Abstract
The present study focuses on the synthesis and characterization of novel bioactive complexes of Cr(III), Co(II), and Mn(II) derived from N′-((3-hydroxynaphthalen-2-yl)methylene)picolinohydrazide (H2L), a Schiff-base ligand featuring multifunctional donor sites. The primary objective was to investigate the coordination behavior, stability, and biological efficacy of these metal chelates. The complexes were synthesized via direct metal–ligand reactions and characterized using elemental analysis, FT-IR spectroscopy, UV–Vis, PXRD, 1H NMR spectroscopy, MS, magnetic susceptibility measurements, and thermogravimetric analysis. Spectroscopic evidence supported tetrahedral geometries for the Co(II) and Mn(II) complexes, while the Cr(III) complex exhibited an octahedral arrangement. Thermogravimetric and kinetic studies yielded positive activation free energy (ΔG*) values, indicating nonspontaneous decomposition pathways and high thermal stability. Quantum Theory of Atoms in Molecules (QTAIM) and reduced density gradient (RDG) analyses were performed to elucidate the nature of ligand–receptor interactions. Biological assessments revealed promising results, as DNA degradation assays demonstrated notable nuclease-like activity, particularly for the Co(II) complex. Antibacterial potency was evaluated via minimum inhibitory concentration (MIC), where the Mn(II) complex exhibited the strongest activity (0.313 mg/mL), followed by Co(II) and Cr(III) (0.625 mg/mL each), and the free ligand (1.250 mg/mL). Cytotoxicity testing against HeLa, HCT-116, and MCF-7 cancer cell lines showed high anticancer efficacy for the [CoL]·2H2O complex with IC50 values of 7.76 ± 0.4, 10.23 ± 0.8, and 6.88 ± 0.4 μM, respectively. Molecular docking studies using an induced fit protocol highlighted the strong noncovalent interactions of the complexes with DNA targets relevant to each cell line.


1. Introduction
Exploring coordination chemistry is an important goal in the search for innovative materials with distinct properties for a wide range of applications. Hydrazones are a distinct type of the Schiff-base ligand defined by the presence of the reactive (HN–NCH) functional group within their molecular formulas. Hydrazones are a specialized subclass of the Schiff base family, with a diverse range of biological and physiological activities. These actions include antitumor, antitubercular, antifungal, antioxidant, antibacterial, anti-inflammatory, and antimalarial effects. − Picolinate hydrazones are characterized by the existence of additional donor atoms, including the oxygen atom of the (CO) group and nitrogen atoms of both the (CN)Py and (CN)az groups. Previous studies have demonstrated that picolinic acid, a metabolite of L-tryptophan, exhibits significant biological potential, functioning as a neuroprotective, immunomodulatory, antimicrobial, and antiproliferative agent within the body. ,− Chromium picolinate has been identified as a dietary supplement that facilitates fat reduction and muscle enhancement in humans while also demonstrating efficacy in lowering blood glucose levels in diabetic patients. − Our research introduces a structurally distinct picolinate-based hydrazone ligand (H2L) featuring pyridine-N-oxide functionality. This design not only enriches the coordination chemistry landscape with multiple donor sites (CO, CNpy, and CNaz) but also enhances the ligand’s versatility in forming stable tridentate, dibasic complexes with transition metals. Compared to earlier reports by Deepa et al., Toyama et al., and Parvarinezhad et al., which involved simpler ligand systems, our synthesized ligand facilitates more robust coordination environments, as confirmed by extensive spectral, thermal, and magnetic analyses (IR, UV–visible, XRD, 1H NMR, MS, and TGA).
Furthermore, the prepared compounds transcend the isolated biological evaluations seen in previous works ,, by demonstrating a comprehensive profile of bioactivities, including anticancer, antibacterial, and antioxidant effects, DNA degradation, and BSA binding properties for the Co(II), Cr(III), and Mn(II) chelates. Molecular docking analysis provides additional insights into protein–ligand interactions, reinforcing the biomedical relevance of our complexes. Taken together, this work presents a distinctive contribution through the integration of an innovative ligand architecture, diverse coordination behavior, and broad-spectrum biological efficacy, clearly distinguishing it from prior studies and advancing the field of bioinorganic chemistry.
2. Experimental Section
2.1. Instruments
All of the chemical compounds involved in this investigation were of analytical grade and utilized directly without undergoing any further purification. The chemicals, reagents, and instruments used in the characterization are described in the Supporting Information. The DFT calculations of the ligand were performed using Gaussian 09 software with B3LYP/6-311G (d) basis set, while the isolated metal chelates were achieved with B3LYP/6-311G (d) for the C, H, N, and O atoms, and LANL2DZ basis set for the metal atom.
2.2. Preparation of the H2L Ligand and Its Complexes
The preparation of the H2L ligand involved the condensation of 3-hydroxy-2-naphthaldehyde and picolinohydrazide in a 1:1 molar ratio (Scheme ) at 90 °C for 4 h. An orange precipitate of the H2L ligand was obtained. The obtained precipitate was separated through filtration, subsequently recrystallized from ethyl alcohol, and dried under vacuum over anhydrous calcium chloride (CaCl2). The metal chelates were synthesized through refluxing a 1:1 molar ratio of the ligand and the corresponding metal salts, including CoCl2, CrCl3·6H2O, and MnCl2·4H2O, at 150 °C for 6 h (Scheme ). The solid precipitate was separated by filtration, washed, and subsequently dried over anhydrous CaCl2.
1. Preparation of H2L and Its Complexes.

2.3. Biological Applications
2.3.1. Antioxidant Activity Using the DPPH Assay
The antiradical activity of the examined compounds was evaluated using the DPPH• colorimetric assay with ascorbic acid as the standard. Serial dilutions were prepared by mixing equal volumes of the sample solution and methanol. A 0.135 mM solution of DPPH• was introduced to each dilution, followed by incubation of the resulting mixtures in a dark environment at 37 °C for a duration of 30 min. − Absorbance was measured at 517 nm. The residual DPPH• percentage was determined by using the stratified application of the following equation:
| 1 |
The % DPPH• remaining was plotted against the sample concentration (mg/mL) using an exponential curve to determine the IC50 value, representing the antioxidant quantity required to decrease the primary DPPH• concentration by 50%. Lower IC50 values correspond to higher antioxidant capacity. The reported total antioxidant activity values represent the mean of three independent experimental replicates.
2.3.2. Antibacterial Activities
The antibacterial potency of the investigated compounds was assessed utilizing the well diffusion assay technique outlined in previously published literature. , The antibacterial potency of agents against various bacterial strains was estimated by an agar well diffusion assay technique. The agar plate surface is prepared by evenly spreading a specific volume of the bacterial inoculum. A 923 mm diameter hole is then aseptically created using a sterile cork borer or tip. A 100 μL sample at the desired concentration is introduced into this well. The agar plates are then incubated under optimal conditions tailored to the specific test microorganism. During incubation, the antibacterial agent diffuses through the agar medium, effectively inhibiting the growth of the tested microbial strains.
2.3.3. Determination of the Minimum Inhibitory Concentration (MIC)
Serial dilutions of the sample at concentrations ranging from 5 to 0.078 mg/mL were used to establish the minimum concentration of the sample that would still exhibit antibacterial potency against Escherichia coli. The control consisted solely of inoculated broth, which was incubated for 1 day at 37 °C. The MIC endpoint is defined as the lowest sample concentration at which no visible microbial growth is observed in the tubes. The visual turbidity of the tubes was observed both prior to and following incubation to validate the MIC value, and the optical density (O.D.) was recorded at 600 nm to confirm the findings.
2.3.4. DNA Degradation Activity
The DNA degradation property of the examined compounds was estimated using Escherichia coli (E. coli) DNA as a substrate. Electrophoretic analysis was performed in a 1% agarose gel, prestained with ethidium bromide, under the conditions of 80 V for 2 h in Tris-(hydroxymethyl)methane buffer. Visualization and documentation of the gel were conducted using a UV transilluminator for precise observation of degradation patterns. ,
2.3.5. Polyacrylamide Gel Electrophoresis (SDS-PAGE)
Electrophoresis was performed using 10 mg of bovine serum albumin (BSA) treated with H2L ligands or metal complexes at a concentration of 5 mM. The prepared mixtures were incubated at 37 °C for 2 h, after which a six times sample buffer was added to the incubated mixtures. The stacking gel was made using a 4.0% (w/v) polyacrylamide solution in Tris buffer (pH 6.8), whereas the separating gel was made with a 15% (w/v) polyacrylamide solution in Tris buffer (pH 8.8) and 0.1% (w/v) SDS. Protein sample migration was conducted at 45 V for 15 min following the loading of samples in the stacking gel. Electrophoretic separation was then induced at 90 V in the running gel for 3 h. The protein bands were visualized by staining the SDS-PAGE gel with bromophenol blue, followed by destaining with a solution composed of methanol, glacial acetic acid, and distilled water.
2.3.6. Antitumor Activities
The antitumor potency of the tested samples was meticulously investigated by an MTT assay. The evaluation targeted specific cancer cell lines, including HeLa (epithelioid carcinoma of the cervix), MCF-7 (mammary gland breast adenocarcinoma), and HCT-116 (human colorectal adenocarcinoma). The inhibitory properties of the tested samples on cell proliferation were assessed by the MTT assay. Cells were cultured in RPMI-1640 medium with supplements and placed at a density of 1.0 × 104 cells/well. After treatment with varying compound concentrations, MTT solution was added, and formazan crystals were dissolved in DMSO solution. The absorbance was measured at 570 nm, and the cell viability (%) was calculated. ,
2.4. Molecular Docking with Receptors 5IAE, 2W3L, and 3W2S
Molecular docking simulations, a key component of computational drug design, offer mechanistic insights into drug–target interactions. By providing detailed information on binding affinity and orientation, these simulations support the rational development of new pharmaceutical entities.
2.4.1. Ligand and Protein Preparation
For molecular docking studies, the synthesized compounds underwent optimization using Ligprep (Schrödinger suite) with default settings. The three-dimensional structures of HeLa (PDB ID: 5IAE) https://www.rcsb.org/structure/5IAE, HCT-116 (PDB ID: 1YWN) https://www.rcsb.org/structure/1YWN, and MCF-7 (PDB ID: 3W2S) https://www.rcsb.org/structure/3W2S were retrieved from the protein database. − Protein structure preparation was performed using the Protein Preparation Wizard (Maestro v. 14.1, Schrödinger suite), encompassing error correction, side-chain reconstruction, and structural repair. The protonation states were determined using PROPKA at pH 7.0, followed by energy minimization with the OPLS4 force field until the heavy atom RMSD converged to 0.30 Å. The bond orders were determined, and hydrogen atoms were incorporated into the structure at a controlled pH of 7.0 ± 2.0. Final restrained minimization, employing the OPLS 2005 force field, was performed with heavy atom RMSD convergence set at 0.3 Å.
2.4.2. Induced Fit Docking
Molecular docking simulations were conducted utilizing the Schrödinger induced fit docking (IFD) protocol. A 10 Å grid box, centered on the native ligand coordinates of 5IAE, 1YWN, and 3W2S, was generated. Ligand flexibility was addressed with Glide (v. 10.4) using a 2.5 kcal/mol ring conformation sampling window, while receptor flexibility was accounted for using Prime. The IFD protocol employed modified van der Waals and Coulombic parameters and temporary side-chain removal. The docking of the minimized compounds was performed with standard sampling, a potential scaling factor of 0.5, and a maximum of five retained poses. IFD scores, reflecting protein–ligand interaction and total system energy, were used for pose ranking.
2.4.3. Docking Validation
The reliability of the docking procedure was assessed through redocking experiments. The cocrystallized ligands were removed and redocked into their respective binding sites using the Schrödinger suite software, without altering any function parameters. This validation step was performed to ensure the accurate reproduction of ligand binding within the active site and to minimize deviations from the experimentally determined cocrystallized structures.
3. Results and Discussion
The metal complexes (10–4 M in DMF) exhibited molar conductance values in the range of 6.32–9.44 Ω–1 cm2 mol–1, indicative of their nonelectrolytic nature. Table S1 provides an overview of the various physical properties and analytical data of the isolated compounds. Elemental analysis data, combined with spectral studies, were utilized to infer the structural formulas and propose the geometries of the complexes. Moreover, all complexes are stable in air and insoluble in most organic solvents, except DMSO and DMF. Additionally, the M–Cl bond in our system functions as a terminal coordinating halide, contributing to the stabilization of the metal center through σ-donation.
3.1. Infrared Spectra
The IR spectra of H2L and the synthesized complexes were assigned from 4000 to 400 cm–1 (Table S2) and are represented graphically in Figure S1. The spectrum of H2L displays two strong vibrational peaks at about 1662 and 1528 cm–1, which are attributed to the ν(CO) and ν(CN)py groups, respectively. The band at 1619 cm–1 is attributed to the azomethine group. The broad bands at 3172 and 3464 cm–1 are assigned to the ν(NH) and ν(OH) groups, respectively. On the other hand, the strong peak at 1033 cm–1 is assigned to the ν(N–N) vibration. The infrared spectra of the studied complexes were discussed by comparing their spectra with those of the free ligands. Consequently, the IR spectrum of the [CoL]·2H2O complex shows that the H2L ligand acts as a binegative tetradentate Schiff base (ONNO) by deprotonating both OH groups of the enolic and aromatic rings and (CN)az, (CN)Py groups. This bonding mode is obtained through the following:
-
(i)
The absence of vibration bands of the (CO) and (NH) groups, alongside the emergence of new bands associated with ν(CN) and ν(C–O) vibrations, indicates the enolization of the carbonyl group.
-
(ii)
The absence of OH vibrations serves as evidence of deprotonation.
-
(iii)
The observed shift of the ν(CN)az and ν(N–N) vibrations to a new position signifies their sharing in the coordination.
-
(iv)
The movement of the ν(CN)Py vibration suggests its involvement in complex formation.
Although in the [Cr(L)(H2O)2Cl]·H2O complex, the H2L ligand acts as a bibasic tridentate Schiff base (ONO) via the nitrogen of the azomethine group and deprotonated of both enolic and phenolic OH groups, this coordination mode is proposed by the following:
-
(i)
The disappearance of the ν(NH) and ν(CO) bands, accompanied by the emergence of a new band attributed to ν(–CN), confirms the enolization form.
-
(ii)
The absence of an (OH) group signal suggests that deprotonation occurred.
-
(iii)
The observed shift of the ν(CN)az and ν(N–N) vibrations to a new position signify their participation in the coordination.
-
(iv)
There were no changes in the positions of the ν(CN)py and δ(CN)py bands, indicating that this functional group was not involved in the chelation process.
In the [Mn(HL)] complex, the H2L ligand acts as a mononegative tetradentate Schiff base (ONNO) via the (CO) and (CN)az groups, and a deprotonated OH group of the aromatic rings. This bonding mode is explained as follows:
-
(i)
The presence of the stretching vibrations of amine and carbonyl groups validates the keto structure, while the observed shift of the ν(CO) vibrational band to another position provides evidence of its involvement in the chelation process.
-
(ii)
The disappearance of the OH vibrational band provides compelling evidence for the occurrence of deprotonation.
-
(iii)
The observed shift of the ν(CN)az and ν(N–N) vibrations to a new position signifies their participation in the coordination.
-
(iv)
The movement of the position of the ν(CN)Py group reveals that this group is employed in complexation.
3.2. Nucleic Magnetic Resonance (NMR) Analysis
The NMR spectrum of the synthesized H2L was obtained in DMSO-d 6 solvent. The proton NMR spectrum of the H2L ligand (Figure S2) exhibited a singlet signal at 12.214 ppm, which was attributed to the phenolic proton. This upper chemical shift value may be attributed to the influence of hydrogen bonding. The singlet peak at 9.83 ppm is assigned to the amino proton. These signals disappeared upon deuteration (Figure S3). The characteristic signal appears at 8.77 ppm, which is related to the azomethine proton. The multiple peaks observed within the chemical shift range of 7.24–8.22 ppm correspond to the aromatic protons. The H2L ligand was analyzed by recording its 13C NMR spectrum, as presented in Figure S4. Multiple bands were seen in the downfield region, within the range of 119.42–138.61 ppm, relating to the aromatic nature of the carbons present in the ligand structure. The triplet signals at 149.03, 149.12, and 149.42 ppm are attributed to the carbon of the (CN)py, (CN)az, and (OC–NH) groups, respectively. The signals at 160.52 and 158.83 ppm are assigned to the carbon of the (CO) and C–OH groups, respectively.
3.3. Mass Spectra
The EI-MS spectra were analyzed to validate the proposed molecular structures of the H2L ligand and its complexes. The mass spectrum of H2L (Figure S5) revealed a molecular ion peak [M]+ at m/z = 291.31, which corresponds to its molecular weight. The proposed fragmentation pathway of H2L is outlined and visually presented in Scheme . The EI-MS of the the [CoL]·2H2O complex (Figure S6) offered strong evidence supporting its suggested molecular formula and a molecular weight of 384.26. The molecular ion peak at m/z = 384.30 (69.70%) corresponds to (C17H15CoN3O4). The EI-MS spectrum of this complex shows distinct fragments that produce peaks of varying intensities at different m/z values, such as at 128.17 (100%) (C10H8), 221.07 (17.40%) (C7H5CoN3O2), 142.04 (7.31%) (C10H6O2), 205.98 (18.45%) (C7H5CoN3O), 120.03 (1027%) (C6H4N2O), and 227.99 (13.39%) (C11H7CoNO). The mass spectrum of the [Cr(L)(H2O)2Cl]·H2O complex (Figure S7) confirmed the proposed molecular formula. The molecular ion peaks at m/z = 430.79 (52.66%) correspond to (C17H17ClCrN3O5). Finally, EI-MS of the [Mn(HL)] complex (Figure S8) confirmed its molecular formula with a molecular ion peak at m/z = 345.54 (65.21) corresponding to (C17H12MnN3O2). Other fragments appear at different m/z values, such as at 57.01 (100%) (CHN2O), 263.37 (26.11%) (C14H10MnNO), 113.95 (27.32%) (C9H6), and 306.08 (31.57%) (C14H9MnN3O2).
2. Mass Fragmentation Path of the H2L Ligand.
3.4. Electronic Spectrum and Magnetic Susceptibility Studies
The electronic spectra of the H2L ligand and its metal complexes were recorded in DMF solvent at room temperature with a concentration of 10–4 M. The magnetic moments and spectral absorption bands are shown graphically in Figure S9 and are reported in Table S3. The electronic spectrum of H2L exhibits three absorption bands: the first one, located at 270 nm, is attributed to the π→π* transition of the aromatic rings. The two bands at 329 and 361 nm are attributed to the n→π* transition of carbonyl and (CN) of both the pyridyl and azomethine groups. The other bands at higher wavelengths within the range of 384–430 nm are due to charge transfer. These bands show a shift toward higher wavenumbers in the metal chelates due to chelation. , In the cobalt complex, the bands at 275 and 327 nm are assigned to the π→π* and n→π* transitions, respectively. The two bands at 358 and 382 nm are due to charge transfer, while the absorption band at 450 nm is attributable to the 4A2→4T1 (P) transition. These results indicate the existence of a tetrahedral geometry for the [CoL]·2H2O complex, proven by its magnetic moment of 4.36 (B.M.). This value lies within the expected range for the tetrahedral arrangement of divalent cobalt complexes, typically between 4.2 and 4.8 μB. The UV-visible spectrum of the [Cr(L)(H2O)2Cl]·H2O complex exhibits absorption bands in the wavelength range of 275–497 nm. The two bands at 275 and 336 nm correspond to the ligand field transitions. The bands at 464 and 498 nm are assigned to the 4A2g(F) →4T1g(F) and 4A2g(F) →4T2g(F) transitions, respectively. , As expected, the υ3 band at about 300 nm, which corresponds to the 4A2g→ 4T1g(P) transition, is obscured by the charge transfer bands. , Furthermore, the μeff value of 3.56 μB supports the octahedral geometry proposed for this complex. In the [Mn(HL)] complex, the bands at 362 and 382 nm correspond to the LMCT transition. The absorption at 465 nm is related to 6A1→4T2, corresponding to the tetrahedral Mn(II) environment. , The measured magnetic moment μeff was calculated to be 5.58 μB, corroborating the characterization of a high-spin tetrahedral Mn(II) complex.
3.5. Powder XRD Studies
Due to the inability to separate single crystals of the synthesized compounds for X-ray crystallographic analysis, powder X-ray diffraction data were obtained for structural description. X-ray diffraction analysis was conducted to provide comprehensive insights into the crystalline nature of H2L and its chelates. The diffractograms of the studied compounds (Figure ) display several sharp peaks, indicating their polycrystalline nature. The computer software BIOVIA Material Studio, the reflex powder indexing module, and the TREOR9 indexing algorithm were employed to calculate the theoretical values of 2θo, crystal system, space group, and lattice parameters. Table S4 presents the crystallite-particulate parameters, including the 2θo values, interplanar spacing (d-values, Å), relative intensity, dislocation density (δ), and crystal strain (ε) of the investigated complex, all calculated using standard equations. − The diffraction data was employed to determine the average crystallite size (D) using the Scherrer equation
| 2 |
where λ= 1.5406 Å, θ represents the Bragg diffraction angle, and β denotes the full width at half-maximum (fwhm) of the peak. , Based on the highest intensity peaks, the grain sizes of the studied compounds were found to be 57.65, 18.42, 81.75, and 81.21 nm for the H2L ligand, [CoL]·2H2O, [Cr(L)(H2O)2Cl]·H2O, and [Mn(HL)] complexes, respectively. According to the δ values, the polycrystalline phase of the ligand and its complexes increased in the following order: [Cr(L)(H2O)2Cl]·H2O (1.79 × 10–3) > [Mn(HL)] (2.05 × 10–3) > H2L(2.84 × 10–3) > [CoL]·2H2O (7.63 × 10–3 nm–2).
1.

Experimental and simulated diffractograms of the H2L ligand (a), [CoL]·2H2O (b), [Cr(L)(H2O)2Cl]·H2O (c), and [Mn(HL)] (d) complexes.
3.6. Thermal Studies
3.6.1. Thermogravimetric Analysis (TG)
A detailed thermal investigation of the Cr(III), Co(II), and Mn(II) chelates was performed using thermogravimetric analysis (TGA) in the temperature range of 25–800 °C (Table S5). The goal of this analysis was to establish the type of water molecules found inside these complexes, namely, whether they exist in crystalline form or are coordinated to the metal centers. The structural characteristics and thermal stability of the complexes are clarified in this work. The thermogram of the [CoL]·2H2O complex displayed five decomposition steps (Figure ). The first stage of fragmentation appeared in the temperature range of 28 to 137 °C, which corresponded to the loss of two hydrated water molecules (weight loss: found 8.93%; calcd 9.36%). The second thermal step, occurring in the temperature range of 218–326 °C, is accompanied by a measured weight loss of 11.25% after the elimination of CO + 0.5N2 molecules from the complex structure. The third step within the temperature range of 327–450 °C corresponds to the loss of the C6H4CNOH fragment (weight loss: found 31.40%; calcd 31.02%). The fourth stage of the thermal curve occurred in the temperature range of 451–481 °C, with a weight loss percentage of 20.98%, corresponding to the removal of the C5H5N composite. The final residue upon decomposition, attributed to cobalt, , was detected at temperatures greater than 670 °C, and the residual weight was 15.75% (calcd: 15.34%). The thermodecomposition behavior of the [Cr(L)(H2O)2Cl]·H2O complex is marked by four steps of decomposition, as shown in Figure S10. The initial stage of decomposition, observed in the range of 24–100 °C, is due to the loss of a hydrated water molecule, and a mass loss value of 4.56% (calcd: 4.18%) was recorded. The second thermal degradation, occurring within the 126–360 °C temperature range, is associated with the evolution of two coordinated water and HCl molecules from the complex, and consequently an observed mass loss of 16.16% (calcd: 16.82%). The third thermolysis step in the temperature range of 361–515 °C is attributed to the elimination of the C6H4CNOH fragment with the release of N2 gas, and the weight loss observed is 34.40% (calcd: 34.14%). The final degradation process, in the temperature range of 516–733 °C, is attributed to the loss of the C5H5 moiety, with a mass loss observed of 14.88%. The remaining part at temperatures greater than 734 °C accounted for 30.00% of the original mass and is due to the presence of residual organic carbon species and chromium monoxide (CrO). The thermogravimetric trace shows no noticeable mass loss below 200 °C, which indicates that there are no coordinated or lattice water molecules in the structure of the Mn(II) complex (Figure S11). The first step of the thermal decomposition occurs within the temperature range of 204–333 °C and involves the release of N2 and H2 gases, with a weight loss of 17.03%. The second thermal decomposition stage, occurring between 334 and 413 °C, corresponds to the elimination of the C5H4N fragment from the complex structure, accompanied by an observed mass loss of 21.74%. The final thermodecomposition stage, between 414 and 664 °C, corresponds to the loss of the C6H4 moiety and the concomitant release of C2H2 gas from the complex structure with a mass loss of 30.03%. At temperatures exceeding 665 °C, the remaining residue consists predominantly of manganese oxide (MnO) and residual organic carbon, accounting for 31.02% of the initial sample mass.
2.
Thermogram of the [CoL]·2H2O complex.
3.6.2. Thermodynamics and Kinetics Studies
Based on the thermogram, the kinetic and thermodynamic parameters of the complexes were determined by using the Coats–Redfern and Horowitz–Metzger equations. The data supporting these calculations are documented in Tables and S6 and are visually represented in Figures S12–S17. The Eyring equations were mathematically applied to investigate and analyze the kinetic and thermodynamic parameters of the complexes. The negative values of ΔS* observed for the investigated complexes suggest that they are more ordered than the original ligand, indicating a slow thermal decomposition process. Conversely, the positive values of activation entropy (ΔS*) may imply that the disorder of the disintegrated fragments increases notably quicker than that of the undecomposed species. , The endothermic nature of all fragmentation processes is evidenced by the positive values of activation enthalpy. The complexes under study display thermal stability, and the positive signs of ΔG* demonstrate that all decomposition steps are nonspontaneous. ,
1. Kinetic Parameters Calculated by the Coats–Redfern Method for the Co(II) and Cr(III) Complexes.
| complex | peak | mid temp (K) | E a (kJ/mol) | A (S–1) | ΔH* (kJ/mol) | ΔS* (kJ/mol·K) | ΔG* (kJ/mol) |
|---|---|---|---|---|---|---|---|
| [CoL]·2H2O | 1st | 317 | 113.23 | 6.63 × 1016 | 110.60 | 0.0766 | 86.31 |
| 2nd | 509 | 416.68 | 4.78 × 1040 | 412.44 | 0.5294 | 142.97 | |
| 3rd | 647 | 1057.1 | 7.44 × 1083 | 1052.1 | 1.354 | 175.94 | |
| 4th | 738 | 1843.1 | 9.62 × 10128 | 1836.9 | 2.216 | 201.0 | |
| 5th | 899 | 784.68 | 4.45 × 1043 | 777.21 | 0.5815 | 254.42 | |
| [Cr(L)(H2O)2Cl]·H2O | 1st | 311 | 101.41 | 1.29 × 1015 | 98.83 | 0.0440 | 85.13 |
| 2nd | 521 | 371.75 | 3.47 × 1034 | 367.41 | 0.4117 | 152.92 | |
| 3rd | 704 | 231.19 | 7.46 × 1014 | 225.34 | 0.0326 | 202.33 | |
| 4th | 817 | 320.17 | 3.14 × 1018 | 313.38 | 0.1008 | 231.00 | |
| [Mn(HL)] | 1st | 572 | 210.07 | 1.08 × 1017 | 205.32 | 0.0757 | 162.00 |
| 2nd | 637 | 329.21 | 7.16 × 1024 | 323.92 | 0.2246 | 180.84 | |
| 3rd | 748 | 270.64 | 4.25 × 1016 | 264.42 | 0.0658 | 215.21 |
3.7. DFT Studies
3.7.1. Geometry Optimization
Figure S18 displays the optimized structures of the H2L ligand and [CoL]·2H2O, [Cr(L)(H2O)2Cl]·H2O, and [Mn(HL)] complexes. Tables S7 and S14 comprehensively detail the bond lengths and angles of the synthesized compounds. The bond angles and lengths vary depending on the metal center and the coordinating environment. The bond angles in the [CoL]·2H2O and [Mn(HL)] complexes show tetrahedral structures with sp3 hybridization, similar to those of tetrahedral Co(II) and Mn(II) complexes. The bond angles in the [Cr(L)(H2O)2Cl]·H2O chelate are typical of the octahedral environment observed in Cr(III) chelates with d2sp3 hybridization. Consistency in bond length elongation for active groups such as CO, CN, and OH was observed upon coordination with metal ions. When the metal–nitrogen (M–N) bond forms, the carbon–nitrogen (C–N) bond weakens due to coordination with the imine group’s nitrogen atom.When the metal–nitrogen (M-N) bond is formed, the carbon–nitrogen (C–N) bond becomes weaker because of coordination with the imine group’s nitrogen atom (CN). This weakening is caused by the nitrogen lone pair’s participation in the formation of the coordination bond with the metal ion, which previously contributed to the double-bond nature of the CN bond. As a result, the electron density of the CN bond is lowered, resulting in decreased strength. The formation of the (M–O) bond induces weakening, thereby facilitating chelation of the carbonyl group (CO).
3.7.2. Frontier Molecular Analysis
The HOMO and LUMO energies define a molecule’s capacity to donate or accept electrons. The chemical reactivity is determined by the energy gap between these two orbital levels. The small energy band gap suggests that the compound demonstrates strong chemical reactivity, considerable biological potency, and strong polarizability. The obtained HOMO and LUMO of the studied compounds are shown in Figures and S19. The computed energies of the HOMO and LUMO are used to calculate various chemical parameters, including electronegativity (χ), chemical potential (μ), hardness (η), softness (σ), and electrophilicity index (ω) (Table S15). A higher ΔE value suggests hardness, whereas a lower ΔE value indicates softness. A soft molecule exhibits greater reactivity compared to a hard molecule due to its ability to donate electrons readily to electron acceptors. The ionization potential is relative to the energy level of the HOMO, whereas the electron affinity is relative to the energy level of the LUMO, both of which are positive. An increase in the ionization potential and electron affinity indicates that the element has greater electronegativity. The metal complexes display smaller band gap energies compared to those of the H2L ligand. This demonstrates their enhanced electron transfer capacity, decreased kinetic stability, increased chemical reactivity, and significant biological activity. Because of their improved electron transport capabilities, these complexes are expected to interact more effectively with macromolecular structures, allowing for easier binding to proteins or DNA. All investigated compounds demonstrate a negative chemical potential (μ), revealing that the inclusion process occurs spontaneously and that the compounds are thermodynamically stable. The chemical potential quantitatively represents the inclination of electrons to diverge from an equilibrium configuration, which decreased as follows: [Mn(HL)] complex (−3.54) > H2L ligand (−4.05) > [CoL]·2H2O complex (−4.17) > [Cr(L)(H2O)2Cl]·H2O (−4.41) eV.
3.
Frontier molecular orbitals (HOMO and LUMO) of the H2L ligand.
3.7.3. Molecular Electrostatic Potential (MEP)
MEP diagrams are a valuable index for visualizing the distribution of electric charges on a molecule’s surface. This understanding helps us understand the molecule’s chemical and physical features. The establishment of many intermolecular interactions, including hydrogen bonding, drug–receptor binding, and enzyme–substrate interactions, is mostly dictated by the electrostatic potential distribution of the molecule. MEP is a well-known method for providing a comprehensive understanding of the reactive sites of compounds, especially identifying areas that react to electrophilic and nucleophilic interactions. The MEP mapping of the H2L Schiff base and its chelates may be classified into three distinct regions based on their color representation, as seen in Figure . The red regions correspond to areas of high electronic density, which aid electrophilic attacks. In contrast, the green regions represent zones of neutral electrostatic potential, while the blue region signifies regions of low electronic density, indicating nucleophilic attack behavior. The regions of negative electrostatic potential are concentrated around electronegative atoms with red color, including the oxygen atoms in carbonyl and hydroxyl groups, nitrogen of the pyridine ring, and nitrogen in the imine group. The hydrogen atom bonded to nitrogen in the (NH) group is positioned within the blue region, which displays a distinctly positive electrostatic potential, rendering it highly vulnerable to nucleophilic attack.
4.

ESP maps of the studied compounds.
3.8. Topology Analyses
3.8.1. Quantum Theory of Atoms in Molecules (QTAIM)
QTAIM analysis was conducted to characterize ligand–receptor interactions, focusing on bond critical points (BCPs). Topological parameters (Table ) and electron density maps (Figure ) were generated. Contour maps showed a decrease in electron density (ρ) from the nuclear centers, and relief maps identified BCPs as the electron density minimum along the bond paths. Negative values of the electronic energy density H(r) and Laplacian of electron density ∇2ρr confirmed the covalent bond characteristics, while positive values indicated noncovalent interactions. The covalent bond character was supported by the calculated negative H(r) and ∇2ρr values (Table ). The G(r)/V(r) ratio, used to differentiate the interaction types, revealed covalent interactions G(r)/V(r) < 0.5 for most bonds, except for BCP 6 (N3–H27) in ligand H2L, which exhibited a noncovalent (van der Waals) interaction G(r)/V(r) > 1.
2. Calculated Topological Parameters (in a.u) of Ligand H2L.
| BCP no. | atoms | ρr | ∇2ρr | V r | G r | H r | G r/V r |
|---|---|---|---|---|---|---|---|
| 1 | C2–N3 | 0.339 | –0.933 | –0.747 | 0.257 | –0.490 | 0.344 |
| 2 | C1–H23 | 0.282 | –0.975 | –0.319 | 0.037 | –0.281 | 0.118 |
| 3 | C1–C2 | 0.311 | –0.876 | –0.419 | 0.100 | –0.319 | 0.239 |
| 4 | N8–H27 | 0.333 | –1.653 | –0.513 | 0.050 | –0.463 | 0.097 |
| 5 | N3–C4 | 0.338 | –0.964 | –0.719 | 0.239 | –0.480 | 0.332 |
| 6 | N3–H27 | 0.020 | 0.082 | –0.014 | 0.017 | 0.003 | 1.228 |
| 7 | C7–N8 | 0.313 | –0.882 | –0.630 | 0.205 | –0.425 | 0.325 |
| 8 | C1–C6 | 0.310 | –0.874 | –0.419 | 0.100 | –0.318 | 0.239 |
| 9 | N8–N9 | 0.359 | –0.691 | –0.520 | 0.173 | –0.346 | 0.334 |
| 10 | C4–C7 | 0.258 | –0.633 | –0.272 | 0.057 | –0.215 | 0.209 |
| 11 | C5–C6 | 0.311 | –0.879 | –0.422 | 0.101 | –0.321 | 0.240 |
| 12 | C4–C5 | 0.311 | –0.874 | –0.419 | 0.100 | –0.319 | 0.240 |
| 13 | C6–H26 | 0.283 | –0.984 | –0.318 | 0.036 | –0.282 | 0.113 |
| 14 | N9–C21 | 0.370 | –0.692 | –0.991 | 0.409 | –0.582 | 0.413 |
| 15 | C7–O10 | 0.410 | –0.253 | –1.318 | 0.627 | –0.690 | 0.476 |
| 16 | C5–H25 | 0.285 | –0.999 | –0.319 | 0.035 | –0.284 | 0.108 |
| 17 | C11–C21 | 0.273 | –0.699 | –0.313 | 0.069 | –0.244 | 0.221 |
| 18 | C11–C12 | 0.294 | –0.796 | –0.369 | 0.085 | –0.284 | 0.231 |
| 19 | C12–C13 | 0.321 | –0.913 | –0.459 | 0.115 | –0.344 | 0.251 |
| 20 | C13–H28 | 0.276 | –0.928 | –0.316 | 0.042 | –0.274 | 0.132 |
| 21 | C13–C14 | 0.295 | –0.792 | –0.376 | 0.089 | –0.287 | 0.237 |
| 22 | C11–C16 | 0.314 | –0.880 | –0.433 | 0.107 | –0.327 | 0.246 |
| 23 | C14–C15 | 0.291 | –0.774 | –0.363 | 0.085 | –0.278 | 0.233 |
| 24 | C17–H30 | 0.280 | –0.957 | –0.318 | 0.039 | –0.279 | 0.124 |
| 25 | C17–C18 | 0.319 | –0.909 | –0.447 | 0.110 | –0.337 | 0.246 |
| 26 | C15–C16 | 0.300 | –0.820 | –0.387 | 0.091 | –0.296 | 0.235 |
| 27 | C14–C17 | 0.296 | –0.801 | –0.377 | 0.088 | –0.289 | 0.234 |
| 28 | C18–C19 | 0.297 | –0.807 | –0.380 | 0.089 | –0.291 | 0.235 |
| 29 | C18–H31 | 0.282 | –0.969 | –0.319 | 0.038 | –0.281 | 0.120 |
| 30 | C15–C20 | 0.295 | –0.797 | –0.373 | 0.087 | –0.286 | 0.233 |
| 31 | C19–C20 | 0.320 | –0.915 | –0.450 | 0.111 | –0.339 | 0.246 |
| 32 | C16–H29 | 0.284 | –0.988 | –0.318 | 0.036 | –0.283 | 0.112 |
| 33 | C19–H32 | 0.282 | –0.967 | –0.319 | 0.039 | –0.281 | 0.121 |
| 34 | C12–O22 | 0.279 | –0.358 | –0.695 | 0.303 | –0.392 | 0.436 |
| 35 | C2–H24 | 0.286 | –0.999 | –0.319 | 0.034 | –0.284 | 0.108 |
| 36 | C20–H33 | 0.281 | –0.967 | –0.318 | 0.038 | –0.280 | 0.120 |
| 37 | C21–H34 | 0.280 | –0.955 | –0.311 | 0.036 | –0.275 | 0.116 |
| 38 | O22–H35 | 0.366 | –2.535 | –0.776 | 0.071 | –0.705 | 0.092 |
5.
Relief and contour maps of ligand H2L.
3.8.2. Reduced Density Gradient
RDG analysis was employed to characterize the noncovalent interactions. RDG, a dimensionless quantity, serves as a metric for evaluating weak interactions in real space, according to electron density, and is defined by the equation
| 3 |
This method enables the discovery of areas with less electron density, indicating weak interactions, while high-density gradient values correspond to strong interactions. Scatter plots and color-filled contour maps of the RDG for the investigated compounds were generated using the Multiwfn program (Figure ). The second eigenvalue of the Hessian matrix (λ2) was employed as a criterion to differentiate between bonded interactions, characterized by λ2 < 0, and nonbonded interactions, marked by λ2 > 0. In this investigation, the ρ function associated with the λ2 sign exhibited a range spanning from −0.05 to 0.05 atomic units (a.u.). The RDG spectra were color-coded as red, green, and blue. The red peaks in the λ2 > 0 region indicate steric repulsion within the ring. The peaks observed in the λ2 ≈ 0 region represent London dispersion or dipole–dipole forces. The blue peaks in the ρ > 0 and λ2 < 0 regions signify electrostatic interactions, such as hydrogen and halogen bonds.
6.
NCI and RDG graphs of ligand H2L.
3.9. Biological Applications
3.9.1. Antioxidant Potency
The antiradical potency of the prepared compounds was assessed by measuring the IC50 values, with lower IC50 values indicating greater antioxidant activity. We evaluated the antiradical activity of ascorbic acid as a reference antioxidant, and the findings are shown in Figure . H2L exhibited moderate antioxidant properties with an IC50 of 0.156 ± 0.0099 mg/mL. The [Cr(L)(H2O)2Cl]·H2O, [Mn(HL)], and [CoL]·2H2O complexes are more active than the ligand alone, with IC50 values of 0.0258 ± 0.0058, 0.0341 ± 0.0082, and 0.0839 ± 0.0057 mg/mL, respectively. This suggests that the complex formation-induced structural changes redistribute the electrostatic charge, thereby increasing the activity. The interaction between the ligand and the positively charged metal induces a transfer of electron density away from the ligand, thereby enhancing its polarization. Overall, the findings suggest that complexation can greatly enhance the bioactivity of organic ligands, with the [Cr(L)(H2O)2Cl]·H2O complex identified as the most potent antioxidant among the compounds tested, suggesting their potential for further investigation and development as therapeutic agents targeting pathological conditions associated with oxidative stress. ,
7.

DPPH scavenging activities (IC50 values) of the studied compounds.
This enhancement can be attributed to metal–ligand complexation that leads to structural and electronic modifications that redistribute electron density, thereby augmenting the free radical scavenging ability of the ligand. Structural modifications of the ligand induced by coordination with 3d-transition metals result in distinct variations in the biological activity of the corresponding complexes. Existing literature corroborates that metal complexation significantly influences the microbiological and cytotoxic profiles of ligands. The reaction mechanism is likely associated with electron transfer and radical stabilization initiated via coordination of the ligand to the positively charged metal centers. Transition metal complexation increases the electron-withdrawing nature of the ligand, enhancing its polarization and enabling hydrogen atom transfer or single-electron transfer mechanisms. For example, the presence of solvent molecules in [Cr(L)(H2O)2Cl]·H2O and the tendency of Cr(III) to strongly accept an electron raise the redox potential of the complex, making it the most effective antioxidant compound among those investigated. The interaction with the positively charged metal results in an electron density shift away from the ligand, thereby polarizing it further. The findings concur with previous observations that biologically active metal complexes of ligands have greater antioxidant activity compared to free ligands due to synergistic metal–ligand interactions.
Overall, the findings suggest that complexation can greatly enhance the bioactivity of organic ligands, with the [Cr(L)(H2O)2Cl]·H2O complex identified as the most potent antioxidant among the compounds tested, suggesting their potential for further investigation and development as therapeutic agents targeting pathological conditions associated with oxidative stress. ,
3.9.2. Antibacterial Activity
The synthesized ligand (H2L), along with its [Cr(L)(H2O)2Cl]·H2O, [CoL]·2H2O, and [Mn(HL)] chelates, were evaluated for their in vitro antibacterial potency against various microbial strains comparable to Ciprofloxacin antibiotic as the reference documented in Table .
For the Escherichia coli strain, the findings demonstrate that the [Mn(HL)] chelate exhibits greater antibacterial potential than the ligand and other complexes. The antibacterial activity for this study decreased as Mn(II) > Cr(III) = Co(II)> H2L.
For the Salmonella typhimurium and Enterobacter cloacae strains, all the tested compounds did not exhibit antibacterial activities.
For the Klebsiella pneumonia strain, only the Mn(II) complex exhibited antibacterial activity.
For the Bacillus subtilis strain, Cr(III) exhibits higher activity than H2L, while the [Mn(HL)] and [CoL]·2H2O complexes are inactive against Bacillus subtilis.
For the Bacillus cereus strain, the order of activity was Cr(III) > Mn(II) > H2L, while the cobalt complex did not exhibit antibacterial activity.
For Staphylococcus aureus and Staphylococcus epidermidis strains, the antibacterial activities decreased in the order Cr(III) > Co(II) > Mn(II) > H2L.
3. Antibacterial Activities of the Studied Complexes .
| inhibition
zones in mm |
||||||
|---|---|---|---|---|---|---|
| microorganisms | H2L | Mn(II) | Cr(III) | Co(II) | control (DMSO) | antibiotic |
| Gram-negative Bacteria | ||||||
| Escherichia coli | 14 | 20 | 18 | 16 | 25 | |
| Salmonella typhimurium | 14 | |||||
| Klebsiella pneumonia | 14 | 20 | ||||
| Enterobacter cloacae | ||||||
| Gram-positive Bacteria | ||||||
| Bacillus subtilis | 14 | 19 | 25 | |||
| Bacillus cereus | 11 | 16 | 17 | 15 | ||
| Staphylococcus aureus | 12 | 15 | 17 | 16 | 23 | |
| Staphylococcus epidermidis | 12 | 16 | 17 | 16 | 23 | |
No inhibition zone means <9 mm.
The antibacterial application of metal complexes varies due to several factors, including the metal’s biological compatibility, its interactions with enzymatic systems, the generation of oxidative stress, the complex’s stability, and the inherent toxicity of the metal. The results demonstrated that most of the studied metal complexes exhibited greater antibacterial activity than H2L. This enhanced activity can be attributed to the improved lipophilicity of the complexes resulting from chelation.
3.9.3. Determination of Minimum Inhibitory Concentration (MIC)
All of the complexes showed strong antibacterial activity against the Escherichia coli strain. Given their strong antibacterial properties, these complexes were subjected to minimum inhibitory concentration (MIC) testing against Escherichia coli (Table ). After 24 h of incubation at 37 °C, turbidity was observed starting from a concentration of 0.625 mg/mL H2L, indicating an MIC of 1.250 mg/mL. In the Mn(II) complex, turbidity was observed starting from a concentration of 0.156 mg/mL, indicating an MIC of 0.313 mg/mL. In the Cr(III) complex, turbidity was observed starting from a concentration of 0.313 mg/mL, indicating an MIC of 0.625 mg/mL. In Co(II), turbidity was observed starting from a concentration of 0.313 mg/mL, indicating an MIC of 0.625 mg/mL. Based on the MIC values, the activity of the tested samples decreased in the following order: Mn(II) (0.313 mg/mL) > Co(II) (0.625 mg/mL) = Cr(III) (0.625 mg/mL) > L (1.250 mg/mL). Overall, the findings indicate that Mn(II) has strong efficacy against bacterial pathogens, indicating that it could be a viable candidate for antimicrobial drug development.
4. Measured Optical Densities at a Wavelength of 600 nm for E. coli for the Investigated Compounds.
| OD600
|
|||
|---|---|---|---|
| concentration (mM) | Cr(III) | Mn(II) | Co(II) |
| 5.000 | 0.02 | 0.091 | 0.066 |
| 2.500 | 0.08 | 0.120 | 0.067 |
| 1.250 | 0.04 | 0.074 | 0.112 |
| 0.625 | 0.014 | 0.053 | 0.042 |
| 0.313 | 1.074 | 0.037 | 0.940 |
| 0.156 | 0.981 | 0.022 | 0.985 |
| 0.078 | 1.040 | 0.817 | 1.242 |
3.9.4. DNA Degradation Activity
DNA degradation activity was assessed by using agarose gel electrophoresis. In the experiment, DNA alone and DNA with DMSO solvent were utilized as controls, and no damage to DNA was observed in lanes 1 and 2 (Figure S20). DNA cleavage activity gradually increased with increasing concentrations of the investigated substances. Figure S20 shows the results using ligand concentrations of 1, 3, and 5 mM and their complexes. The findings reveal a concentration-dependent influence on DNA cleavage. At a higher concentration of 5 mM, all the compounds exhibited complete DNA degradation. At a concentration of 3 mM, significant DNA cleavage activity was observed, with the highest effect for the Co(II) complex. Finally, at a low concentration of 1 mM, the examined compounds exhibited moderate DNA degradation activity. DNA intercalation refers to the insertion of planar molecules between the stacked base pairs of double-stranded genomic DNA, disrupting its native structure. This process leads to a notable decrease in the helical twist of the DNA structure. The observed differences in cleavage efficiency among the compounds might be attributed to variations in their DNA-binding affinities. Therefore, based on the demonstrated DNA-binding properties of the investigated samples, it was deemed pertinent to evaluate their potential anticancer properties.
3.9.5. Polyacrylamide Gel Electrophoresis (SDS-PAGE)
The effect of the examined samples on the BSA protein was assessed, and the results are shown in Figure S21. The disappearance of the band in the [Cr(L)(H2O)2Cl]·H2O complex suggests strong BSA binding and potential aggregation that prevents the protein from entering the gel, as observed in lane six. A significant effect is observed in lane four, which is assigned to the [CoL]·2H2O complex. A weak effect was observed for the [Mn(HL)] complex, while the H2L ligand displayed no effect compared to the control. The varied effects of metal complexes on BSA may be influenced by multiple factors, including the chemical structure of the complexes, the type of metal ions, and protein–metal interactions.
3.9.6. Anticancer Activity
The anticancer properties of the H2L Schiff base and its corresponding metal chelates were assessed utilizing the MTT assay, targeting HeLa, MCF-7, and HCT-116 cell lines. The isolated compounds were prepared at different concentrations (1.56, 3.125, 6.25, 12.5, 25, 50, and 100 μM) in DMSO solvent. The effects of the compounds on the three cell lines were evaluated based on the IC50 values, representing the drug concentration required to reduce cell viability by 50%. The percentages of relative cell viability at various concentrations are shown in Figure , and the IC50 values are tabulated in Table . The results were compared with those of doxorubicin, which served as the standard reference.
The [CoL]·2H2O complex exhibited very strong potency against the HeLa, HCT-116, and MCF-7 cell lines with IC50 values of 7.76 ± 0.4, 10.23 ± 0.8, and 6.88 ± 0.4 μM, respectively.
The [Mn(HL)] complex displayed very strong anticancer potency against the MCF-7 cell line with an IC50 of 9.46 ± 0.7 μM. Moreover, it demonstrated strong activity against HeLa and HCT-116 cell lines with IC50 values of 13.47 ± 1.2 and 15.56 ± 1.3 μM, respectively.
The [Cr(L)(H2O)2Cl]·H2O complex exhibited strong antitumor properties against MCF-7 cells, with an IC50 value of 17.79 ± 1.3 μM and moderate activities against HeLa and HCT-116 cell lines with IC50 values of 26.10 ± 1.8 and 22.49 ± 1.5 μM, respectively.
The H2L ligand exhibited moderate anticancer properties against HeLa and MCF-7 cells with IC50 values of 42.65 ± 2.5 and 29.74 ± 1.9 μM, respectively, while it exhibited weak activity against cancer HCT-116. According to the results, the anticancer potency of the tested compounds follows the order [CoL]·2H2O> [Mn(HL)]> [Cr(L)(H2O)2Cl]·H2O > H2L ligand.
8.

Cell viability of the studied compounds against the human cancer (A) HeLa, (B) HCT-116, and (C) MCF-7 cell lines.
5. Anticancer Activities of the Studied Compounds against Human Cell Lines.
| in vitro
cytotoxicity, IC50 (μM) |
|||
|---|---|---|---|
| compound | HeLa | HCT-116 | MCF-7 |
| doxorubicin | 5.57 ± 0.4 | 5.23 ± 0.3 | 4.17 ± 0.2 |
| H2L ligand | 42.65 ± 2.5 | 67.15 ± 3.6 | 29.74 ± 1.9 |
| [CoL]·2H2O | 7.76 ± 0.4 | 10.23 ± 0.8 | 6.88 ± 0.4 |
| [Cr(L)(H2O)2Cl]·H2O | 26.10 ± 1.8 | 22.49 ± 1.5 | 17.79 ± 1.3 |
| [Mn(HL)] | 13.47 ± 1.2 | 15.56 ± 1.3 | 9.46 ± 0.7 |
The results indicate that the metal complexes exhibit greater efficacy in inhibiting cancer cell proliferation than their corresponding free ligands. These findings suggest that the chelation of metal ions to H2L Schiff bases markedly improves their anticancer efficacy and alters their biological activity profile. Enhanced activity may be attributed to the influence of the metal ions, which increase the acidity of the ligand, thereby promoting the formation of strong hydrogen bonds. The high activity of the Co(II) and Mn(II) complexes in influencing cancer cell processes can be attributed to their adequate lipid solubility, which facilitates metal transport across cellular membranes and significant thermodynamic stability to ensure that they reach the target site intact. The reduced activity of the Cr(III) complex could be attributed to its low lipid solubility, which hinders the transport of the metal atom to the precise target site. As a result, the cancer cell wall remains unaffected, and no interference occurs with cellular processes, allowing normal cell activity to proceed uninterrupted.
3.10. Molecular Docking
3.10.1. Induced Fit Docking Simulation
Molecular docking simulations were performed to estimate the binding affinities of the synthesized compounds toward the target DNA helices of the selected receptors, providing detailed insights into their interaction profiles (Figures and S22–S25). The induced fit docking (IFD) protocol was employed, which accounts for both ligand and receptor flexibility. This iterative procedure involves initial Glide docking with softened potentials, followed by prime side-chain prediction and minimization, and subsequent redocking of the ligands into the refined receptor structures. The IFD methodology, which utilizes the Refinement module in Prime v 7.7, was selected to accurately predict the binding modes of the studied compounds within the active sites of 5IAE, 1YWN, and 3W2S (Tables and ). For HeLa and HCT-116 inhibition, the [Cr(L)(H2O)2Cl]·H2O complex exhibited significant negative G emodel values (−38.119 kcal/mol for HeLa and −53.954 kcal/mol for HCT-116), indicative of stable binding. These values were attributed to substantial electrostatic (G eCoul = −29.795 kcal/mol for HeLa and −21.777 kcal/mol for HCT-116) and van der Waals (G evdW = −18.418 kcal/mol for HeLa and −17.213 kcal/mol for HCT-116) energy contributions, as determined by the OPLS4 force field. Furthermore, favorable G score (−7.765 and −6.415 kcal/mol), RMSD (1.662 and 2.486 Å), hydrophobic (G lipo = −1.256 and −1.092 kcal/mol), and IFDscore (−559.43 and −618.63 kcal/mol) values confirmed the high binding affinity of [Cr(L)(H2O)2Cl]·H2O with the 5IAE and 1YWN receptors. Interaction analysis revealed a single aromatic ring interaction with TYR204 (5.09 Å) in HeLa and a hydrogen bond between CN and ARG840 (2.49 Å) in HCT-116 (Figure S22).
9.
3D poses of free ligand H2L with the (a) HeLa, (b) HCT-116, and (c) MCF-7 targets.
6. Molecular Docking IFD Scores of the Investigated Compounds against Selected Targets.
| compound | IFD score | RMSD (Å) | interactions | type | distance (Å) |
|---|---|---|---|---|---|
| HeLa (PDB ID: 5IAE) | |||||
| original ligand | –576.22 | 0.920 | PHE250→(CO) | H-bond | 1.88 |
| PHE250←(H2O)→(CO) | H-bond | 1.90. 1.73 | |||
| TRP214→(C–O–) | H-bond | 2.39 | |||
| SER65←(H2O)→(CO) | H-bond | 1.97, 1,93 | |||
| SER65→(C–O–) | H-bond | 2.34 | |||
| ARG64---(C–O–) | salt bridge | 2.96 | |||
| GLY122→(CO) | H-bond | 2.31 | |||
| GLY122←(H2O)→(CO) | H-bond | 1.74, 1.70 | |||
| GLN161→(CO) | H-bond | 1.89 | |||
| CYS163→(CO) | H-bond | 2.79 | |||
| NH→SER205 | H-bond | 2.38 | |||
| ARG207---(C–O–) | salt bridge | 2.98 | |||
| ARG207---(C–O–) | salt bridge | 3.94 | |||
| ARG207→(CO) | H-bond | 1.80 | |||
| NH→ARG207 | H-bond | 1.60 | |||
| ARG207←(H2O)→(C–O–) | H-bond | 1.98, 2.25 | |||
| H2L ligand | –555.87 | 1.157 | SER209→CN | H-bond | 2.34 |
| ARG207←(H2O)→(C–O–) | H-bond | 1.91, 1.97 | |||
| ARG207←(H2O)→(CO) | H-bond | 1.91, 1.84 | |||
| ARG207---(C–O–) | salt bridge | 2.84 | |||
| [Cr(L)(H2O)2Cl]·H2O | –559.43 | 1.662 | TYR204---Ar ring | π–π stacking | 5.09 |
| [CoL]·2H2O | –547.37 | 1.464 | TRP206---Ar ring | π–π stacking | 5.23 |
| ARG207---(C–N–) | salt bridge | 4.62 | |||
| [Mn(HL)] | –549.25 | 0.905 | ARG207→(C–O–) | H-bond | 2.40 |
| TRP206---Ar ring | π–π stacking | 3.81 | |||
| TRP206---Ar ring | π–π stacking | 4.11 | |||
| TRP206---Ar ring | π–cation | 5.66 | |||
| doxorubicin | –553 | 1.942 | TYR214→(H2O)→(OH) | H-bond | 1.81, 1.55 |
| TYR204←(H2O)→(CO) | H-bond | 1.81, 1.87 | |||
| OH→SER205 | H-bond | 2.07 | |||
| ARG207→(OH) | H-bond | 2.10 | |||
| HCT-116 (PDB ID: 1YWN) | |||||
| original ligand | –623.21 | 0.590 | NH→GLU883 | H-bond | 2.15 |
| NH→GLU883 | H-bond | 2.05 | |||
| ASP1044→(CO) | H-bond | 1.86 | |||
| NH2→GLU915 | H-bond | 1.80 | |||
| H2L ligand | –611.89 | 2.352 | ASP1044←H2O→(CN) | H-bond | 1.99, 1.87 |
| ASP1044→(CO) | H-bond | 2.57 | |||
| LYS866→H2O→(CN) | H-bond | 2.11, 1.87 | |||
| LYS866---(C–O–) | salt bridge | 2.85 | |||
| [Cr(L)(H2O)2Cl]·H2O | –618.63 | 2.486 | ARG840→(CN) | H-bond | 2.49 |
| [CoL]·2H2O | –606.82 | 2.845 | LYS866---(C–O–) | salt bridge | 3.63 |
| [Mn(HL)] | –608.94 | 2.647 | solvent exposure | ||
| doxorubicin | –607.98 | 1.913 | ASP1044→NH2 | H-bond | 1.98 |
| ASP1044←H2O→(C–O–) | H-bond | 1.92, 1.77 | |||
| OH→GLU915 | H-bond | 1.82 | |||
| MCF-7 (PDB ID: 3W2S) | |||||
| original ligand | –673.03 | 0.823 | MET893→(CN) | H-bond | 1.94 |
| CYS775→(H2O)→(CN) | H-bond | 2.48, 1.94 | |||
| THR854→(H2O)→(CN) | H-bond | 1.86, 2.39 | |||
| THR854←(H2O)→(CN) | H-bond | 2.62, 2.39 | |||
| NH→PHE856 | H-bond | 2.07 | |||
| PHE856---Ar ring | π–π stacking | 5.31 | |||
| NH→GLY857 | H-bond | 2.37 | |||
| LYS745→(CO) | H-bond | 1.67 | |||
| Cl→LEU788 | halogen bond | 3.09 | |||
| ligand | –658.63 | 1.812 | PHE856---Ar ring | π–π stacking | 4.86 |
| LYS745---Ar ring | π–cation | 3.64 | |||
| [Cr(L)(H2O)2Cl]·H2O | –658.17 | 1.308 | LYS745---(C–N–) | salt bridge | 2.82 |
| PHE723---Py-ring | π–π stacking | 5.37 | |||
| [CoL]·2H2O | –649.25 | 1.368 | solvent exposure | ||
| [Mn(HL)] | –663.18 | 1.278 | LYS745→ (C–O–) | H-bond | 3.09 |
| NH→GLY857 | H-bond | 1.97 | |||
| doxorubicin | –658.35 | 1.079 | LYS745→(C–O–) | H-bond | 1.84 |
| OH→GLY857 | H-bond | 1.74 | |||
| OH→GLU749 | H-bond | 1.87 | |||
| (NH3)→ASP855 | H-bond | 2.07 | |||
| (NH3)---ASP855 | salt bridge | 2.91 | |||
| (NH3)→ASN842 | H-bond | 2.73 | |||
| OH→ASN842 | H-bond | 1.85 | |||
7. Induced Fit Docking Scores of the Investigated Compounds against Selected Targets (kcal/mol).
| compound | G score | G evdW | G eCoul | G energy | G emodel | G hbond | G lipo | IFDscore |
|---|---|---|---|---|---|---|---|---|
| HeLa (PDB ID: 5IAE) | ||||||||
| original ligand | –17.346 | –37.314 | –79.701 | –117.014 | –244.430 | –4.547 | –1.354 | –576.220 |
| H2L ligand | –6.484 | –25.547 | –20.673 | –46.220 | –69.196 | –0.416 | –1.259 | –555.870 |
| [Cr(L)(H2O)2Cl]·H2O | –7.765 | –18.418 | –29.795 | –48.213 | –38.119 | –0.780 | –1.256 | –559.430 |
| [CoL]·2H2O | –4.599 | –32.146 | –0.454 | –32.601 | –42.487 | 0.000 | –0.958 | –547.370 |
| [Mn(HL)] | –4.974 | –27.744 | –10.759 | –38.503 | –44.176 | –0.223 | –1.735 | –549.250 |
| doxorubicin | –11.095 | –21.689 | –34.161 | –55.850 | –99.426 | –0.231 | –1.497 | –553.000 |
| HCT-116 (PDB ID: 1YWN) | ||||||||
| original ligand | –14.678 | –59.698 | –18.167 | –77.865 | –140.736 | –1.126 | –3.953 | –623.210 |
| H2L ligand | –7.615 | –38.024 | –6.092 | –44.116 | –63.179 | –0.456 | –1.985 | –611.890 |
| [Cr(L)(H2O)2Cl]·H2O | –6.415 | –17.213 | –21.777 | –38.990 | –53.954 | –0.563 | –1.092 | –618.630 |
| [CoL]·2H2O | –7.716 | –31.411 | –6.259 | –37.670 | –51.285 | 0.000 | –1.650 | –606.820 |
| [Mn(HL)] | –6.681 | –37.771 | –7.835 | –45.606 | –61.006 | 0.000 | –2.744 | –608.940 |
| doxorubicin | –8.996 | –43.242 | –13.954 | –57.196 | –59.170 | –0.045 | –1.926 | –607.980 |
| MCF-7 (PDB ID: 3W2S) | ||||||||
| original ligand | –14.377 | –73.002 | –22.734 | –95.737 | –154.482 | –1.119 | –5.353 | –673.030 |
| H2L ligand | –5.561 | –37.043 | –0.627 | –37.670 | –49.019 | –0.133 | –1.646 | –658.630 |
| [Cr(L)(H2O)2Cl]·H2O | –6.901 | –38.553 | –20.461 | –59.014 | –32.654 | –0.096 | –2.085 | –658.170 |
| [CoL]·2H2O | –4.081 | –42.893 | –1.996 | –44.889 | 33.088 | 0.000 | –2.332 | –649.250 |
| [Mn(HL)] | –12.402 | –40.857 | –10.445 | –51.302 | –75.511 | –0.520 | –2.114 | –663.180 |
| doxorubicin | –11.259 | –51.537 | –21.390 | –72.927 | –107.495 | –0.232 | –2.222 | –658.350 |
Glide score.
Glide van der Waals energy.
Glide Coulomb energy.
Glide energy.
Glide model energy.
Glide hydrogen bonding.
Glide lipophilic contact plus phobic attractive term in the glide score.
Induced fit docking score.
For MCF-7 inhibition, the free ligand H2L demonstrated a highly negative G emodel (−49.019 kcal/mol), driven by significant electrostatic (G eCoul = −0.627 kcal/mol) and van der Waals (G evdW = −37.043 kcal/mol) energy contributions, indicating stable binding. Additionally, favorable G score (−5.591 kcal/mol), RMSD (1.812 Å), hydrophobic (G lipo = −1.646 kcal/mol), and IFDscore (−658.63 kcal/mol) values indicated strong binding affinity with the 3W2S receptor. Interaction analysis showed π–π stacking with PHE856 (4.86 Å) and π–cation stacking with LYS745 (3.64 Å) (Figure c).
Although molecular docking is acknowledged as a valuable tool, it has inherent limitations due to its dependence on algorithms, scoring functions, and active site sensitivity. Moreover, it does not fully account for experimental factors, microbial characteristics, or drug transport mechanisms.
3.10.2. Docking Validation
Redocking experiments were performed to validate the docking protocol and assess its efficiency using identical docking parameters. The native peptide inhibitor in the 5IAE receptor exhibited precise binding within the active site pocket, characterized by three salt-bridge interactions with ARG64 and ARG207, and 13 hydrogen bonds involving residues PHE250, TRP214, SER65, ARG64, GLY122, GLN161, CYS163, SER205, and ARG207. The redocked native ligand in the 1YWN receptor formed six hydrogen bonds with distances ranging from 1.80 to 2.15 Å. The native inhibitor in the 3W2S receptor interacted with the active site pocket through seven hydrogen bonds, including three water-mediated interactions, a π–π stacking interaction between an aromatic ring and PHE856 (5.31 Å), and a halogen bond between Cl and LEU788 (3.9 Å). An overlay analysis of the redocked ligands with the native cocrystallized structures was conducted (Figure ). The redocked inhibitors bound to the 5IAE, 1YWN, and 3W2S receptors exhibited IFD scores of −576.22, −623.21, and −673.03 kcal mol–1, respectively, and RMSD values of 0.920, 0.590, and 0.823 Å, respectively. The low RMSD values of 0.920, 0.590, and 0.823 Å upon redocking confirm the reliability and accuracy of the chosen IFD protocol for the SIAE, 1YWN, and 3W2S receptors, respectively.
10.
Superimposition of the redocked original ligand (represented in gray) and the native cocrystallized ligand (depicted in green) within the binding pocket of the (a) HeLa, (b) HCT-116, and (c) MCF-7 receptors.
4. Conclusion
In summary, the effective synthesis and full characterization of H2L and its transition metal complexes were carried out through an extensive series of physicochemical, analytical, and spectroscopic investigations. Spectroscopic evidence confirmed the tetrahedral geometries of the Co(II) and Mn(II) complexes and octahedral coordination spheres of the Cr(III) complexes. Quantum chemical calculations with QTAIM and RDG provided a wealth of information on bonding and noncovalent interactions, which proved to be decisive for the stabilization of ligand–receptor complexes and rational drug design. Thermogravimetric analysis, supported by activation energy calculations, revealed variations in the thermal stabilities of the complexes. The optical band gap values also indicate that these materials are well-suited for solar energy collection as well as the assembly of photovoltaic devices. Owing to its pronounced DNA-binding affinity and superior anticancer activity, the Co(II) complex emerges as a promising candidate for in vivo application as an antitumor agent. Its potential mechanism involves the inhibition of DNA replication within malignant cells, thereby contributing to the suppression of tumor development. The complexes exhibited structure-dependent and selective interactions with bovine serum albumin (BSA). The rise in anticancer activity upon metal coordination arises from the increased ligand acidity involving strong hydrogen bonding interactions. Remarkably, computational docking demonstrated that the [Cr(L)(H2O)2Cl]·H2O complex exhibited high binding affinities with HeLa and HCT-116 DNA helices and a selective affinity of the free ligand with the MCF-7 cell line. These multifaceted findings affirm the promise of these compounds as multifunctional candidates for photovoltaic and therapeutic applications.
Supplementary Material
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.5c05921.
Chemicals, instruments, 1H NMR, 13C NMR, FT-IR, UV–vis, and mass spectra of the ligand and its complexes; figures and tables; and details on the XRD, thermal curves, kinetic data, and theoretical studies. (PDF)
Yasmeen G. Abou El-Reash: Data curation, project administration, and review and editing; Saja Abdulrahman Althobaiti: data curation; Sahar Abdalla: data curation; Gaber M. Abu El-Reash: conceptualization and formal analysis; and Mahdi A. Mohammed: methodology, writing original draft, software, visualization, and review and editing.
This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) (grant number IMSIU-DDRSP2501).
The authors declare no competing financial interest.
References
- Panicker R. R., Sivaramakrishna A.. Remarkably flexible 2,2′:6′,2″-terpyridines and their group 8–10 transition metal complexes – Chemistry and applications. Coord. Chem. Rev. 2022;459:214426. doi: 10.1016/j.ccr.2022.214426. [DOI] [Google Scholar]
- Kargar H., Fallah-Mehrjardi M., Munawar K. S.. Metal complexes incorporating tridentate ONO pyridyl hydrazone Schiff base ligands: Crystal structure, characterization and applications. Coord. Chem. Rev. 2024;501:215587. doi: 10.1016/j.ccr.2023.215587. [DOI] [Google Scholar]
- Basaran E., Sogukomerogullari H. G., Tılahun Muhammed M., Akkoc S.. Synthesis of Novel Cu(II), Co(II), Fe(II), and Ni(II) Hydrazone Metal Complexes as Potent Anticancer Agents: Spectroscopic, DFT, Molecular Docking, and MD Simulation Studies. ACS Omega. 2024;9(38):40172–40181. doi: 10.1021/acsomega.4c06202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deepa S., Mathangi N., Mudavath R., Shekhar I., Aparna A. V., Sarala Devi C.. Heteroleptic complexes of hydrazone Scaffold of picolinoyl N- oxide and 2,4 dihydroxy phenyl moieties; evaluation of antioxidant activity, DNA and protein binding properties and in vitro antiproliferation studies. Inorg. Chim. Acta. 2025;574:122391. doi: 10.1016/j.ica.2024.122391. [DOI] [Google Scholar]
- Argirova M., Cherneva E., Mihaylova R., Momekov G., Yancheva D.. New metal complexes of 1H-benzimidazole-2-yl hydrazones: Cytostatic, proapoptotic and modulatory activity on kinase signaling pathways. Arch. Biochem. Biophys. 2025;764:110245. doi: 10.1016/j.abb.2024.110245. [DOI] [PubMed] [Google Scholar]
- Yadav V. K., Nath M.. Diorganotin(IV) derivatives of salicylaldehyde hydrazones: Synthesis, structural characterization, crystal structure of dimethyltin(IV) complexes, and CT-DNA binding. Inorg. Chem. Commun. 2024;170:113283. doi: 10.1016/j.inoche.2024.113283. [DOI] [Google Scholar]
- Kanchanadevi S., Fronczek F. R., Immanuel David C., Nandhakumar R., Mahalingam V.. Investigation of DNA/BSA binding and cytotoxic properties of new Co(II), Ni(II) and Cu(II) hydrazone complexes. Inorg. Chim. Acta. 2021;526:120536. doi: 10.1016/j.ica.2021.120536. [DOI] [Google Scholar]
- Bashir M., Dar A. A., Yousuf I.. Syntheses, Structural Characterization, and Cytotoxicity Assessment of Novel Mn(II) and Zn(II) Complexes of Aroyl-Hydrazone Schiff Base Ligand. ACS Omega. 2023;8(3):3026–3042. doi: 10.1021/acsomega.2c05927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mohammed M. A., Fetoh A., Ali T. A., Youssef M. M., Abu El-Reash G. M.. Fabrication of novel Fe (III), Co (II), Hg (II), and Pd (II) complexes based on water-soluble ligand (NaH2PH): structural characterization, cyclic voltammetric, powder X-ray diffraction, zeta potential, and biological studies. Appl. Organomet. Chem. 2023;37(1):e6910. doi: 10.1002/aoc.6910. [DOI] [Google Scholar]
- Grant R. S., Coggan S. E., Smythe G. A.. The Physiological Action of Picolinic Acid in the Human Brain. Int. J. Tryptophan Res. 2009;2:71–79. doi: 10.4137/IJTR.S2469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pathak N., Ekta R., Nitesh K., Suvarna G. K., Rao C. M.. A Review on Anticancer Potentials of Benzothiazole Derivatives. Mini-Rev. Med. Chem. 2020;20(1):12–23. doi: 10.2174/1389557519666190617153213. [DOI] [PubMed] [Google Scholar]
- Moreira J. M., Vieira S. d. S. F., Correia G. d. D., de Almeida L. N., Finoto S., Brandl C. A., Msumange A. A., Galvão F., Pires de Oliveira K. M., Caneppele Paveglio G.. et al. Synthesis and Characterization of Novel Hydrazone Complexes: Exploring DNA/BSA Binding and Antimicrobial Potential. ACS Omega. 2025;10(7):7428–7440. doi: 10.1021/acsomega.5c00069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parvarinezhad S., Ramezanipoor S., Kubicki M., Salehi M.. Zn(II), Mn(III), and Co(III) complexes of Schiff base derived from 2-hydroxy-1-naphthaldehyde: Synthesis, spectral surveys, single crystal structure studies, Hirshfeld surface analysis, density functional theory calculation, molecular electrostatic potential, and molecular docking evaluations. Appl. Organomet. Chem. 2024;38(6):e7477. doi: 10.1002/aoc.7477. [DOI] [Google Scholar]
- Cefalu W. T., Wang Z. Q., Zhang X. H., Baldor L. C., Russell J. C.. Oral Chromium Picolinate Improves Carbohydrate and Lipid Metabolism and Enhances Skeletal Muscle Glut-4 Translocation in Obese, Hyperinsulinemic (JCR-LA Corpulent) Rats. J. Nutr. 2002;132(6):1107–1114. doi: 10.1093/jn/132.6.1107. [DOI] [PubMed] [Google Scholar]
- Ahmad H., Ahmed Z., Khan R.. Effect of Chromium Picolinate Supplementation on Diabetic Profile and Nutritional Status of the Type-2 Diabetic Adult Population ¨C A Randomized Controlled Trial. J. Food Nutr. Res. 2016;4(8):535–542. doi: 10.12691/JFNR-4-8-8. [DOI] [Google Scholar]
- Georgaki M.-N., Tsokkou S., Keramas A., Papamitsou T., Karachrysafi S., Kazakis N.. Chromium supplementation and type 2 diabetes mellitus: an extensive systematic review. Environ. Geochem. Health. 2024;46(12):515. doi: 10.1007/s10653-024-02297-5. [DOI] [PubMed] [Google Scholar]
- Deepa S., Nayakini A., Mudavath R., Lakshmi P. J., Sudeepa K., Devi C. S.. Novel Tridentate Hydrazone of Heterocyclic N-oxide and Its Transition Metal Complexes: Evaluation of DNA/BSA Binding Interactions, In Vitro and In Silico Biological Properties. ChemistrySelect. 2025;10(5):e202404749. doi: 10.1002/slct.202404749. [DOI] [Google Scholar]
- Toyama M., Onishi Y., Tanaka N., Nagao N.. Complexes exhibiting trans(Cl)-RuCl2(dmso-S)2 geometry with methyl-picolinate-type neutral N,O-ligands: Synthesis, structural characterization, and chemical behavior analysis in aqueous solutions. J. Mol. Struct. 2025;1321:139964. doi: 10.1016/j.molstruc.2024.139964. [DOI] [Google Scholar]
- Abdullah T. B., Behjatmanesh-Ardakani R., Faihan A. S., Jirjes H. M., Abou-Krisha M. M., Yousef T. A., Kenawy S. H., Al-Janabi A. S.. Cd (II) and Pd (II) Mixed Ligand Complexes of Dithiocarbamate and Tertiary Phosphine LigandsSpectroscopic, Anti-Microbial, and Computational Studies. Molecules. 2023;28(5):2305. doi: 10.3390/molecules28052305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chai L.-Q., Zhang X.-F., Tang L.-J.. Crystallographic, spectroscopic, TD/DFT calculations and Hirshfeld surface analysis of cadmium (II) coordination polymer containing pyridine ring. J. Mol. Struct. 2021;1245:131028. doi: 10.1016/j.molstruc.2021.131028. [DOI] [Google Scholar]
- Kitts D. D., Wijewickreme A. N., Hu C.. Antioxidant properties of a North American ginseng extract. Mol. Cell. Biochem. 2000;203(1–2):1–10. doi: 10.1023/A:1007078414639. [DOI] [PubMed] [Google Scholar]
- Damena T., Demissie T. B., Bitew Alem M., Zeleke D., Desalegn T.. In vitro biological activities of novel Ni(II)-2-chloroquinoline derivative complex: Synthesis, characterization and computational studies. Results Chem. 2023;6:101090. doi: 10.1016/j.rechem.2023.101090. [DOI] [Google Scholar]
- Damena T., Desalegn T., Mathura S., Getahun A., Bizuayehu D., Alem M. B., Gadisa S., Zeleke D., Demissie T. B.. Synthesis, Structural Characterization, and Computational Studies of Novel Co(II) and Zn(II) Fluoroquinoline Complexes for Antibacterial and Antioxidant Activities. ACS Omega. 2024;9(34):36761–36777. doi: 10.1021/acsomega.4c05560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Damena T., Zeleke D., Desalegn T., Demissie T. B., Eswaramoorthy R.. Synthesis, Characterization, and Biological Activities of Novel Vanadium(IV) and Cobalt(II) Complexes. ACS Omega. 2022;7(5):4389–4404. doi: 10.1021/acsomega.1c06205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Damena T., Alem M. B., Zeleke D., Demissie T. B., Desalegn T.. Synthesis and computational studies of novel cobalt(II) and oxovanadium(IV) complexes of quinoline carbaldehyde derivative ligand for antibacterial and antioxidant applications. J. Mol. Struct. 2023;1280:134994. doi: 10.1016/j.molstruc.2023.134994. [DOI] [Google Scholar]
- Parejo I., Codina C., Petrakis C., Kefalas P.. Evaluation of scavenging activity assessed by Co(II)/EDTA-induced luminol chemiluminescence and DPPH* (2,2-diphenyl-1-picrylhydrazyl) free radical assay. J. Pharmacol Toxicol. Methods. 2000;44(3):507–512. doi: 10.1016/S1056-8719(01)00110-1. [DOI] [PubMed] [Google Scholar]
- Gupta S., Prakash A., Savita S., Satya, Hashmi K., Mishra P., Veg E., Khan T., Patel D. K., Joshi S.. Studies of some Mixed Ligand-Metal complexes of 4-((2-(phenylcarbamothioyl)hydrazinylidene) methyl) benzoic acid in search of potential antimicrobial activity. J. Mol. Struct. 2024;1318:139319. doi: 10.1016/j.molstruc.2024.139319. [DOI] [Google Scholar]
- Youssef M. M., Al-Omair M. A., Ismail M. A.. Synthesis, DNA affinity, and antimicrobial activity of 4-substituted phenyl-2,2′-bichalcophenes and aza-analogues. Med. Chem. Res. 2012;21(12):4074–4082. doi: 10.1007/s00044-011-9964-y. [DOI] [Google Scholar]
- Parvekar P., Palaskar J., Metgud S., Maria R., Dutta S. J. B. I. i. D.. The minimum inhibitory concentration (MIC) and minimum bactericidal concentration (MBC) of silver nanoparticles against Staphylococcus aureus. Biomater. Investig. Dent. 2020;7(1):105–109. doi: 10.1080/26415275.2020.1796674. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang H., Yao H., Ni R., Wang R., Ren J., Qiao H., Zhang Y., Zhang Z., Wang J.. Insights into interaction of quaternary ammonium salt cationic surfactants with different branched-chain lengths and DNA: Multi-spectral analysis, viscosity method, and gel electrophoresis. Int. J. Biol. Macromol. 2025;299:140095. doi: 10.1016/j.ijbiomac.2025.140095. [DOI] [PubMed] [Google Scholar]
- Adoui F., Saadi S., Lemmadi S., Benelouezzane C., Guemra I., Boughellout H., Benatallah L., Nedjar N.. Unravelling the behaviour of camel milk caseins during hydrolysis and separation: Insights into proteomic profiles of caseins using high performance liquid chromatography and polyacrylamide gel electrophoresis. Food Humanity. 2025;4:100491. doi: 10.1016/j.foohum.2024.100491. [DOI] [Google Scholar]
- Devi P., Singh K., Kumar B., Singh J. K.. Synthesis, spectroscopic, antimicrobial and in vitro anticancer activity of Co+2, Ni+2, Cu+2 and Zn+2 metal complexes with novel Schiff base. Inorg. Chem. Commun. 2023;152:110674. doi: 10.1016/j.inoche.2023.110674. [DOI] [Google Scholar]
- Sharfalddin A. A., Inas Muta’eb Alyounis E., Emwas A.-H., Jaremko M.. Biological efficacy of novel metal complexes of Nitazoxanide: Synthesis, characterization, anti-COVID-19, antioxidant, antibacterial and anticancer activity studies. J. Mol. Liq. 2022;368:120808. doi: 10.1016/j.molliq.2022.120808. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schrödinger Suite 2025-1. LigPrep; Schrödinger, LL: New York, NY, 2025. [Google Scholar]
- Maciag J. J., Mackenzie S. H., Tucker M. B., Schipper J. L., Swartz P., Clark A. C.. Tunable allosteric library of caspase-3 identifies coupling between conserved water molecules and conformational selection. Proc. Natl. Acad. Sci. U.S.A. 2016;113(41):E6080–E6088. doi: 10.1073/pnas.1603549113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miyazaki Y., Matsunaga S., Tang J., Maeda Y., Nakano M., Philippe R. J., Shibahara M., Liu W., Sato H., Wang L., Nolte R. T.. Novel 4-amino-furo[2,3-d]pyrimidines as Tie-2 and VEGFR2 dual inhibitors. Bioorg. Med. Chem. Lett. 2005;15(9):2203–2207. doi: 10.1016/j.bmcl.2005.03.034. [DOI] [PubMed] [Google Scholar]
- Sogabe S., Kawakita Y., Igaki S., Iwata H., Miki H., Cary D. R., Takagi T., Takagi S., Ohta Y., Ishikawa T.. Structure-Based Approach for the Discovery of Pyrrolo[3,2-d]pyrimidine-Based EGFR T790M/L858R Mutant Inhibitors. ACS Med. Chem. Lett. 2013;4(2):201–205. doi: 10.1021/ml300327z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lu C., Wu C., Ghoreishi D., Chen W., Wang L., Damm W., Ross G. A., Dahlgren M. K., Russell E., Von Bargen C. D.. et al. OPLS4: Improving Force Field Accuracy on Challenging Regimes of Chemical Space. J. Chem. Theory Comput. 2021;17(7):4291–4300. doi: 10.1021/acs.jctc.1c00302. [DOI] [PubMed] [Google Scholar]
- Sherman W., Beard H. S., Farid R.. Use of an Induced Fit Receptor Structure in Virtual Screening. Chem. Biol. Drug Des. 2006;67(1):83–84. doi: 10.1111/j.1747-0285.2005.00327.x. [DOI] [PubMed] [Google Scholar]
- Jacobson M. P., Pincus D. L., Rapp C. S., Day T. J. F., Honig B., Shaw D. E., Friesner R. A.. A Hierarchical Approach to All-Atom Protein Loop Prediction. Proteins. 2004;55(2):351–367. doi: 10.1002/prot.10613. [DOI] [PubMed] [Google Scholar]
- Thanh N. D., Hai D. S., Thi Huyen L., Giang N. T. K., Thu Ha N. T., Tung D. T., Thi Le C., Van H. T. K., Toan V. N.. Synthesis and in vitro anticancer activity of 4H-pyrano[2,3-d]pyrimidine–1H-1,2,3-triazole hybrid compounds bearing D-glucose moiety with dual EGFR/HER2 inhibitory activity and induced fit docking study. J. Mol. Struct. 2023;1271:133932. doi: 10.1016/j.molstruc.2022.133932. [DOI] [Google Scholar]
- Lu F., Astruc D.. Nanocatalysts and other nanomaterials for water remediation from organic pollutants. Coord. Chem. Rev. 2020;408:213180. doi: 10.1016/j.ccr.2020.213180. [DOI] [Google Scholar]
- Colthup, N. B. ; Daly, L. ; Wiberley, S. E. . Introduction to Infrared and Raman Spectroscopy; Academic Press, 1975. [Google Scholar]
- Bellamy, L. J. The Infrared Spectra of Complex Molecule; Chapman and Hall, 1975. [Google Scholar]
- Tolan D. A., Kashar T. I., Yoshizawa K., El-Nahas A. M.. Synthesis, spectral characterization, density functional theory studies, and biological screening of some transition metal complexes of a novel hydrazide–hydrazone ligand of isonicotinic acid. Appl. Organomet. Chem. 2021;35(6):e6205. doi: 10.1002/aoc.6205. [DOI] [Google Scholar]
- Zahirović A., Osmanković I., Osmanović A., Višnjevac A., Magoda A., Hadžalić S., Kahrović E.. Interaction of Copper(II) Complexes of Bidentate Benzaldehyde Nicotinic Acid Hydrazones with BSA: Spectrofluorimetric and Molecular Docking Approach. Acta Chim. Slov. 2023;70(1):74–85. doi: 10.17344/acsi.2022.7826. [DOI] [PubMed] [Google Scholar]
- Mahmoud M. A., Sallam S. A., Ads A. M., El-Nour K. M. A.. Assessing the antimicrobial and cytotoxic activities of VO(IV), Cr(III), Cu(II)-metformin-schiff-base’s complexes enhanced by gold nanoparticles. J. Mol. Struct. 2024;1312:138559. doi: 10.1016/j.molstruc.2024.138559. [DOI] [Google Scholar]
- Al-Farraj E. S., Fetoh A.. Synthesis of new Fe(III), Co(II), and Cr(III) complexes of N-(benzo[d]thiazol-2-ylcarbamothioyl)benzamide (H2L2): Structural characterization and biological activities. Appl. Organomet. Chem. 2023;37(11):e7248. doi: 10.1002/aoc.7248. [DOI] [Google Scholar]
- Fetoh A., Mohammed M. A., Youssef M. M., El-Reash G. M. A.. Investigation (IR, UV–visible, fluorescence, X-ray diffraction and thermogravimetric) studies of Mn(II), Fe(III) and Cr(III) complexes of thiosemicarbazone derived from 4- pyridyl thiosemicarbazide and monosodium 5-sulfonatosalicylaldehyde and evaluation of their biological applications. J. Mol. Struct. 2023;1271:134139. doi: 10.1016/j.molstruc.2022.134139. [DOI] [Google Scholar]
- Zihon M., Kassa A., Tigineh G. T., Chanie G., Tesfaye D., Gebrezgiabher M., Metto M., Alem M. B., Abebe A., Thomas M.. Manganese(II) resorcinolate complex: synthesis, characterizations, electrochemical behavior, and antibacterial activities. J. Appl. Electrochem. 2025;55:727–737. doi: 10.1007/s10800-024-02194-w. [DOI] [Google Scholar]
- Khedr A. M., Mandour H. S. A., Wahdan K. M., El-Ghamry H. A.. New nanometric metal complexes based on thiazole derivative as potential antitumor and antimicrobial agents: Full structural elucidation, docking simulation and biological activity studies. J. Mol. Liq. 2024;401:124665. doi: 10.1016/j.molliq.2024.124665. [DOI] [Google Scholar]
- Venkatesh G., Vennila P., Kaya S., Ahmed S. B., Sumathi P., Siva V., Rajendran P., Kamal C.. Synthesis and Spectroscopic Characterization of Schiff Base Metal Complexes, Biological Activity, and Molecular Docking Studies. ACS Omega. 2024;9(7):8123–8138. doi: 10.1021/acsomega.3c08526. [DOI] [PMC free article] [PubMed] [Google Scholar]
- El-Metwaly N., Farghaly T. A., Elghalban M. G.. Synthesis, analytical and spectral characterization for new VO (II)-triazole complexes; conformational study beside MOE docking simulation features. Appl. Organomet. Chem. 2020;34(4):e5505. doi: 10.1002/aoc.5505. [DOI] [Google Scholar]
- Mohammed M. A., Fetoh A., Youssef M. M., Abu El-Reash G. M., El-Reash Y. G. A., Ali T. A.. Synthesis of a new Cd (II) complex as ionophores and its application in construction of a highly selective and sensitive cadmium (II) sensor in petroleum water samples. Appl. Organomet. Chem. 2023;37(8):e7186. doi: 10.1002/aoc.7186. [DOI] [Google Scholar]
- Alayyafi A. A., Diab M. A., El-Sonbati A. Z., Negm E., El-Zahed M. M., Morgan S. M.. Synthesis of metal(II) acetate complexes with a heterocyclic azo dye ligand: Characterization, geometrical molecular structures, antioxidant, biological activities, DNA binding and molecular docking studies. J. Mol. Struct. 2025;1322:140223. doi: 10.1016/j.molstruc.2024.140223. [DOI] [Google Scholar]
- Mohammed M. A., Ali T. A., Youssef M. M., Abou El-Reash Y. G., Al-Farraj E. S., Adam F. A., El-Moneim A. A., Abu El-Reash G. M., Fetoh A.. Elucidation, cyclic voltammetry, computational studies, and biological activities of Ni(II), Cu(II), Cd(II), and Ce(III) complexes with a water-soluble isonicotinohydrazide-based Schiff base ligand. Appl. Organomet. Chem. 2024;38(9):e7618. doi: 10.1002/aoc.7618. [DOI] [Google Scholar]
- Elsherbeny M., El-Ayaan U., Abou El-Reash Y. G., Abu El-Reash G. M.. Characterization, DFT, DNA, Biological Applications, and In Silico Studies of Mn(II), VO(II), Pd(II), and Cd(II) Complexes Derived From Chromone Isonicotinic Hydrazone Ligand. Appl. Organomet. Chem. 2025;39(2):e7822. doi: 10.1002/aoc.7822. [DOI] [Google Scholar]
- Belal D. M., El-Ayaan U. I., El-Gamil M. M., Younis A. M., El-Reash G. M. A.. Fluorescence, cyclic voltammetric, computational, and spectroscopic studies of Mn(II), Co(II), Pd(II), Zn(II) and Cd(II) complexes of salen ligand and their biological applications. J. Mol. Struct. 2023;1271:134142. doi: 10.1016/j.molstruc.2022.134142. [DOI] [Google Scholar]
- Fayed T. A., Gaber M., El-Nahass M. N., Diab H. A., El-Gamil M. M.. Synthesis, Structural characterization, thermal, molecular modeling and biological studies of chalcone and Cr(III), Mn(II), Cu(II) Zn(II) and Cd(II) chelates. J. Mol. Struct. 2020;1221:128742. doi: 10.1016/j.molstruc.2020.128742. [DOI] [Google Scholar]
- Coats A. W., Redfern J.. Kinetic parameters from thermogravimetric data. Nature. 1964;201(4914):68–69. doi: 10.1038/201068a0. [DOI] [Google Scholar]
- Horowitz H. H., Metzger G.. A new analysis of thermogravimetric traces. Anal. Chem. 1963;35(10):1464–1468. doi: 10.1021/ac60203a013. [DOI] [Google Scholar]
- Frost A., Pearson R.. Kinetics and mechanism. J. Phys. Chem. A. 1961;65(2):384. doi: 10.1021/j100820a601. [DOI] [Google Scholar]
- Kenawy I. M. M., Hafez M. A. H., Lashein R. R.. Thermal Decomposition of Chloromethylated Poly(styrene)-PAN Resin and Its Complexes with Some Transition Metal Ions. J. Therm. Anal. Calorim. 2001;65(3):723–736. doi: 10.1023/A:1011999325998. [DOI] [Google Scholar]
- Fayed T. A., Gaber M., Abu El-Reash G. M., El-Gamil M. M.. Structural, DFT/B3LYP and molecular docking studies of binuclear thiosemicarbazide Copper (II) complexes and their biological investigations. Appl. Organomet. Chem. 2020;34(9):e5800. doi: 10.1002/aoc.5800. [DOI] [Google Scholar]
- Maravalli P. B., Goudar T. R.. Thermal and spectral studies of 3-N-methyl-morpholino-4-amino-5-mercapto-1,2,4-triazole and 3-N-methyl-piperidino-4-amino-5-mercapto-1,2,4-triazole complexes of cobalt(II), nickel(II) and copper(II) Thermochim. Acta. 1999;325(1):35–41. doi: 10.1016/S0040-6031(98)00548-6. [DOI] [Google Scholar]
- Yusuff K. K. M., Sreekala R.. Thermal and spectral studies of 1-benzyl-2- phenylbenzimidazole complexes of cobalt(II) Thermochim. Acta. 1990;159:357–368. doi: 10.1016/0040-6031(90)80121-E. [DOI] [Google Scholar]
- El-Gammal O. A., Abu El-Reash G. M., El-Gamil M. M.. Binuclear copper(II), cobalt(II) and Nickel(II) complexes of N1-ethyl-N2-(pyridin-2-yl) hydrazine-1,2-bis(carbothioamide): Structural, spectral, pH-metric and biological studies. Spectrochim. Acta, Part A. 2012;96:444–455. doi: 10.1016/j.saa.2012.05.046. [DOI] [PubMed] [Google Scholar]
- Hussain R., Rubab S. L., Maryam A., Ashraf T., Arshad M., Lal K., Sumrra S. H., Ashraf S., Ali B.. Synthesis, Spectroscopic and Nonlinear Optical Properties, and Antimicrobial Activity of Cu(II), Co(II), and Ni(II) Complexes: Experimental and Theoretical Studies. ACS Omega. 2023;8(45):42598–42609. doi: 10.1021/acsomega.3c05322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ayisha Begam K., Kanagathara N., Marchewka M. K., Lo A.-Y.. DFT, hirshfeld and molecular docking studies of a hybrid compound - 2,4-Diamino-6-methyl-1,3,5-triazin-1-ium hydrogen oxalate as a promising anti -breast cancer agent. Heliyon. 2022;8(8):e10355. doi: 10.1016/j.heliyon.2022.e10355. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pandey N., Mehata M. S., Pant S., Tewari N.. Structural, Electronic and NLO Properties of 6-aminoquinoline: A DFT/TD-DFT Study. J. Fluoresc. 2021;31(6):1719–1729. doi: 10.1007/s10895-021-02788-z. [DOI] [PubMed] [Google Scholar]
- Chatterjee S., Afzal M., Mandal P. C., Modak R., Guin M., Konar S.. Exploration of supramolecular interactions, Hirshfeld surface, FMO, molecular electrostatic potential (MEP) analyses of pyrazole based Zn(II) complex. J. Indian Chem. Soc. 2024;101(10):101275. doi: 10.1016/j.jics.2024.101275. [DOI] [Google Scholar]
- Zinatloo-Ajabshir S., Rakhshani S., Mehrabadi Z., Farsadrooh M., Feizi-Dehnayebi M., Rakhshani S., Dušek M., Eigner V., Rtimi S., Aminabhavi T. M.. Novel rod-like [Cu(phen)2(OAc)]·PF6 complex for high-performance visible-light-driven photocatalytic degradation of hazardous organic dyes: DFT approach, Hirshfeld and fingerprint plot analysis. J. Environ. Manage. 2024;350:119545. doi: 10.1016/j.jenvman.2023.119545. [DOI] [PubMed] [Google Scholar]
- Nasertayoob P., Shahbazian S.. Revisiting the foundations of quantum theory of atoms in molecules (QTAIM): The variational procedure and the zero-flux conditions. Int. J. Quantum Chem. 2008;108(9):1477–1484. doi: 10.1002/qua.21665. [DOI] [Google Scholar]
- Bakheit A. H., Al-Salahi R., Al-Majed A. A.. Thermodynamic and Computational (DFT) Study of Non-Covalent Interaction Mechanisms of Charge Transfer Complex of Linagliptin with 2,3-Dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) and Chloranilic acid (CHA) Molecules. 2022;27(19):6320. doi: 10.3390/molecules27196320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arulaabaranam K., Muthu S., Mani G., Ben Geoffrey A. S.. Speculative assessment, molecular composition, PDOS, topology exploration (ELF, LOL, RDG), ligand-protein interactions, on 5-bromo-3-nitropyridine-2-carbonitrile. Heliyon. 2021;7(5):e07061. doi: 10.1016/j.heliyon.2021.e07061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sagaama A., Issaoui N., Al-Dossary O., Kazachenko A. S., Wojcik M. J.. Non covalent interactions and molecular docking studies on morphine compound. J. King Saud Univ. Sci. 2021;33(8):101606. doi: 10.1016/j.jksus.2021.101606. [DOI] [Google Scholar]
- Sogukomerogullari H. G., Başaran E., Kepekçi R. A., Türkmenoğlu B., Sarıoğlu A. O., Köse M.. Novel europium(III), terbium(III), and gadolinium(III) Schiff base complexes: Synthesis, structural, photoluminescence, antimicrobial, antioxidant, and molecular docking studies. Polyhedron. 2025;265:117275. doi: 10.1016/j.poly.2024.117275. [DOI] [Google Scholar]
- Ahmed Y. M., Elgendi M. A., Omar M. M., Mohamed G. G., Deghadi R. G.. Synthesis, characterization, antimicrobial, antioxidant studies, molecular docking and DFT calculations of novel Schiff base and its metal complexes. J. Mol. Struct. 2025;1326:141076. doi: 10.1016/j.molstruc.2024.141076. [DOI] [Google Scholar]
- Elkanzi N. A. A., Alsirhani A. M., Ali A. M., Abdou A., Alali I.. Design, characterization, antioxidant activity, and molecular docking study of sulfadiazine derivative-based Ni (II) and Cu (II) complexes. J. Mol. Liq. 2025;423:126937. doi: 10.1016/j.molliq.2025.126937. [DOI] [Google Scholar]
- Swiderski G., Jabłońska-Trypuć A., Kalinowska M., Swisłocka R., Karpowicz D., Magnuszewska M., Lewandowski W.. Spectroscopic, Theoretical and Antioxidant Study of 3d-Transition Metals (Co(II), Ni(II), Cu(II), Zn(II)) Complexes with Cichoric Acid. Materials. 2020;13(14):3102. doi: 10.3390/ma13143102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jabłońska-Trypuć A., Wydro U., Wołejko E., Świderski G., Lewandowski W.. Biological Activity of New Cichoric Acid–Metal Complexes in Bacterial Strains, Yeast-Like Fungi, and Human Cell Cultures In Vitro. Nutrients. 2020;12(1):154. doi: 10.3390/nu12010154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kostova I., Balkansky S.. Metal complexes of biologically active ligands as potential antioxidants. Curr. Med. Chem. 2013;20(36):4508–4539. doi: 10.2174/09298673113206660288. [DOI] [PubMed] [Google Scholar]
- Yavuz Ş., Köse D. A., Özkinali S., Tatar D., Veyisoğlu A.. Synthesis and characterization of 4-hydroxy-3-((2-hydroxy-3-methoxybenzylidene)amino)benzene-sulfonic acid and its metal complexes and their antimicrobial activity. J. Mol. Struct. 2025;1322:140364. doi: 10.1016/j.molstruc.2024.140364. [DOI] [Google Scholar]
- Daisylet B. S., Raphael S. J., Kumar P., Mohan A., Dasan A.. Exploring photocatalytic, antimicrobial, and biofilm inhibition properties of cuminaldehyde derived benzimidazole metal complexes. J. Mol. Struct. 2025;1328:141311. doi: 10.1016/j.molstruc.2025.141311. [DOI] [Google Scholar]
- Ismail M. A., Youssef M. M., Arafa R. K., Al-Shihry S. S., El-Sayed W. M.. Synthesis and antiproliferative activity of monocationic arylthiophene derivatives. Eur. J. Med. Chem. 2017;126:789–798. doi: 10.1016/j.ejmech.2016.12.007. [DOI] [PubMed] [Google Scholar]
- Todorov L. T., Kostova I. P.. Coumarin-transition metal complexes with biological activity: current trends and perspectives. Front. Chem. 2024;12:2296–2646. doi: 10.3389/fchem.2024.1342772. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Czarnomysy R., Radomska D., Muszyńska A., Hermanowicz J. M., Prokop I., Bielawska A., Bielawski K.. Evaluation of the Anticancer Activities of Novel Transition Metal Complexes with Berenil and Nitroimidazole. Molecules. 2020;25(12):2860. doi: 10.3390/molecules25122860. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sumi M., Nevaditha N. T., Sindhu Kumari B.. Synthesis, structural evaluation, antioxidant, DNA cleavage, anticancer activities and molecular docking study of metal complexes of 2-amino thiophene derivative. J. Mol. Struct. 2023;1272:134091. doi: 10.1016/j.molstruc.2022.134091. [DOI] [Google Scholar]
- Al-Farraj E. S., Mohammed M. A.. Comprehensive Exploration of Water-Soluble Schiff Base Complexes of Ni (II), Cu (II), Mn (II), and Ce (III): Electrochemical, Computational, and Biological Studies. Appl. Organomet. Chem. 2025;39(4):e70122. doi: 10.1002/aoc.70122. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.







