | −8.7 | −8.3 | −7.1 |
Metronidazole + secnidazole + tizoxanide + caffeine | −8.2 | −8.8 | −8.1 | −7.0 |
This journal is © The Royal Society of Chemistry 2025 |