Molecular docking and computational assessment of spectroscopic analysis of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate as a potential antibacterial agent

Wissam Habibia, Saadia Ouizatb, Mohamed Chellegui*cd, Bushra Shakoore, Marwa Alaqarbehf, Mohamed Adel Sayedg, Mostafa Khouilia, Abdessamad Tounsib, Haydar A. Mohammad-Salimhi and Mohamed Anouar Harradj
aMolecular chemistry, Materials and Catalysis laboratory, Faculty of Sciences and technology, Sultan Moulay Slimane University, Beni-Mellal 23000, Morocco
bEnvironmental, Ecological, and Agro-Industrial Engineering Laboratory, Sultan Moulay Slimane University, Beni-mellal 23000, Morocco
cLaboratory of Organic Chemistry (LR17ES08), Faculty of Sciences, University of Sfax, 3038 Sfax, Tunisia. E-mail: mohamed.chellegui.etud@fss.usf.tn
dNamur Institute of Structured Matter, University of Namur, Rue de Bruxelles, 61, B-5000 Namur, Belgium
eSynthetic and Natural Products Drug Discovery Lab., Department of Chemistry, Government College Women University, Faisalabad, 38000, Pakistan
fDepartment of Chemistry, Faculty of Science, Applied Science Private University, Amman, 11931, Jordan
gPharmaceutical Chemistry Department, Faculty of Pharmacy, Egyptian Russian University, Badr City, Cairo 11829, Egypt
hDepartment of Chemistry, Faculty of Science, University of Zakho, Zakho 42002, Kurdistan Region, Iraq
iTCCG Lab, Scientific Research Center, University of Zakho, Zakho 42002, Kurdistan Region, Iraq
jInterdisciplinary Research Laboratory in Bioresources, Environment and Materials (LIRBEM) - École Normale Supérieure - Cadi Ayyad University, Morocco

Received 30th April 2025 , Accepted 11th July 2025

First published on 11th August 2025


Abstract

A novel enamino ester, ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3, was synthesized and structurally characterized. Density functional theory (DFT) calculations at the B3LYP-D3/6-311++G(d,p) level were used to investigate its molecular geometry and predict NMR chemical shifts, which showed good agreement with experimental data. Electronic properties, including the HOMO–LUMO gap, Fukui functions, and molecular electrostatic potential (MEP) surface, indicated both chemical stability and potential reactive sites. Molecular docking analysis revealed promising antibacterial activity, with compound 3 showing favorable binding to Bacillus subtilis laccase, E. coli DNA gyrase B, and key residues in Staphylococcus aureus. ADME-Tox predictions confirmed acceptable pharmacokinetic properties and low toxicity, notably with no significant hERG inhibition or neurotoxicity risk. Overall, compound 3 emerges as a stable and bioactive molecule with potential for further antibacterial drug development.


1. Introduction

Chemical compounds known as enaminones consist of an amino group linked to a carbonyl group by a C[double bond, length as m-dash]C double bond. Enamine chemistry is an emerging area of organic synthesis. These ethylenes are push–pull in nature, with the amino group pushing the electron density because it is nucleophilic and the carbonyl attracting it because it is electrophilic. Heterogeneous catalysis, a method within green chemistry, facilitates the synthesis of enamines, which serve as versatile chemical intermediates.1 The use of these molecules as building blocks in chemical synthesis has attracted attention in recent years.2,3

These compounds demonstrate considerable reactivity in diverse reduction, oxidation, photochemical processes, and nucleophilic and electrophilic substitutions.4,5 Moreover, they have served as precursors for numerous pharmaceutical compounds exhibiting anti-epileptic,6 antibacterial,7 anticonvulsant,8 anticancer,9 and antiparasitic activities,10 along with physiologically and therapeutically active compounds.11,12

Minimal research has been conducted on quantum chemical calculations and their correlation with the experimental properties of enamines, as indicated by the bibliographic study.13–18 Enaminones, including α,β-unsaturated esters bearing an amino substituent at the β-position such as ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate, represent a valuable class of compounds with diverse applications in heterocyclic chemistry and drug design.19–21 This work details the synthesis and characterization of an enaminoester. The attributes of the synthesized compound's molecular structure were examined using theoretical calculations based on density functional theory (DFT). DFT calculations have proven to be powerful tools for probing molecular structures, energetics, and reactivity patterns in various chemical systems. These methods provide valuable insights that complement experimental findings.22,23 Frontier molecular orbitals (FMO), including the highest occupied molecular orbital (HOMO)–lowest unoccupied molecular orbital (LUMO) energy gap, were also calculated using the DFT framework. The relative electrophilicity and nucleophilicity of the investigated compound was evaluated using the condensed Fukui function and the molecular electrostatic potential (MEP) surface. Molecular docking analysis was used to investigate the biological activities of the compound. The study concentrated on the active bending residues linked to hydrogen bonds and the bending energy of a specific chemical that demonstrates bacterial activity. Molecular docking simulations are widely employed to investigate non-covalent interactions and to predict the preferred binding orientation of small molecules within host systems or active sites. These approaches are particularly useful for exploring recognition processes and complement quantum chemical methods in studying supramolecular assemblies.24,25 The physico-chemical properties and ADME-Tox predictions of the prospective bioactive compound were assessed. We also find inspiration in the recent developments in the application of enaminones and β-enamino esters in drug discovery, specifically antimicrobial activity. Previous studies have demonstrated the biological relevance of structurally congruent scaffolds to antibacterial, anticancer, and anti-inflammatory applications.19–21 Taking advantage of these findings, we synthesized and characterized a novel β-enamino ester and employed a combined theoretical and molecular docking approach for the investigation of its chemical reactivity and antibacterial properties. This integrated approach via DFT, Fukui, MEP, and docking methods is a comprehensive assessment of the reactivity and biological relevance of the compound and is consistent with more recent literature on computational drug design.22–26

2. Experimental methods

2.1. Materials

Analytical grade compounds, such as ethyl 3-oxo-3-phenylpropanoate and 4-chloro-2-methyl-aniline, were acquired from Janssen Chemica and Sigma-Aldrich.

2.2. Instrumentation

A Bruker Advanced 300 spectrometer at 300 MHz and 75 MHz in CdCl3 solvent was used to obtain 1H and 13C NMR spectra. Banc Kofler apparatus was used to determine the melting points. Fourier-transform infrared (FTIR) spectra were obtained utilizing a Nicolet is 5 Thermo scientific spectrometer. Pellets were composed of 100 mg of finely powdered KBr and 2 mg of the sample.

2.3. Synthesis of the β-enaminoester

15 mg of ZnAl2O4@ZnO was added to a magnetically stirred mixture of ethyl 3-oxo-3-phenylpropanoate (0.32 g; 1.7 mmol) and 4-chloro-2-methyl-aniline (0.23 g; 1.7 mmol) and the resulting solution was agitated under ambient conditions for 20 minutes. Upon completion of the reaction, 10 mL of distilled water was introduced to the residue and subsequently extracted with diethyl ether (3 × 25 mL). The organic layer was dehydrated with Na2SO4, and the solvent was evaporated at reduced pressure. The pure β-enamino ester was isolated using column chromatography on silica gel, employing a hexane/ethyl acetate solvent system. With a yield of 82%, the purified product was obtained as a bright colourless powder.

M.p. 116–118 °C. 1H NMR (CDCl3, 300 MHz) δ/ppm: 1.13 (t, J = 6.6 Hz, 3H), 1.16–1.22 (m, 6H), 3.07–3.17 (m, 3H), 3.98 (q, J = 6.6 Hz, 2H), 4.53 (s, 1H), 7.01–7.15 (m, 2H, Ar), 9.73 (br s, 1H, NH). 13C NMR (75 MHz, CDCl3): δ 12.31, 15.64, 17.65, 56.25, 83.12, 123.56, 133.74, 133.49, 133.73, 135.39, 157.26, 168.21.

3. Computational details

3.1. Density functional theory (DFT) calculations

The quantum calculations were carried out using density functional theory (DFT) with the Gaussian 16 software.27 The geometric optimization of the studied compound was carried out with the hybrid functional B3LYP, which combines the three-parameter exchange potential of Becke and the correlation functional of Lee, Yang and Parr,28 using the basic set 6-311++G(d,p). The D3 correction was also included to account for dispersion correction energies.29 The vibrational frequencies were calculated at the same level of calculation to verify the nature of the minimum obtained. Under the same level of calculation, the frontier molecular orbitals (FMO), in particular the highest occupied molecular orbital (HOMO) and the lowest unoccupied orbital (LUMO), as well as the molecular electrostatic potential (MEP), were calculated and identified.

Following the approach of conceptual DFT (CDFT),30–32 the global chemical reactivity indices, such as the global hardness (η),33 the global softness (S), the absolute electronegativity (χ), the global electrophilicity (ω)34 and the nucleophilicity (N),35 have been calculated from the energies of the EHOMO and ELUMO boundary orbitals obtained, using the equations:

Hardness

 
η = ELUMOEHOMO/2 (1)
Softness
 
S = 1/η (2)
Electronegativity
 
χ = −EHOMOELUMO/2 (3)
Electrophilicity index
 
ω = χ2/2η (4)
Nucleophilicity index
 
N = EmoleculeHOMOETCEHOMO (5)
TCE is the tetracyano-ethylene reference.

In addition, the electrophilic P+k(r) and nucleophilic Pk(r) Parr functions35,36 were calculated by analyzing the Mulliken atomic spin density for the anionic and cationic forms of the compound studied, following the method proposed by Domingo.37 The local reactivity indices derived from these functions make it possible to identify the atoms or the molecular sites likely to behave as donors or acceptors of electrons.38 These parameters provide essential information to better understand the chemical reactivity of the different regions of the molecule.

Nuclear magnetic resonance chemical shifts have been calculated with the gauge including the atomic orbital (GIAO) approach and compared with experimental spectra to determine the chemically significant area.

Gauss view 6.0 software39 is used to obtain the highest occupied and lowest unoccupied molecular orbital maps (HOMO–LUMO), band energy gap and molecular electrostatic potential map (MEP) for identifying the potential region.

Non-covalent interaction (NCI) analysis, electron localization function (ELF), localized orbital locator (LOL), and Fukui function were carried out by Multiwfn,40 which is a multifunctional wave function analysis program and all isosurface maps were rendered by the VMD program.41

3.2. Molecular docking simulations

Molecular operating environment (MOE) 2019.010242,43 was used for the preparation of both protein and ligands, molecular docking, and evaluation of ligand–protein interaction through visualization of poses and scoring function. The docking process was performed via the Amber10 protocol using docking placement: triangular matcher, rescoring: London dG, forcefield and refinement: affinity dG.

Copper-dependent laccase (CotA) [PDB id: 1of0], DNA gyrase B [PDB id: 6kzv] and wild-type Staph. aureus penicillin binding protein 4 (PBP4) [PDB id: 5tw8], are considered as important bacterial proteins in Bacillus subtilis, E. coli and Staph. aureus that play crucial roles in antimicrobial drug design through evolving research.44,45 These isozymes were retrieved from the protein data bank (https://www.rcsb.org) with Bacillus subtilis [PDB id: 1of0], E. coli [PDB id: 6kzv] and Staph. aureus [PDB id: 5tw8]. The protein structure was checked and repaired through automatic correction and fixation order of MOE. Hydrogen atoms were added to the structure during the protonation step. The co-crystalized ligands were used to determine the binding sites. Then, the co-crystallized water molecules and bound ligands were removed. After docking completion, observation and filtration of the results through scoring values and visualized poses were performed.

MOE was used to prepare a 3D model library from the active and selective target compound (A). This compound was subjected to an energy minimization process and automatic calculation of the partial charges. Finally, this prepared library was saved in the form of an MDB file to be used in the docking calculations with target isozymes.

Validation of the docking protocol is performed via calculation of the root mean square deviation (RMSD). The RMSD is predicted via redocking of the co-crystalized ligand on its target enzyme then superimposing the redocked co-crystalized ligand on its original co-crystallized bound conformation. During this study, the RMSD of copper-dependent laccase (CotA) [PDB id: 1of0], DNA gyrase B [PDB id: 6kzv] and wild-type Staph. aureus penicillin binding protein 4 (PBP4) [PDB id: 5tw8], appeared in the acceptable range with values equal to 1.7295, 1.6839 and 1.2139 Å, respectively.

3.3. Molecular dynamic simulations

Compound A/(PBP4) of Staph. aureus complex was subjected to molecular dynamic analysis study. The MD simulations were carried out by iMod server (iMODS) (https://imods.chaconlab.org/) which gives a convenient interface for this enhanced normal mode analysis (NMA) methodology in inner coordinates46 at 300 K constant temperature and 1 atm constant pressure. Finally, 50 ns molecular dynamic simulation was carried out for the target complex.

3.4. ADMET LAB 2.0 – drug likeness and ADME studies

ADMET LAB 2.0 platform (https://admetmesh.scbdd.com/) is a freely accessible web tool that has gathered the most relevant computational techniques to provide a global estimation of the pharmacokinetics profile of small molecules. Their methodologies were chosen by the web tool creators for their robustness, as well as their ease of interpretation, to allow effective translation to medicinal chemistry. Some of these approaches were updated by web tool designers utilising open-source algorithms, while others were unmodified versions of the original authors’ methods.47 The molecular structure of the target compound A was uploaded into the ADMET LAB 2.0 web tool section using the simplified molecular-input line-entry specification (SMILES) nomenclature technique using Marvin sketch software 19.19 and then the result report was generated.

4. Results and discussion

The β-enaminoester was synthesized by condensation reaction of an amine with a dicarbonyl in the presence of ZnAl2O4@ZnO as a catalyst according to Scheme 1. The reaction was carried out solvent-free at room temperature with good yield. The compound obtained underwent analysis using NMR.
image file: d5nj01851j-s1.tif
Scheme 1 Synthesis of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3.

4.1. Synthesis of the title compound

The synthesis of the title compound was carried out in accordance with the current study concentrated on the theoretical and structural analysis of ethyl 3-(4-chloro-2-methylanilino)-3-phenylpropanoate 3. The title compound was synthesized according to the experimental procedures outlined in Scheme 1. Briefly, a stoichiometric quantity of ethyl 3-oxo-3-phenylpropanoate 1 and 2,4,6-trimethyl-phenylamine 2 was mixed in the presence of ZnAl2O4@ZnO at room temperature. The reaction completion was followed by TLC. At the end of the reaction followed by workup steps and purification by column chromatography, the desired product was isolated as a colourless powder with a yield of 82%.

Scheme attempts to illustrate the proposed mechanism for this reaction. According to this hypothesis, the reaction begins with amines bearing electron-donating substituents and possessing nucleophilic properties. Initially, the amine stabilizes the cationic intermediate, which is then converted into an intermediate, continuing until the final regeneration of the catalyst (Scheme 2).


image file: d5nj01851j-s2.tif
Scheme 2 Proposed reaction mechanism for the formation of the β-enaminoester.

4.2. Spectroscopic studies

Nuclear magnetic resonance (NMR) spectroscopy is a physicochemical technique for determining the structural properties of an organic compound. DFT calculations have been demonstrated to be a useful method for predicting NMR spectra and investigating the relationship between molecular structure and chemical shifts (Table 1). As a result, the use of experimental and theoretical methods allows for the evaluation of molecule structure.
Table 1 Experimental and B3LYP/6-311+G(2d,p) computed 1H-13C NMR of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3
Atoms Exp. value δ (ppm) Calc. value δ (ppm) Atoms Exp. value δ (ppm) Calc. value δ (ppm)
H1 10.15 10.29 C1 158.92 164.52
H2 5.01 4.88 C2 91.94 95.94
H4a 4.2 4.92 C3 169.89 174.33
H4b 4.13 3.6 C4 59.13 61.45
H5a 1.32 1.56 C5 15.61 16.5
H5b 1.29 0.81 C6 135.93 142.4
H5c 1.27 1.09 C7 127.43 131.99
H7 7.27 7.63 C8 128.16 132.6
H8 7.24 7.49 C9 128.69 133.31
H9 7.34 7.4 C10 127.95 131.56
H10 7.24 7.18 C11 128.25 133.12
H11 7.25 7.08 C12 137.69 142.96
H13 6.16 6.04 C13 124.39 126.71
H14 7.07 6.58 C14 126.34 129.49
H16 7.22 7.07 C15 127.43 140.58
H18a 2.38 1.94 C16 128.45 133.78
H18b   2.44 C17 129.45 136.06
H18c   2.37 C18 18.13 19.52


The experimental 1H NMR spectrum of the enamino ester (Scheme 3) shows aromatic protons between 6.15 and 7.33 ppm, calculated in the range of 6.03–7.62 ppm in chloroform (Fig. S1, ESI). We also note the presence of a signal in the form of a singlet at 5 ppm, due to the proton of the methylene group, and a quadruplet at 4.20–4.13 ppm, due to the (O–CH2–CH3) protons. In addition, a singlet at 10.15 ppm is assigned to the (N–H) proton of the isoxazoline ring (Fig. S1, ESI). The calculated chemical shift values of the (H1–N– (10.29 ppm)), (O–CH2–CH3 (3.59–4.91 ppm)) and (C[double bond, length as m-dash]CH (4.88 ppm)) protons were obtained using the DFT method and were in good agreement with the experimental values (Table 1).


image file: d5nj01851j-s3.tif
Scheme 3 Structure of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3.

For 13C NMR (Fig. S2 and S3, ESI), the chemical shifts calculated by the DFT method show excellent agreement with the experimental values. The theoretical spectrum of the aromatic carbons is observed between 126 and 142 ppm, while in the experimental spectrum it is detected between 124 and 135 ppm. Chemical shifts for the CH carbon of the isoxazoline ring were observed at 91 ppm, which can be explained by the attractive effect of the oxygen in the isoxazoline ring. Furthermore, the signal at 59 ppm corresponds to the CH2 group, whereas the calculated value is 61 ppm. These results are in good agreement with the experimental results. We also noted the presence of two signals at 156 ppm and 177 ppm, which indicate that they correspond to the C[double bond, length as m-dash]N carbon of the isoxazoline ring and to the carbonyl (C[double bond, length as m-dash]O), while the calculated values are observed at 164 ppm and 174 ppm, respectively. The FT-IR spectrum of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3, presented in Fig. S4 (ESI), exhibits a broad absorption band of weak intensity at approximately 3300 cm−1, which can be assigned to the N–H stretching vibration. A strong absorption band at 1541 cm−1 is attributed to the stretching vibration of the C[double bond, length as m-dash]C bond. The aromatic C–H stretching modes are observed at 3050 cm−1, while the N–C stretching vibration is identified at 1294 cm−1.

4.3. DFT-based computational chemistry approach

4.3.1. Frontier molecular orbital analysis. The highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO), called frontier molecular orbitals (FMOs), play a crucial role in quantum chemistry, helping us to understand the reactivities of a large number of organic molecules.

The FMOs make it possible to explain the electronic properties by studying the behavior of the electrons, in particular during an excitation from the HOMO to the LUMO. The HOMO energy (EHOMO) tells us how easily a molecule can donate electrons, while the LUMO energy (ELUMO) reveals its ability to accept them.

In Fig. 1, the FMOs of the studied compound are represented. The green and red regions show the spatial distribution of the molecular phases, with green representing negative phases and red representing positive ones. This visualization helps us see where electrons are likely to move and interact.


image file: d5nj01851j-f1.tif
Fig. 1 Frontier molecular orbitals (HOMO and LUMO) of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 in the ground state, calculated at the B3LYP-D3/6-311++G(d,p) level.

According to this figure, it can be seen that the HOMO involves the π electrons of the phenyl ring, and the non-binding doublets of the chlorine atom, as well as those of the amine and oxygen groups of the ester function. These sites are particularly reactive in intermolecular interactions, where the molecule can play a role as an electron donor. Moreover, the molecule can also act as an electron acceptor by mobilizing its anti-binding orbitals, which constitute the LUMO.

The analysis of frontier molecular orbital levels provides precise insights into the molecule's ability to act as either an electron donor or acceptor (Table 2).

Table 2 Electronic properties (in eV) of the ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 calculated at the B3LYP-D3/6-311++G(d,p) level of approximation
ELUMO EHOMO Egap χ η S ω N
−1.67 −6.01 4.34 3.84 2.17 0.46 3.40 3.49


The value of the energy of the HOMO (−6.01 eV) indicates that the molecule has a low tendency to donate electrons, which makes it a relatively inefficient electron donor. On the other hand, the energy of the LUMO (−1.67 eV) shows a moderate ability to accept electrons. The relatively wide energy gap (4.34 eV) reflects good chemical stability as well as moderate reactivity under standard conditions. These results suggest that the molecule presents a balance between the behaviors of donor and acceptor of electrons, without manifesting a marked preference for one or the other.

With a dipole moment of μ = 2.15 D, the molecule has an intermediate polarity, which could influence its molecular interactions, via dipole–dipole interactions. Although this polarity is less marked than that of highly polar molecules, it allows the molecule to interact effectively in polar environments while maintaining moderate chemical stability (Fig. 2).


image file: d5nj01851j-f2.tif
Fig. 2 B3LYP-D3/6-311++G(d,p) optimized geometry and dipole moment vector of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3.

With an electronegativity of χ = 3.84 eV, this molecule shows a strong ability to attract electrons, which suggests a tendency to act as an electron acceptor in chemical reactions. Its hardness (η = 2.17 eV) indicates a moderate reactivity, allowing it to participate in various chemical processes. Finally, the electrophilicity index (ω = 3.40 eV) confirms its significant ability to accept electrons, reinforcing its potential role in chemical interactions.

The nucleophilicity index (N = 3.49 eV) highlights the ability of the molecule to participate in electron transfers while maintaining a balanced reactivity profile. These quantum indexes show that the molecule has a moderate and versatile reactivity, capable of acting both as an electron donor and acceptor. Its dipole moment, combined with its electronegativity and its electrophilicity index, suggests an ability to engage in chemical reactions while maintaining a notable stability. This makes it a potentially interesting candidate for applications where electronic interactions play a key role.

The DOS plot shows the distribution of electronic states and their contribution to molecular orbitals. A high density of states around the HOMO–LUMO gap reveals the delocalization of π-electrons, enhancing the ability of the title molecule to communicate electronically within biological systems. From the peaks at the HOMO and LUMO energy levels, strong electronic interactions are exhibited; the contributions arise majorly from the aromatic rings and a carbonyl group (Fig. 3).


image file: d5nj01851j-f3.tif
Fig. 3 B3LYP-D3/6-311++G(d,p) density of states (DOS) plot of the title molecule.
4.3.2. Analysis of the MEP. The result obtained by the analysis of the FMOs is confirmed by the molecular electrostatic potential (MEP) map. In the end, the MEP map allows for the visualization of regions in the molecule that are likely to participate in electrostatic interactions with other chemical species, such as electrophiles or nucleophiles.9

According to Fig. 4, the red-colored areas on the MEP map corresponding to regions with a negative electrostatic potential (−0.07131), are associated with the non-binding doublets of the oxygen of the ester group, and the non-binding doublet of the chlorine atom, as well as with the p electrons of the phenyl ring. The oxygen atom O5 and its neighboring atoms constitute electron-rich centers, which makes them likely to act as electron donors during the formation of new bonds during chemical reactions. On the other hand, the regions of low electron density, characterized by a positive electrostatic potential (0.07131), correspond to the hydrogen atoms bonded to the carbon atoms. A particularly marked contribution comes from the hydrogen of the amine group. These zones are characteristic of atoms capable of accepting electrons, indicating significant electrophilic reactivity.


image file: d5nj01851j-f4.tif
Fig. 4 B3LYP-D3/6-311++G(d,p) electrostatic potential distribution in ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3.
4.3.3. Localization of reactive sites: approach using Parr functions. To analyze in detail the chemical reactivity of our molecule and identify its active sites, we have also determined the local reactivity indices, such as the electrophilic P+k(r) and nucleophilic Pk(r) Parr functions. These indices are essential to identify the atoms or molecular sites likely to act as donors or acceptors of electrons.11

The P+k(r) measures the ability of a site to donate electrons, thus revealing its nucleophilic potential (Fig. 5). Conversely, the Pk(r) evaluates its ability to accept electrons. High values for these indices make it possible to locate the sites suitable for electrophilic or nucleophilic reactions, which facilitates a thorough understanding of the reaction mechanisms at the molecular scale (Fig. 5).


image file: d5nj01851j-f5.tif
Fig. 5 B3LYP-D3/6-311++G(d,p) electrophilic and nucleophilic Parr functions (η = 0.02) of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3.

According to Fig. 5 representing the electron distribution of the Parr functions, the C32 atom appears as the most electrophilic site, with a value of 0.19, revealing a strong capacity to accept electrons. Other notable electrophilic sites include the C1, C17 and C27, with values of 0.16, 0.11 and 0.16, respectively, suggesting that they are also conducive to electrophilic interactions. These results indicate that these carbon atoms are the most reactive to interact with electron-rich centers. Moreover, the Pk(r) functions show that the C2 atom has a high value of 0.54, which makes it a very active electron donor site. The O4 and O5 oxygen atoms present a very low Pk(r) value of 0.04 and 0.08, respectively. The N6 and C22 exhibit a Parr(k) value of 0.31 and 0.11, respectively.

To validate the identification of the reactive sites, we compared results from three local reactivity descriptors: condensed Fukui functions, Parr functions, and the molecular electrostatic potential (MEP) map (see Section 4.3.2). All methods consistently indicated that atoms such as C32 and C2 are key electrophilic and nucleophilic centers, respectively, thereby reinforcing the accuracy and reliability of our reactivity predictions.

4.3.4. Reduced density gradient analysis. Electron density (ED) and its derivatives can be used to highlight weak interactions, such as intermolecular and even covalent interactions in real space, using the reduced density gradient (RDG).

The results were calculated by Multiwfn and the VMD program.48,49 RDG analysis of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 by Multiwfn gives a visual representation of non-covalent interactions and strong repulsion (Fig. 6). While green indicates a van der Waals interaction, the dashed red hues suggest repulsive contacts (steric effect) and blue indicates attractive interactions (hydrogen bonding). Repulsive interactions were prototyped on the ester group indicating a significant steric influence in the molecule (Fig. 6). van der Waals interactions were found between the oxygen and hydrogen atom of the amine. The ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 exhibits strong attractive interactions, as shown by its blue colour (Fig. 6).


image file: d5nj01851j-f6.tif
Fig. 6 Binding configuration of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 and corresponding interaction analysis. The 2D interaction diagram reveals hydrogen bonding (green dashed lines) and hydrophobic interactions (red curved lines).
4.3.5. AIM topological analyses of the electron density of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3. AIM topological analysis of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 provides information about the nature of intramolecular interactions within the molecule. The electron density (ρ) at the bond critical point (BCP) was found to be 0.0314 a.u., and the Laplacian of the electron density (∇2ρ) was calculated to be 0.112 a.u. These are indicative of the presence of a weak, non-covalent interaction (a hydrogen bond) between the oxygen and hydrogen atoms of the molecule. The positive Laplacian confirms the closed-shell nature of this interaction, typical of hydrogen bonding rather than covalent bonding. These interactions stabilize the molecule's three-dimensional structure and may be relevant to its general reactivity and biological activity. The ensuing QTAIM molecular graph (see Fig. 7) makes these topological properties visually apparent, affording greater insight into the electronic environment governing the behavior of the molecule in chemical and biological systems.
image file: d5nj01851j-f7.tif
Fig. 7 QTAIM molecular graphs of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3.
4.3.6. ELF and LOL analysis. The topological analyses of the electron localization function (ELF) and localized orbital locator (LOL) are tools used for performing covalent bonding analysis. ELF and LOL reveal a high probability of finding an electron pair on the molecular surface. The colour map and relief map of ELF and LOL are produced by Multiwfn software. The ELF and LOL are shown in Fig. 8, where the relief map with a broad or contracted peak area delineates the electron environment around each atom.
image file: d5nj01851j-f8.tif
Fig. 8 Relief map with projection of the electron localization function of the title compound of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3.

ELF explains the electron pair density while LOL explains the maximum overlapping localised orbitals due to the orbital gradient.50,51 The ELF map is designed in the range 0.0 to 1.0; however, the region below 0.5 shows a delocalised electron region and LOL explains the overlap of the most localised orbitals and LOL reaches large values >0.5 in regions where electron density is dominated by electron localisation.

The colour shades of the ELF and LOL maps confirm the existence of bonding and non-bonding electrons, while the red colour around the hydrogen atoms, with a maximum value, indicates the existence of bonding and non-bonding electrons.

The high ELF or LOL values indicated by the red colour around the hydrogen atoms show a strong localisation of electrons due to the existence of a covalent bond, a lone pair of electrons or a nuclear shell in the hydrogen atom. The blue colour around the carbon atoms in the phenyl and pyridine rings evokes a cloud of delocalised electrons around them. While the blue circles around the nitrogen and pyridine atoms evoke a cloud of delocalised electrons around them.

The blue circles around the nitrogen and chlorine atoms indicate the region of electron depletion between the inner and valence layers, the inner layer and the valence layer. The central region of the hydrogen atom in the LOL is white because the electron density exceeds the upper limit of the colour scale.

4.4. Molecular docking simulations

The target ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 was selected as a hit effective antibacterial inhibitor that revealed predicted inhibition for Bacillus subtilis, E. coli and Staph. aureus via docking against an important isozyme, named copper-dependent laccase (CotA) [PDB id: 1of0], DNA gyrase B [PDB id: 6kzv] and wild-type Staph. aureus penicillin binding protein 4 (PBP4) [PDB id: 5tw8], respectively. Molecular operating environment 2019.0102. was used for the preparation steps of either the enzymes or target compound and the docking process of the previously mentioned compound into the active site of the microbial proteins.

Concerning Bacillus subtilis isozyme, compound 3 revealed a probable effective antibacterial inhibition for the copper-dependent laccase. It showed a binding score equal to −4.5689 kcal mol−1 in addition to an interaction pose that includes three hydrophobic interactions with the key amino acid residues of the enzyme His419, Gly417 and Gly323 (Fig. 9).


image file: d5nj01851j-f9.tif
Fig. 9 (a) 2-D and (b) 3-D interaction docking poses for ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 Bacillus subtilis isozyme copper-dependent laccase (CotA).

Regarding E. coli's DNA gyrase B isozyme, compound 3 generally escalated as a potent antibacterial hit. It showed better binding score (−6.4421 kcal mol−1) to the target protein (Table 3). Moreover, it succeeded to bind Glu50, Asp73 and Asn46 via three strong hydrogen bonds in addition to two hydrophobic interactions with Thr165 and Asn46, key residues of the target protein, respectively (Fig. 10 and Table 3).

Table 3 Docking scores and featured interaction of the target compounds against Bacillus subtilis, E. coli and Staph. aureus isozymes (CotA, DNA gyrase B and PBP4)
Compound Microorganism/PDB id S-score (kcal mol−1) Involved receptor residues Type of bond interaction
3 (B. subtilis)/1of0 −4.5689 His419 Hydrophobic
Gly417 Hydrophobic
Gly323 Hydrophobic
(E. coli)/6kzv −6.4421 Glu50 H-bonding
Asp73 H-bonding
Asn46 H-bonding
Thr165 Hydrophobic
Asn46 Hydrophobic
(Staph. aureus)/5tw8 −6.3545 Ser116 H-bonding
Ala182 Hydrophobic
Ser263 Hydrophobic



image file: d5nj01851j-f10.tif
Fig. 10 (a) 2-D and (b) 3-D interaction docking poses for ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 against E. coli isozyme DNA gyrase B.

Finally, regarding Staph. aureus, ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 revealed a good binding score of −6.3545 kcal mol−1 via hydrogen bonding and hydrophobic interactions with the key amino acid residues Ser116, Ala182 and Ser263 (Fig. 11 and Table 3).


image file: d5nj01851j-f11.tif
Fig. 11 (a) 2-D and (b) 3-D interaction docking poses for ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 against Staph. aureus isozyme wild-type Staph. aureus penicillin binding protein 4 (PBP4).

Fig. 12 displays the co-crystallized and docked ligand conformations with an RMSD value range from 1.60 to 1.80 Å. The very low RMSD values indicate that the binding pose prediction is accurate and supports the docking results obtained for compound 3 with its bacterial target.


image file: d5nj01851j-f12.tif
Fig. 12 Superposition and RMSD of the co-crystal (green) and re-docked (gray) structure in the active site of the targets: (A) 6kzv, (B) 1of0 and (C) 5tw8.

4.5. Molecular dynamic simulations

The outcomes of the molecular dynamics (MD) simulation and normal mode analysis (NMA) for the docked complex of the wild-type Staphylococcus aureus penicillin-binding protein 4 (PBP4) isozyme with antibacterial ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3, performed using the iMOD server (iMODS; https://imods.chaconlab.org), are presented in Fig. 13. The simulation aimed to evaluate the flexibility of protein residues in response to binding with ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3. The main-chain deformability graph (Fig. 13A) highlights regions with higher peaks, representing areas of increased protein deformability. These peaks suggest flexibility and the repositioning of amino acid residues to enhance their interaction with ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3. The experimental B-factor analysis (Fig. 13B) identifies the precise positions of atoms and accounts for crystal irregularities. Higher B-factor values correspond to greater atomic mobility and flexibility. The B-factor graph illustrates significant fluctuations in protein residues across the atomic index range of 0–2500. Notably, the ligand-enzyme complex of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 exhibited a low eigenvalue of 8.684631 × 10−4 (Fig. 13-C), indicating enhanced deformability and substantial protein flexibility. Additionally, the covariance map (Fig. 13-D) reveals strong correlations between ligand and enzyme residues. Correlated motions are denoted in red, uncorrelated motions in white, and anti-correlated motions in blue, emphasizing the dynamic interplay between the protein and ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3.
image file: d5nj01851j-f13.tif
Fig. 13 Molecular dynamics simulation of ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3/wild-type Staph. aureus penicillin binding protein 4 (PBP4) of Staph. aureus complex by iMODS server. (A) Deformability, (B) B-factor values, (C) eigenvalue and (D) covariance model.

4.6. ADMET LAB 2.0 – drug likeness and ADME studies

The predictions of ADMET properties for ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 were calculated using ADMET LAB 2.0 web tool (https://admetlab3.scbdd.com/). Compound 3 was characterized by better human intestinal absorption (HIA) results, which was supported by being accepted by the Lipinski rule. In addition, these hit compounds succeeded in passing through the blood brain barrier (BBB) indicating their suitability to treat brain bacterial infections. They did not exhibit any PAINS structural toxicity alerts. Concerning metabolism, ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 exhibited an inhibitory effect only for CYP1A2, CYP2C19 and CYP2C9. Moreover, it acts as a substrate for CYP1A2, CYP2C19 and CYP3A4 (Table 4). Regarding the toxicity profile, ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3 showed low tendency to act as a hERG blocker with no effects on the respiratory system or ability to induce any drug-neurotoxicity. The hit compound 3 failed to show a complete safety profile because of exhibiting drug-induced liver injury (DILI) ability in addition to a carcinogenicity alert, which necessitates future drug structure modification (Table 5). Finally, studying clearance and excretion of the hit molecule showed that compound 3 exhibited a clearance value equal to 3.478 mL min−1 kg−1 which is considered as a moderate value.
Table 4 The metabolic effect of compound A on CYP-enzymes
Compound Inhibitor of Substrate for
A CYP1A2, CYP2C19 and CYP2C9 CYP1A2, CYP2C19 and CYP3A4


Table 5 Drug-likeness properties of compound 3 and ADME-Tox prediction
Property Predicted value Property Prediction
MW 247.34 CYP2D6 substrate No
Num. rotatable bonds 5 CYP3A4 substrate No
Num. H-bond acceptor 3 CYP1A2 inhibitor Yes
Num. H-bond donors 1 CYP2C19 inhibitor Yes
log[thin space (1/6-em)]P (o/w) 3.72 CYPC2C9 inhibitor No
Human intestinal absorption (%) 92.986 CYP2D6 inhibitor No
VDss (log[thin space (1/6-em)]L Kg−1) 0.251 CYP3A4 inhibitor No
P-glycoprotein substrate No Real OCT2 substrate No
P-glycoprotein I, II inhibitor No AMES toxicity No
BBB permeability (log BB) −0.301 hERG I, II inhibitor No
CNS permeability (log PS) −2.128 Hepatotoxicity No
Skin permeability −2.614 Skin sensitization Yes


5. Conclusion

The synthesis and characterisation of an enamino ester, ethyl 3-(4-chloro-2-methylphenylamino)-3-phenylacrylate 3, has been detailed in this paper. The molecular structure of the synthesized compound was investigated using density functional theory (DFT) calculations at the B3LYP-D3/6-311++G(d,p) level of approximation. DFT has proven to be a reliable tool for predicting NMR spectra and exploring the relationship between molecular structure and chemical shifts. The theoretical results show good agreement with the experimental data, supporting the accuracy of the computational approach.

Additionally, the DFT framework was employed to compute molecular orbitals, including the LOMO–HUMO energy gap. The condensed Fukui function and the molecular electrostatic potential (MEP) surface were employed to assess the relative electrophilicity and nucleophilicity of the molecule being studied. The results obtained demonstrate the current molecule's ability to engage in chemical reactions while preserving significant stability, positioning it as a potentially intriguing candidate for applications where electronic interaction plays a crucial role.

The biological activities of the compound were investigated using molecular docking analysis. The study concentrated on the active bending residues linked to hydrogen bonds and the bending energy of a specific chemical that demonstrates bacterial activity. Regarding Bacillus subtilis isozyme, compound 3 demonstrated a likely effective antibacterial inhibition for the copper-dependent laccase. Concerning the compounds targeting the DNA gyrase B isozyme of E. coli, there has been a notable increase in their potency as antibacterial agents. In conclusion, Staph. aureus compound 3 demonstrated a favorable binding score through hydrogen bonding and hydrophobic interactions with the critical amino acid residues Ser116, Ala182, and Ser263.

The assessment covered the physico-chemical characteristics and the ADME-Tox forecasts of the prospective bioactive compound. In terms of the toxicity profile, compound 3 demonstrated a minimal inclination to function as an hERG blocker, exhibiting no impact on the respiratory system or capacity to induce any drug-related neurotoxicity. Ultimately, the investigation into the clearance and excretion of the target molecule revealed that compound 3 demonstrated a clearance value that is regarded as moderate.

Author contributions

Wissam Habibi, Saadia Ouizat, Mohamed Chellegui, Bushra Shakoor, Marwa Alaqarbeh, Mohamed Adel Sayed, and Mostafa Khouili: writing, investigation, validation, methodology; Abdessamad Tounsi, Haydar A. Mohammad-Salim and Mohamed Anouar Harrad: editing, reviewing, and supervision. All authors have read and approved the final version of the manuscript.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the ESI.

Acknowledgements

M. C. is grateful to UNamur for granting him the status of scientific collaborator. M. C. gratefully acknowledges the F. R. S.-FNRS for financially supporting his research stay at UNamur.

References

  1. A. Z. A. Elassar and A. A. El-Khair, Recent developments in the chemistry of enaminones, Tetrahedron, 2003, 59, 8463–8480 CrossRef CAS .
  2. L. Pan, X. Bi and Q. Liu, Recent developments of ketene dithioacetal chemistry, Chem. Soc. Rev., 2013, 42, 1251–1286 RSC .
  3. L. Zhang, J. Dong, X. Xu and Q. Liu, Chemistry of ketene N, S-acetals: an overview, Chem. Rev., 2016, 116, 287–322 CrossRef CAS .
  4. S. Nikolovaa, et al., Synth. Commun., 2013, 43, 326–336 CrossRef .
  5. N. A. Al-Awadi, M. R. Ibrahim, M. H. Elnagdi, E. John and Y. A. Ibrahim, Enaminones in a multicomponent synthesis of 4-aryldihydropyridines for potential applications in photoinduced intramolecular electron-transfer systems, Beilstein J. Org. Chem., 2012, 8, 441–447 CrossRef CAS PubMed .
  6. M. Khurana, N. N. Salama, K. R. Scott, N. N. Nemieboka, K. S. Bauer Jr and N. D. Eddington, Preclinical evaluation of the pharmacokinetics, brain uptake and metabolism of E121, an antiepileptic enaminone ester, in rats, Biopharm. Drug Dispos., 2003, 24, 397–407 CrossRef CAS PubMed .
  7. N. J. Thumar and M. P. Patel, Synthesis, characterization, and antimicrobial evaluation of carbostyril derivatives of 1H-pyrazole, Saudi. Pharm., 2011, 19, 75–83 CrossRef CAS .
  8. P. L. Jackson, C. D. Hanson, A. K. Farrell, R. J. Butcher, J. P. Stables, N. D. Eddington and K. R. Scott, Enaminones 12. An explanation of anticonvulsant activity and toxicity per Linus Pauling's clathrate hypothesis, Eur. J. Med. Chem., 2012, 5, 42–51 CrossRef .
  9. S. M. Riyadh, Enaminones as building blocks for the synthesis of substituted pyrazoles with antitumor and antimicrobial activities, Molecules, 2011, 16, 1834–1853 CrossRef CAS .
  10. A. M. El-Shennawy, A. H. Mohamed and M. Abass, Studies on parasitologic and haematologic activities of an enaminone derivative of 4-hydroxyquinolin-2 (1H)-one against murine schistosomiasis mansoni, MedGenMed, 2007, 9, 15–33 Search PubMed .
  11. N. Kaur, 5-Membered Heterocycle Synthesis Using Iodine, Elsevier, 2023 Search PubMed .
  12. J. M. Taylor, The design and synthesis of acylated enamino esters as potential inhibitors of serine proteases, 1993 Search PubMed.
  13. M. A. Harrad, A. Semane, M. Badereddine and A. Tounsi, ZnAl2O4@ ZnO an effective, heterogeneous catalyst for the synthesis of bis-(β-enaminones) and bis-(β-enaminoesters), Chem. J. Mold., 2024, 19, 74–82 CrossRef CAS .
  14. S. Atlas, K. S. Etsè, Z. V. Guillermo, M. Maatallah, A. Tounsi and M. A. Harrad, Structural and quantum study of newly synthesized methyl (Z)-3-((4-fluorophenyl) amino) but-2-enoate, Adv. J. Chem., Sect. A, 2025, 8, 809–824 CAS .
  15. M. Loughzail, K. S. Etsè, Z. V. Guillermo, R. Touzani, A. Moliterni, A. Tounsi and M. A. Harrad, Synthesis, characterization, Hirshfeld and admet estimation studies of novel 3-(2,4,6-trimethyl-phenylamino)-but-2-enoate, Chem. J. Mold., 2024, 19, 83–92 CrossRef CAS .
  16. M. A. Harrad, B. Boualy, L. El Firdoussi and M. Ait Ali, Aluminum phosphate catalyzed free solvent preparation of β-enamino esters, Am. J. Chem., 2012, 2, 271–276 CrossRef .
  17. M. A. Harrad, R. Outtouch, M. Ait Ali, L. El Firdoussi, A. Karim and A. Roucoux, Ca (CF3COO)2: An efficient Lewis acid catalyst for chemo-and regio-selective enamination of β-dicarbonyl compounds, Catal. Commun., 2010, 11, 442–446 CrossRef CAS .
  18. M. A. Harrad, I. Houssini, B. Boualy, A. Ouahrouch, M. AitAli and M. Loughzail, Natural Phosphate as New, Highly Efficient and Reusable Heterogeneous Catalyst for the Selective Preparation of Beta-Enaminoesters under Solvent-Free Conditions, Chem. Mater. Res., 2014, 6, 31–37 Search PubMed .
  19. A. H. Abdelrahman, M. E. Azab, M. A. Hegazy, A. Labena and S. K. Ramadan, Design, Synthesis, Antiproliferative Screening, and In Silico Studies of Some Pyridinyl-Pyrimidine Candidates, J. Heterocycl. Chem., 2025, 62, 303–315 CrossRef CAS .
  20. E. A. El-Helw, W. S. Abou-Elmagd, E. M. Hosni, M. Kamal, A. I. Hashem and S. K. Ramadan, Synthesis of Benzo[h]quinoline derivatives and evaluation of their insecticidal activity against Culex pipiens L. larvae, Eur. J. Med. Chem., 2025, 290, 117565 CrossRef CAS .
  21. E. A. El-Helw, M. Asran, M. E. Azab, M. H. Helal, A. Y. Alzahrani and S. K. Ramadan, Synthesis and in silico studies of certain benzo[f]quinoline-based heterocycles as antitumor agents, Sci. Rep., 2024, 14, 15522 CrossRef CAS .
  22. A. Abdou, O. A. Omran, A. Nafady and I. S. Antipin, Structural, spectroscopic, FMOs, and non-linear optical properties exploration of three thiacaix (4) arenes derivatives, Arab. J. Chem., 2022, 15, 103656 CrossRef CAS .
  23. A. M. Abu-Dief, T. El-Dabea, R. M. El-Khatib, A. Abdou, I. O. Barnawi, H. A. Alshehri, K. Al-Ghamdi and A. E. A. A. Mahmoud, Fabrication, physicochemical characterization and theoretical studies of some new mixed ligands complexes based on N-(1H-benzimidazol-2-yl)-guanidine and 1, 10-phenanthroline: DNA interaction, biological applications and molecular docking approach, J. Mol. Struct., 2024, 1310, 138328 CrossRef CAS .
  24. S. H. Rashmi, K. S. Disha, N. Sudheesh, J. Karunakaran, A. Joseph, A. Jagadesh and P. P. Mudgal, Repurposing of approved antivirals against dengue virus serotypes: an in silico and in vitro mechanistic study, Mol. Diversity, 2023, 1–14 Search PubMed .
  25. A. Abdou, H. M. Mostafa and A. M. Abdel-Mawgoud, New Fe(III) and Ni(II) azocoumarin based complexes: synthesis, characterization, DFT, antimicrobial, anti-inflammatory activity, and molecular docking analysis, Russ. J. Gen. Chem., 2023, 93, 3006–3019 CrossRef CAS .
  26. A. S. Elgubbi, E. A. El-Helw, M. S. Abousiksaka, A. Y. Alzahrani and S. K. Ramadan, β-Enaminonitrile in the synthesis of tetrahydrobenzo[b]thiophene candidates with DFT simulation, in vitro antiproliferative assessment, molecular docking, and modeling pharmacokinetics, RSC Adv., 2024, 14, 18417–18430 RSC .
  27. M. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, et al., Gaussian 16, Revision A. 03, Gaussian Inc., Wallingford, CT, 2016 Search PubMed .
  28. C. Lee, W. Yang and R. G. Parr, Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789 CrossRef CAS .
  29. S. Grimme, Density functional theory with London dispersion corrections, WIREs Comput. Mol. Sci., 2011, 1, 211–228 CrossRef CAS .
  30. R. G. Parr and W. Yang, Density-Functional Theory of Atoms and Molecules, Oxford University Press, Oxford, 1989 Search PubMed .
  31. P. Geerlings, F. De Proft and W. Langenaeker, Conceptual density functional theory, Chem. Rev., 2003, 103, 1793–1874 CrossRef CAS PubMed .
  32. L. R. Domingo, M. Ríos-Gutiérrez and P. Pérez, Applications of the conceptual density functional theory indices to organic chemistry reactivity, Molecules, 2016, 21, 748–769 CrossRef .
  33. R. G. Parr and R. G. Pearson, Absolute hardness: companion parameter to absolute electronegativity, J. Am. Chem. Soc., 1983, 105, 7512–7516 CrossRef CAS .
  34. R. G. Parr, L. V. Szentpály and S. Liu, Electrophilicity index, J. Am. Chem. Soc., 1999, 121, 1922–1924 CrossRef CAS .
  35. L. R. Domingo, E. Chamorro and P. Pérez, Understanding the reactivity of captodative ethylenes in polar cycloaddition reactions. A theoretical study, J. Org. Chem., 2008, 73, 4615–4624 CrossRef CAS PubMed .
  36. L. R. Domingo, P. Pérez and J. A. Sáez, Understanding the local reactivity in polar organic reactions through, RSC Adv., 2013, 3, 1486–1494 RSC .
  37. L. R. Domingo and J. A. Sáez, Understanding the mechanism of polar Diels–Alder reactions, Org. Biomol. Chem., 2009, 7, 3576–3583 RSC .
  38. P. K. Chattaraj, B. Maiti and U. Sarkar, Philicity: a unified treatment of chemical reactivity and selectivity, J. Phys. Chem. A, 2003, 107, 4973–4975 CrossRef CAS .
  39. J. M. Millam, Gauss View5.0.8, Gaussian Inc, Wallingford, 2016 Search PubMed .
  40. T. Lu and F. Chen, Multiwfn: A multifunctional wavefunction analyzer, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed .
  41. J. Eargle and Z. Luthey-Schulten, NetworkView: 3D display and analysis of protein RNA interaction networks, Bioinformatics, 2012, 28, 3000–3001 CrossRef CAS .
  42. M. A. Said, A. Albohy, M. A. Abdelrahman and H. S. Ibarhim, Remdesivir analog as SARS-CoV-2 polymerase inhibitor: virtual screening of a database generated by scaffold replacement, RSC Adv., 2012, 12, 22448–22457 RSC .
  43. M. A. Said, A. Albohy, M. A. Abdelrahman and H. S. Ibrahim, Importance of glutamine 189 flexibility in SARS-CoV-2 main protease: Lesson learned from in silico virtual screening of ChEMBL database and molecular dynamics, Eur. J. Pharm. Sci., 2021, 160, 105744 CrossRef CAS .
  44. I. F. Nassar, A. A. H. Abdel-Rahman, E. M. Elnem, M. A. Said and M. G. Abouelenein, Facile Synthesis, In silico Studies, and Biological Assessment of Novel Pyrazolo [3,4-b]pyridine Congeners, Egypt. J. Chem., 2024, 67, 83–97 Search PubMed .
  45. M. G. Abouelenein, A. H. El-boghdady, H. M. Ali and M. A. Said, Molecular Investigations of Novel Pyrano [2,3-c] Pyrazole Congeners as Potential HCoV-229E Inhibitors: synthesis, Molecular Modeling, 3D QSAR, and ADMET Screening, Polycyclic Aromat. Compd., 2025, 45, 251–268 CrossRef .
  46. A. H. Khalil, E. A. Aidy, M. A. Said, R. Kebeish and A. H. Al-Badwy, Biochemical and molecular docking-based assessment of Spirulina platensis's bioactive constituents for their potential application as natural anticancer drug, Algal Res., 2024, 82, 103624 CrossRef .
  47. M. A. Mwaheb, N. M. Reda, M. S. El-Wetidy, A. H. Sheded, F. Al-Otibi, G. A. Al-Hamoud and E. A. Aidy, Versatile properties of Opuntia ficus-indica (L.) Mill. flowers: In vitro exploration of antioxidant, antimicrobial, and anticancer activities, network pharmacology analysis, and In-silico molecular docking simulation, PLoS One, 2024, 19, e0313064 CrossRef CAS PubMed .
  48. T. Lu and F. Chen, Multiwfn: A multifunctional wavefunction analyzer, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS .
  49. A. Aksimentiev, A. Arkhipov, R. Birnbaum, et al., Vmd Tutorial, Published online, 2017, pp. 1–77, https://www.ks.uiuc.edu/Training/Tutorials/.
  50. C. D. Laccase, CotA of Bacillus subtilis Is, 2001.
  51. H. Jacobsen, Localized-orbital locator (LOL) profiles of chemical bonding, J. Comput. Chem., 2008, 86, 695–702 CAS .

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5nj01851j

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2025
Click here to see how this site uses Cookies. View our privacy policy here.