Skip to main content
ACS Omega logoLink to ACS Omega
. 2023 Apr 4;8(15):13605–13625. doi: 10.1021/acsomega.2c07592

Spectroscopic Characterization, Cyclic Voltammetry, Biological Investigations, MOE, and Gaussian Calculations of VO(II), Cu(II), and Cd(II) Heteroleptic Complexes

Anwer G Al-Harazie †,‡,*, Esam A Gomaa , Rania R Zaky , Mahmoud N Abd El-Hady
PMCID: PMC10116629  PMID: 37091434

Abstract

graphic file with name ao2c07592_0027.jpg

A novel hydrazone ligand (o-H2BMP) N-(benzo[d]thiazol-2-yl)-3-oxo-3-(2-(1-(pyridin-2-yl)ethylidene)hydrazinyl)propanamide alongside its Cu(II), Cd(II), and VO(II) complexes were prepared and structurally characterized via various spectroscopic analyses (Fourier transform infrared spectroscopy, UV–visible spectroscopy, 1H/13C NMR spectroscopy, liquid chromatography coupled to mass spectrometry, and electron paramagnetic resonance spectroscopy) as well as by elemental analysis, thermal gravimetry analysis/differential thermal analysis, and magnetic moment measurements. Powder X-ray diffraction analysis was also performed for the free ligand and its metal complexes to determine the crystallographic structures and atomic spacing. It also provided information on unit cell dimensions and the average crystallite size. Furthermore, geometric optimization and computational studies were carried out by applying Gaussian (09) software based on density-functional theory coupled with the B3LYP functional and LANL2DZ/6-31+G(d,p) mixed basis set to evaluate some distinct features such as molecular electrostatic potential, EHOMO, and ELUMO. Moreover, electrochemical measurements were performed for Cu(II) in the absence/presence of the chelating agent to predict the effect of complexation interaction in the solution state study. As part of the biological examination, antioxidant and antimicrobial assays were conducted for each compound individually, in addition to cytotoxicity evaluations via MTT assays for all isolated complexes compared to the corresponding metal salts. The MOE (molecular operating environment) approach was also applied to model the interface between the isolated compounds and proteins that were expressed in breast cancer at the atomic level.

1. Introduction

Hydrazones belong to the Schiff base family and have played an imperative role in the development of coordination chemistry owing to their chelating properties.13 Generally, hydrazones are characterized by an imine group, providing information about the mechanism as well as the interaction in the biochemical system, where these compounds have a significant impact in pharmaceutical fields due to their wide bioactive efficacy.4 As a result of these significant biological applications of metal complexes, they have received a great deal of attention and are considered models of biologically relevant organisms. An extensive range of biological activities were demonstrated for these compounds, including antidiabetic, antifungal, antibacterial, anticancer, antitumor, and antiproliferative activities.5

In particular, 2-acetylpyridine hydrazone derivatives of benzothiazole exhibit a potent toxic effect on lymphomas, leukemias, and breast tumors.6,7 Also, Cu(II) complexes are more potent against rheumatoid arthritis and ulcers due to their role as acidic anti-inflammatory supplementary agents in gastrointestinal damage.8 Moreover, VO(II) complexes showed a very similar strength of activity to standard drugs such as ciprofloxacin (an antibacterial drug) and clotrimazole (an antifungal drug).9 Further, Cd(II) complexes with some hydrazone derivatives, such as ethyl-2-[(pyridine-2-ylmethylene)hydrazino] acetate, showed very similar biological activity to ampicillin as an antibacterial drug.10

In the present study, we aim to synthesize transition-metal complexes of VO(II), Cu(II), and Cd(II) cations using a novel N-(benzothiazol-2-yl)-3-oxo-3-(2-(1-(pyridin-2-yl)ethylidene)hydrazinyl)propanamide in conjunction with experimental and theoretical characterizations. In addition to providing qualitative information about the redox process and thermodynamic parameters, cyclic voltammetry (CV) was also employed to analyze the Cu(II) species. Finally, a biological investigation was conducted, which included tests for antimicrobial, antioxidant, and MTT activity. Molecular docking allowed us to predict the behavior of investigated compounds toward the binding sites of target proteins as well as to investigate fundamental biochemical processes.

2. Materials and Methods

2.1. Chemicals

The chemicals CuCl2·2H2O, CdCl2·H2O, VO(SO4)·5H2O, DMSO, KCl, 3-(benzo[d]thiazol-2-ylamino)-3-oxopropanoyl chloride, hydrazine hydrate, 2-acetyl pyridine, glacial acetic acid, and absolute EtOH were used without any treatment and were provided by BDH, Sigma-Aldrich, and Merck companies.

2.2. Synthesis of Ligand o-H2BMP and Complexes

The synthesis of N-(benzothiazol-2-yl)-3-oxo-3-(2-(1-(pyridin-2-yl)ethylidene) hydrazinyl)-propenamide ligand (o-H2BMP) was accomplished by heating under reflux a mixture of N-(benzo[d]thiazol-2-yl)-3-hydrazinyl-3-oxopropanamide (0.01 m; 2.503 gm) with 1-(pyridin-2-yl)ethan-1-one (0.01 m; 1.21 gm) on addition drops of glacial acetic acid for 3 h.11,12 After completion of the reaction, a newly formed product of yellowish-white precipitate was obtained, which was confirmed by thin layer chromatography (TLC) and separated by filtration, followed by washing with absolute EtOH. In addition, Cu(II), Cd(II), and VO(II) complexes were formed according to Scheme 1.

Scheme 1. Synthesis of o-H2BMP Ligand and Its Metal Complexes.

Scheme 1

2.3. Analysis of o-H2BMP Ligand and Its Complexes

The percentages of C, H, and N for the (o-H2BMP) ligand and its metal complexes were performed in Microanalytical Unit, Ain Shams University, Egypt. Also, the metal cations’ contents were determined by complexometric titrations. The content of sulfate in the VO(II) complex was detected by gravimetric analysis in the form of barium sulfate. The instruments used to elucidate the structures of the isolated compounds were Fourier-transform infrared (FT-IR) spectroscopy, UV–visible spectroscopy, 1H/13C NMR spectroscopy, LC coupled to mass spectrometry (LC–MS), electron spin resonance (ESR) spectroscopy, powder X-ray diffraction (PXRD) analysis, thermogravimetric analysis/differential thermal analysis (TGA/DTA), and magnetic susceptibility balance measurements. CV was performed by a DY 2100 potentiostat with three electrodes (glassy carbon as a working electrode, platinum as an auxiliary electrode, and Ag/AgCl as a reference electrode) that were immersed in a cell containing 0.1 M of KCl solution as a supporting electrolyte13,14 as shown in Scheme 1S.

2.4. Gaussian Calculations

The density-functional theory (DFT) calculations were performed using Gaussian 09 software on applying the B3LYB functional and the LANL2DZ/6-31+G(d,p) mixed basis set to realize the geometric optimized structures of the isolated solid compounds except for the ligand, which is obtained by using the 6-31+G(d,p) basis set.1518

2.5. Antioxidant Activity (DPPH, ABTS)

The DPPH (1,1-diphenyl-2-picrylhydrazyl) and ABTS [2,2-azinobis-(3-ethylbenzothiazoline-6-sulfonic acid)] assays are widely used in antioxidant studies because these methods are simple and very sensitive colorimetric techniques that rely on the production of colorful mono-cationic free radicals within a strong absorption (deep purple at λ 517 nm for DPPH+• and green-blue at λ 734 nm for ABTS+•). A significant decrease in the measured absorbance of these radicals, which act as hydrogen acceptors, is observed when hydrogen donor antioxidants are present. Thus, the inhibition percentage can be calculated using l-ascorbic acid as a reference standard antioxidant,1921 as shown in Scheme 2S.

2.6. Microbial Study

The antimicrobial activity was done by disc diffusion technique21,22 as shown in Scheme 3S. The microbial strains tested are Staphylococcus aureus, Bacillus subtilis, Escherichia coli, Salmonella typhi, and Candida albicans. The percentage of antimicrobial activity and anti-fungal activity for the compounds was evaluated by comparison with gentamycin and clotrimazole standards.23

2.7. MTT Assay

The antitumor efficiency of the free ligand and its metal complexes were examined in vitro against breast cancer cells (MDA-MB-231 and MCF-7). The examined compounds were dissolved in DMSO by 10 mM stock and stored at −20 °C, and cisplatin was used as a positive control. Cells were cultured using Dulbecco’s modified Eagle’s medium in a 96-walled plate and completed with 10% fetal bovine serum with (100 IU/mL) penicillin/streptomycin (100 μg/mL) (Lonza, 17-602E) at 95% air condition with 5% of CO2 in 37 °C. The initial screening and cell viability were assessed by MTT assay,2428 in which the absorbance was measured at λ 570 nm by a Biotek plate reader (Gen5), as shown in Scheme 4S.

2.8. Molecular Docking

MOE (molecular operating environment 2015) software was used to carry out molecular docking calculations of the prepared compounds with protein receptors 6CBZ, 1FDW, 2WTT, 4GL7, 1A53, and 1X7R, which are expressed more during breast cancer. All proteins were obtained as PDB files from the Protein Data Bank. Water molecules connected to these proteins were excluded to avoid interfering with the docking study. The structures were drawn by using ChemDraw 15 software. These compounds were then docked into the active sites of selected proteins individually.29,30

3. Results and Discussion

3.1. Elemental Analyses and Physical Properties

The elemental analysis results with some physical properties are presented in Table 1 in which the calculated and experimental percentages strongly coincided with each other, confirming the proposed chemical formula. The o-H2BMP ligand reacted with the VO(II), Cu(II), and Cd(II) ions to form five- or six-membered ring chelates. According to the outcomes data, the isolated solid complexes have a stoichiometric ratio of 1 M/1 L). The measured molar conductance value for VO(II) complex (10–3 M) at room temperature in DMSO is 18 μS cm2 mol–1. Therefore, it is concluded that the SO42– ion is bonded to the VO(II) ion, indicating its coordinating nature.

Table 1. Physical Properties and Elemental Analysis of o-H2BMP Ligand and Its Complexes.

      (a) physical properties
(b) elemental analysis: found (calcd)
no. compound (chemical formula) m. wt mp color % C % N % H % S % M SO4
1 o-H2BMP (C17H15N5O2S) 353.40 214 yellowish white 57.33 (57.78) 19.83 (19.82) 4.29 (4.28) 8.76 (9.07)
2 [VO(o-H2BMP)(SO4)]·4H2O (C17H23N5O11S2V) 588.46 >300 olive 35.20 (34.70) 11.69 (11.90) 4.31 (3.94) 11.04 (10.90) 8.45 (8.66) 15.8 (16.31)
3 [Cu(o-BMP)(H2O)2]·2H2O (C17H21N5O6SCu) 486.99 >300 brown 42.61 (41.93) 14.85 (14.38) 4.09 (4.35) 6.05 (6.58) 11.31 (13.05)
4 [Cd(o-BMP (H2O)2]·2H2O(C17H21N5O6SCd) 535.86 >300 yellow 38.78 (38.10) 13.65 (13.07) 4.22 (3.95) 6.21 (5.98) 20.83 (20.98)

3.2. FT-IR Spectra

The FT-IR spectrum of the hydrazone ligand (o-H2BMP) revealed various important characteristic bands at 1652, 1696, 1628, 1605, 3216, 3180, and 689 cm–1 that belong to ν(C=O)1, ν(C=O)2, ν(C=N)1 and ν(C=N)2, ν(NH)1, ν(NH)2 and breathing mode of vibration for pyridine ring ν(C=N)py, correspondingly.

The spectrum of the [VO(o-H2BMP)(SO4)]·4H2O complex showed that the (o-H2BMP) coordinated as a neutral bidentate ligand via ν(C=O)1 and ν(C=O)2 groups according to the following variations: (i) the change in the intensity of the two carbonyl bands (C=O)1 and (C=O)2 as well as the shift of these two groups to lower wavenumbers of 1646 and 1678 cm–1, respectively; (ii) the appearance of a new ν(SO42–) at 1041 and 1164 cm–1, predicting the bidentate nature;31,32 and (iii) the presence of new ν(V=O) at 980 cm–1,33,34 and ν(V–O) at 513 cm–1.

While the IR spectrum of both [Cu(o-BMP)(H2O)2]·2H2O and [Cd(o-BMP)(H2O)2]·2H2O complexes deduced the binegative tetradentate chelation via two enolic oxygens (=C–O-)1, (=C–O-)2, and two azomethine nitrogens (C=N)1, (C=N)2 based on the following observations: (i) disappearance of ν(C=O)1, ν(C=O)2, ν(NH)1 and ν(NH)2 with a simultaneous appearance of new enolic ν(C–O)1, ν(C–O)235 and azomethine ν(C=N)*1, ν(C=N)*2 groups at (1271, 1268) cm–1, (1103, 1108) cm–1, (1634, 1638) cm–1, and (1629, 1630) cm–1, correspondingly; (ii) shift of ν(C=N)1, ν(C=N)2 to lower wavenumber at (1605, 1600) cm–1 and (1582, 1589) cm–1, respectively; and (iii) existence of new bands for ν(M–O) at (575, 570) cm–1 and ν(M–N) at (446, 443) cm–1, correspondingly. All the most characteristic FT-IR bands of the ligand and its complexes are listed in the Table 2.

Table 2. Most Characteristic FT-IR Bands of Ligand 1 (o-H2BMP) and Its Complexes.

  ν(C=O)thiazole1 ν(C=O)azo2 ν(C=N)thiazole1 ν(C=N)azo2 ν(NH)thiazole1 ν(NH)azo2 ν(C=N)thiazole*1 ν(C=N)azo*2 ν(=C–O)thiazole1 ν(=C–O)azo2
1 1652 1696 1628 1605 3216 3180
2 1646 1678 1627 1604 3223 3178
3 1605 1582 1634 1629 1271 1103
4 1600 1589 1638 1630 1268 1108

3.2.1. Correlation Analysis

Correlation analysis has been established between the experimental and theoretical absorption bands deduced from the infrared spectroscopic investigations, where the theoretical absorption bands were computed from DFT frequency data and the experimental absorption bands were detected from the measured IR spectra of solid compounds. The converge diagrams in conjugation with the scatter plots were used to clarify the statistical correlation.36 Based on the results in Figures 1 and 2 and Table 1S, it was evident that all compounds had a strong positive linear correlation, with R2 ranging between 0.9728 and 0.9997.

Figure 1.

Figure 1

Experimental and theoretical IR spectra of (1) o-H2BMP, (2) [VO(o-H2BMP)(SO4)]·4H2O, (3) [Cu(o-BMP)(H2O)2]·2H2O, and (4) [Cd(o-BMP)(H2O)2]·2H2O.

Figure 2.

Figure 2

Correlation plots of experimental and theoretical data for the most characteristic IR bands of (1) o-H2BMP, (2) [VO(o-H2BMP)(SO4)]·4H2O, (3) [Cu(o-BMP)(H2O)2]·2H2O, and (4) [Cd(o-BMP)(H2O)2]·2H2O.

3.3. 1H NMR and 13C NMR

The 1H NMR spectrum of o-H2BMP as illustrated in Figure 3 showed amide-iminol tautomerism (Scheme 5S) according to the following observations: (i) imide proton (NH)1 appeared at chemical shift (δ) 12.63 ppm for both tautomers; (ii) imide proton (NH)2 for iminol tautomer appeared at (δ) 10.85 ppm; (iii) enolic proton (O–H) appeared at (δ) 11.04 ppm which predicted the enolization of (o-H2BMP) ligand; (iv) aromatic protons existed in the range of (δ) 7.26–8.54 ppm in both tautomeric structures; (v) two singlet signals for methylene and methyl protons (CH2), (CH3) at (δ) 3.98, δ 2.43 ppm for amide moiety, and (δ) 3.78, 2.52 ppm for iminol moiety, respectively.

Figure 3.

Figure 3

1H NMR spectrum of o-H2BMP.

The 13C NMR spectrum of o-H2BMP as demonstrated in Figure 4 showed the following remarks: (i) aromatic carbons appeared in the range of δ 121.00–132.65 ppm in both tautomeric structures; (ii) cyclic carbons existed in the range of δ 21.73–29.26 ppm for both two tautomeric structures; (iii) in amide tautomer, sharp peaks at (δ) 167.46, 165.58, 158.39, 140.26, and 43.37 ppm are assigned to (C=N)1, (C=O)2, (C=O)1, (C=N)2, and (CH2), respectively; and (iv) in imidic acid form, sharp peaks at δ 169.65, 167.20, 158.11, 149.07, and 42.24 ppm which correspond to (C=N)1, (C=N)*, (C=O)1, (C=N)2, and (CH2), respectively.

Figure 4.

Figure 4

13C NMR spectrum of o-H2BMP.

In a comparison of the 1H NMR spectrum of the free ligand with [Cd(o-BMP)(H2O)2]·2H2O as in Figure 1S, it was noticed that the absence of any protons possessed imide or enolic groups for both tautomers, which predicted the hydrazone reacted as a bi-negative ligand via two enolic groups as a result of the enolization of two carbonyls. It also showed sharp two-singlet peaks at δ 3.3–3.5 ppm for methylene group protons, δ 2.5–2.52 ppm for methyl protons, and multiple δ 7.24–8.62 ppm for aromatic benzene and pyridine ring protons.

3.4. UV–Visible and Magnetism

The qualitative UV–visible spectra were obtained in the DMSO solvent to obtain smooth absorption curves as shown in Figure 2S. Interaligand transition bands appearing at 35,971, 33,898, and 27,778 cm–1 are attributed to n → σ*, π → π*, and n → π* transitions, respectively. A structurally condensed conjugation of chromophores leads to the appearance of a yellowish-white color for the hydrazone ligand. The VO(II) complex displayed a band at 19,685 cm–1 assigned to 2B22E22). This spectrum, accompanied by the magnetic moment value (μeff = 1.79 BM), is attributed to the square pyramidal configuration.3739 Whereas, the electronic spectrum of the Cu(II) complex displayed two transition bands at 16,207 and 14,771 cm–1 assigned to 2B1g2Eg and 2B1g2A1g transitions, respectively, which indicated the existence of a distorted octahedral geometry around the Cu(II) ion.40 Also, the magnetic moment value of [Cu(o-BMP)(H2O)2]·2H2O is 2.13 BM.

3.5. EPR Studies

EPR spectra of VO(II) and Cu(II) complexes (Figure 5) were recorded using a Bruker EMX spectrometer operating at room temperature in the X-band (9.685 GHz) with a 100 kHz modulation frequency.

Figure 5.

Figure 5

EPR spectra of [Cu(o-BMP)(H2O)2]·2H2O and [VO(o-H2BMP)(SO4)]·4H2O complexes.

The EPR spectrum of the Cu(II) complex displayed a four-line hyperfine pattern with spin Hamiltonian parameters (I = 3/2, S = 1/2) that showed axial symmetry similarity for g-tensor parameters (g||, 2.26 > g, 2.09 > ge, 2.0023), indicating that Cu(II) has a dx2y2 ground-state with octahedral geometry.41 These g-tensor parameters were calculated from the EPR spectrum by applying eq 1(4244)

3.5. 1

where h is the Planck constant (6.626 × 10–34 J), ν is the frequency (9.865 MHz), μ the is applied magnetic field in mT, and B is the Bohr magneton (9.27 × 10–24 J/mT). Consequently, the G-factor for axial symmetry could be calculated from eq 2

3.5. 2

Essentially, G is a measure of the exchange interaction between Cu(II) centers in the solid state complex, which is negligible if G is greater than 4, while if G is less than 4, there is a considerable exchange interaction in the Cu(II) complex.45 In addition, the parallel and perpendicular hyperfine parameters (A||, 0.0152 and A, 0.0531) components were estimated from the spectrum. Furthermore, the F-factor was determined by applying eq 3

3.5. 3

Basically, the F-factor is used to quantify the degree of distortion of the Cu(II) complex. Where, if the F-factor value is between (105 and 135), the complex is square-planar, while the tetrahedrally distorted structure can have a much larger value (>135). The F-factor value for the Cu(II) complex was achieved (148.38), predicting the existence of distorted octahedral geometry.

The molecular orbital coefficients, covalent in-plane σ-bonding (α2), and covalent in-plane π-bonding (β2) were calculated using eqs 4 and 5

3.5. 4
3.5. 5

where E is the energy of electronic transition on the UV spectrum, and λ = −828 cm–1 for the free Cu(II) ion. The factor (α2 = 1) represents a full ionic character, whereas (α2 = 0.5) represents absolute covalent bonding, and the small values are negligible. As the β2 value decreases, the covalent in-plane π-bonding covalence increases.46 The Cu(II) complex has very strong in-plane π-bonding, based on the values of α2 = 0.75 and β2 = 0.75. These results are expected due to the ligand having orbitals that combine with the dxy orbital in the copper ion.

The EPR spectrum of the VO(II) complex gave a typical eight-line characteristic pattern of interaction with spin Hamiltonian parameters (S = 1/2 and I = 7/2) and one unpaired electron (3d1) in the VO(II) ion. The EPR parameters (g||, g, A||, A, G, F) calculated for the [VO(o-H2BMP)(SO4)]·4H2O complex are presented in Table 3, while the molecular orbital α2 and β2 coefficients were applied using eqs 6 and 7(47)

3.5. 6
3.5. 7

where λ for VO(II) ion = 170 cm–1, and the Fermi contact (κ), a measure of the d-orbital population for unpaired electron, was calculated from eqs 8 and 10

3.5. 8
3.5. 9
3.5. 10

where P is the constant of direct dipole–dipole interaction between the electron and the magnetic moment of the vanadium nucleus (0.0136 cm–1).48 The VO(II) complex gave (G = 1.5 < 4), which indicated the exchange interaction between VO(II) centers in the solid state, and the F-factor was 111.78, which is less than 135 indicating the square-pyramidal geometry around the VO(II) ion.

Table 3. ESR Parameters for Cu(II) and VO(II) Complexes.

no. g|| g A|| (cm–1) A (cm–1) G F α2 β2
2 1.93 1.95 0.0173 0.0086 1.5 111.78 0.23 2.45
3 2.26 2.09 0.0152 0.0531 2.9 148.38 0.75 0.75

The calculated (β2 = 2.45) value which deviated from the expected range between (0.93–1) indicated the increase of the covalency degree, and this is a measure for the bonding between the dxy orbital of VO(II) with the π-orbitals of the ligand. As a result, the α2 value (α2 = 0.23) showed a very strong in-plane π-bonding.49,50 This more covalent bonding resulted in the appearance of an overlapped hyperfine splitting within small variations between the g-values (g||, 1.93 < g, 1.95 < ge, 2.0023) and A-values (A||, 0.0173 and A, 0.0086).

3.6. LC–MS

The mass spectrum of the o-H2BMP ligand was obtained5153 by using the LC/MSD iQ instrument at a retention time of 2.958 min, as shown in Figure 6. The molecular ion peak [m/z] of the free ligand was observed at 354, which approximately corresponds to the exact molecular weight.

Figure 6.

Figure 6

Mass spectrum of o-H2BMP.

3.7. TGA/DTA

Thermal analysis was measured for the isolated complexes to determine the water molecules’ nature, as coordinated or crystalline Table 3S. TGA curves are illustrated in Figures 3S–5S, and thermal decomposition steps are clarified in Schemes 6S–8S.

The TGA curve of [VO(o-H2BMP)(SO4)]·4H2O validated five successive decomposition steps. An initial temperature range of 25–115 °C corresponds to the removal of four crystalline water molecules (weight loss: found 12.354%; calcd 12.245%). As a next step, the H2SO4 molecule was removed at 116–335 °C (weight loss: found 16.677%; calcd 16.665%). The third step occurred at 336–445 °C, which ascribed to the loss of C7H5NS moiety (weight loss: found 23.084%; calcd 22.972%). Fourth, a loss of C7H5N unit occurred between 446 and 535 °C (weight loss: found 17.075%; calcd 17.517%). The fifth step occurred at 536–800 °C, contributing to the loss of C3H3N3O (weight loss: found 16.646%; calcd 16.50%). VO2 is the final residue left behind after the complex decomposed entirely above 800 °C (remaining weight: found 14.136%; calcd 14.09%).

The TGA curve of [Cu(o-BMP)(H2O)2]·2H2O demonstrated four decomposition stages. In the first step, two crystal water molecules were removed between 25 and 116 °C (weight loss: found 7.125%; calcd 7.40%). Second, both two coordinating water molecules accompanied by the C5H5N moiety were lost at 117–309 °C (weight loss: found 23.258%; calcd 23.641%). The third step appeared at 310–419 °C and was attributed to C7H5NS loss (weight loss: found 27.652%; calcd 27.758%). In the last step, 420–800 °C was utilized for the loss of the C5H3N3O fragment (weight loss: found 25.267%; calcd 24.867%), departing CuO as the final residue (remaining weight: found 16.698%; calcd 16.335%).

The TGA curve of [Cd(o-BMP)(H2O)2]·2H2O revealed five decomposition steps. The first one was performed at 24–107 °C, which led to the removal of two crystal water molecules (weight loss: found 6.60%; calcd 6.724%). The next stage occurred at 108–333 °C and was attributed to the loss of two coordinating water molecules and the C5H5N moiety (weight loss: found 21.507%; calcd 21.485%). During the third step, the C7H5NS moiety was decomposed in the range 336–440 °C (weight loss: found 25.241%; calcd 25.230%). In step four, the remaining part of the organic skeleton, C5H3N3O, was fired at a temperature range of 494–800 °C (weight loss: found 22.569%; calcd 22.60%), leaving CdO as final residual ash (remaining weight: found 24.082%; calcd 23.963%).

3.7.1. Kinetic and Thermodynamic Studies

TGA coupled with DTA diagrams were used to estimate kinetic and thermodynamic parameters using Coats–Redfern and Horowitz–Metzger methods,5457 as shown in Figures 6S–11S. The calculated parameters (Ea, A, ΔH, ΔS, and ΔG) for the thermal degradation steps (Tables 4S–6S) can be concluded as follows: (i) the high activation energy (Ea) values for the thermal decomposition steps indicated the high stability of complexes;58 (ii) all decomposition steps were endothermic processes as a result of (+ΔH) enthalpy values; (iii) the (+ΔS) entropy values indicated the fast decomposition reactions while the (−ΔS) entropy values revealed the slow degradation stages;59 (iv) the (+ΔG) free energy values indicated that the free energy of the final residue is greater than that of the initial complex supporting the non-spontaneous processes.6062

3.8. PXRD Analysis

The PXRD analysis was done at room temperature via the Rigaku Ultima IV X-ray diffractometer with Cu anode radiation Kα (λ = 1.5418 Å) and 2θ measuring from 4 to 70°. The prepared compounds showed diffraction patterns in Figure 7 analogues to crystalline phases natures. These patterns can be used to calculate the crystalline size (S) at the most intense peak on the light of the Debye–Scherrer equation, S = 0.9λ/(β cos θ), where λ = 1.5406 Å, β is the fwhm of the peak, and θ is the Bragg diffraction angle. Furthermore, the d-spacing values can be obtained by using the Bragg equation nλ = 2d sin(θ), assuming (n = 1).63 The PXRD data indexing analysis was performed by QualX software to determine the lattice parameters of the isolated compounds,64 such as the crystal system, cell parameters, and cell volume, which are listed in Table 4. The indexing was established by comparing between 2θ values in the crystallography open database (COD)6568 with the experimental data for isolated compounds, as illustrated in Table 7S.

Figure 7.

Figure 7

PXRD patterns of (1) o-H2BMP, (2) [VO(o-H2BMP)(SO4)]·4H2O, (3) [Cu(o-BMP)(H2O)2]·2H2O, and (4) [Cd (o-BMP)(H2O)2]·2H2O.

Table 4. Crystallographic Data of o-H2BMP Ligand and Its Metal Complexes.

Compound 1 2 3 4
31.832 27.814 31.765 31.674
d-spacing (Å) 2.809 3.2037 2.8137 2.8215
fwhm (β) 0.100 0.100 0.113 0.100
crystal size (nm) 82.58 81.81 73.07 82.55
crystal system triclinic triclinic monoclinic monoclinic
cell parameters a = 8.3018 Å, b = 9.4214 Å, c = 16.2930 Å, α = 83.450°, β = 77.590°, γ = 85.730° a = 7.9043 Å, b = 10.2300 Å, c = 12.5970 Å, α = 102.320°, β = 102.610°, γ = 110.500° a = 8.2021 Å, b = 12.8425 Å, c = 14.6338 Å, α = 90°, β = 91.109°, γ = 90° a = 6.8284 Å, b = 10.6888 Å, c = 18.5505 Å, α = 90°, β = 94.305°, γ = 90°
cell volume 1234.81 Å3 883.04 Å3 1541.17 Å3 1350.13 Å3

The results revealed that the o-H2BMP ligand and the VO(II) complex are polycrystalline with a triclinic crystal system, while both Cu(II) and Cd(II) complexes are polycrystalline within a monoclinic crystal system. The calculated crystal sizes were 82.58, 81.81, 73.07, and 82.55 nm, for the o-H2BMP ligand, VO(II), Cu(II), and Cd(II) complexes, respectively. Whereas, the d-spacing result is 2.809, 3.2037, 2.8137, and 2.8215 nm, correspondingly.

3.9. Gaussian Studies

3.9.1. Molecular Modeling

DFT coupled with the [LANL2DZ/6-31G+(d,p)] mixed basis set6972 was employed to optimize geometrically the molecular structure of all isolated compounds within atom numbering as represented in Figure 8. The results displayed in Table 8Sa,b showed that (i) the bond lengths of the most characteristic functional groups for isolated complexes correlate with the IR band expansions, confirming their participation in the coordination process, and (ii) the bond angles of the [Cu(o-BMP)(H2O)2]·2H2O and [Cd(o-BMP)(H2O)2]·2H2O complexes correlate a distorted octahedral arrangement, while the [VO(o-H2BMP)(SO4)]·4H2O exists as a square-pyramidal geometry supporting the configurations concluded by UV–visible spectral analyses.73

Figure 8.

Figure 8

Molecular modeling of (a) o-H2BMP, (b) [VO(o-H2BMP)(SO4)]·4H2O, (c) [Cu(o-BMP)(H2O)2]·2H2O, and (d) [Cd(o-BMP)(H2O)2]·2H2O.

3.9.2. Frontier Molecular Orbitals and Molecular Chemical Parameters

As a complement to the conceptual density functional theory calculations, the dipole moment (D) and distinct quantum energetic values have been derived as represented in Table 5, such as the total electronic energies (ETotal) and the complex binding energy (Ebinding).74 Furthermore, the energies of frontier molecular orbitals (EHOMO and ELUMO) which are subjected to calculation (Egap) as well as several global and local descriptors of chemical reactivity like the electronegativity (χ), the chemical potential (μ), the hardness (η), the softness(σ), and the electrophilicity index (ω)75 are presented in eqs 1117.

3.9.2. 11
3.9.2. 12
3.9.2. 13
3.9.2. 14
3.9.2. 15
3.9.2. 16
3.9.2. 17
Table 5. Calculated Dipole Moment, Total Energy, Binding Energy, and Chemical Reactivity Parametersa.
compound 1 2 3 4
dipole moment (D) 6.0221 18.0239 2.3230 5.605
ETotal (Hartree) –1479.3152 –2325.0267 –1827.1159 –1679.0503
Ebinding (Hartree) –623.9644 –121.6967 –121.6791
EHOMO (eV) –0.2367 –0.2543 –0.2219 –0.2077
ELUMO (eV) –0.0698 –0.1052 –0.0639 –0.0736
Egap (eV) 0.1669 0.1491 0.1580 0.1341
χ (eV) 0.1533 0.1798 0.1429 0.1407
μ (eV) –0.1533 –0.1798 –0.1429 –0.1407
η (eV) 0.0835 0.0746 0.079 0.0671
σ (eV) 11.9832 13.4138 12.6582 14.9142
ω (eV) 0.1407 0.2167 0.1292 0.1475
a

Note: total energies of metals (V(IV), Cu(II), and Cd(II)) are −221.7474, −226.1040, and −78.0560 Hartree, respectively.

The frontier molecular orbital distribution of the compounds is depicted in Figure 9. The low values of the HOMO and LUMO energy gaps showed that the studied compounds have high chemical reactivity, biological activity, and polarizability. Chemical potential is used to determine the stability of isolated compounds. It can also predict their tendency to chemically react to form new substances, transform into new physical states, or migrate from one spatial location to another.76,77 Both chemical potential energy and electronegativity can also be thought of as depending on chemical composition, where the VO(II) complex possesses the most negative chemical potential magnitude (μ = −0.1798) and the largest electronegativity value (x = 0.1798) among the compounds, enhancing the lowest stability or the most reactivity.

Figure 9.

Figure 9

HOMO and LUMO of (a) o-H2BMP, (b) [VO(o-H2BMP)(SO4)]·4H2O, (c) [Cu(o-BMP)(H2O)2]·2H2O, and (d) [Cd(o-BMP)(H2O)2]·2H2O.

The free ligand is the hardest (lowest soft) base compound with values (η = 0.0835, σ = 11.9832) as a result of bearing the greatest number of highly electronegative donor atoms (O or N centers). This refers to the steadiness of prepared complexes using such hydrazone ligands, where generally the stable complexes are formed between hard acids and hard bases or between soft acids and soft bases. Based on the minimum electrophilicity principle stated as “Electrophilicity will be a minimum (maximum) when both chemical potential and hardness are maxima (minima)”, the order of the compounds according to their electrophilicity index (ω) is VO(II) complex > Cd(II) complex > ligand > Cu(II) complex.

3.9.3. Molecular Electrostatic Potential

The molecular electrostatic potential (MEP) is used to predict the relative reactivity positions in a species for nucleophilic and electrophilic attack and to study the hydrogen bonding interactions, in addition to evaluating some biological significance. The MEPs of the optimized geometry of the o-H2BMP ligand and its complexes were obtained based on their DFT calculation.78

The MEP mapping for an isolated ligand and its complexes can be categorized into three zones according to their colors, as shown in Figure 10: (i) the red region matches the high electronic density zone, which enhances the electrophilic attack interactions; (ii) the green region indicates the neutral electrostatic potential zone; and (iii) the blue region refers to the low electronic density zone, which reflects the nucleophilic attack demeanor.79

Figure 10.

Figure 10

MEP surface maps of (a) o-H2BMP, (b) [VO(o-H2BMP)(SO4)]·4H2O, (c) [Cu(o-BMP)(H2O)2]·2H2O, and (d) [Cd(o-BMP)(H2O)2]·2H2O.

As can be seen for compound 1, the negative electrostatic potential regions are localized over the electronegative atoms (O-carbonyl groups, S-thiazole ring, N-thiazole ring, N-Schiff base moiety, and N-pyridine ring), while the positive electrostatic potential regions are mainly concentered over hydrogen atoms of NH-imine groups. Also, compound 2 has negative electrostatic potential regions focused over the sulfate group and oxovanadium species, causing a strong effect in different interactions. On the other hand, both compounds 3 and 4 have weak negative electrostatic potential regions limited to pyridine nitrogen within intense positive electrostatic potential regions over hydrogen atoms of water molecules.

3.10. Cyclic Voltammetry

3.10.1. CV of Cu(II)

The electrochemical behavior of 3.13 × 10–3 M of Cu(II) cation was studied via CV using a glassy carbon working electrode immersed in 0.1 M of supporting electrolyte KCl dissolved in a 50% DMSO–50% water mixed solvent. The scanning was applied at a potential window (1.2 to −1.2 V) within a 0.1 V/s scan rate at 291.15 K, as shown in Figure 11. The voltammogram showed two peaks in the reduction sweeping and two in the oxidation side, corresponding to the conversions between [Cu(II)/Cu(I)], [Cu(I)/Cu(0)] species. The mechanisms of both reduction and oxidation reactions can be described as follows

3.10.1. 18
3.10.1. 19
3.10.1. 20
3.10.1. 21
Figure 11.

Figure 11

CV of 3.13 × 10–3 M Cu(II) in 0.1 M KCl.

3.10.2. CV of the Cu (II)/o-H2BMP System

The electrochemical behavior for the complexation reaction between o-H2BMP and copper spices was studied with the same conditions and settings as shown in Figure 12. The voltammogram revealed the interaction between copper and ligand species according to the potential shifts to new values and the decreases in both cathodic and anodic peak currents.

Figure 12.

Figure 12

CV of 3.13 × 10–3 M Cu(II) in 0.1 M KCl in the presence of o-H2BMP.

Also, the stability constants (βj), a measure of the strength of the interaction between the metal ion and ligand to form the complex, can be evaluated by Lingane’s eq 22(80)

3.10.2. 22

where both EC° and EM° are the formal peak potentials corresponding to the complex and free metal ions, n is the number of electrons, F is the Faraday constant (96,485.33 Columb/mol), R is a gas constant (8.314 J mol–1 K–1), T is the absolute temperature, j is the molar ratio of the stoichiometric complex, and [L] is the o-H2BMP ligand concentration. The formal potential (E°) was determined as follows

3.10.2. 23

By plotting −log [L] versus ΔE°81 as shown in Figure 13 on the aid of (Epc, Epa, ipa, ipc, E°, j) data present in Table 6, the value of j calculated from the slope was achieved (0.9497 ≈ 1), as explained in eq 24

3.10.2. 24
Figure 13.

Figure 13

Plot of −log[L] vs ΔE°.

Table 6. CV Data of CuCl2 (Cu2+/Cu+) in the Absence and Presence of (o-H2BMP) Ligand at 291.15 K.
M ×10–3 CuCl2 M ×10–3o-H2BMP Epa (V) Epc (V) ipa (μA) ipc (μA) E° (V) ΔE° (V) j
3.13   0.357 0.173 46.537 21.601 0.265   0
3.03 0.3030 0.357 –0.026 41.780 17.748 0.1655 0.100 0.1
2.94 0.5882 0.353 –0.034 38.257 18.545 0.1595 0.106 0.2
2.86 0.8571 0.354 –0.037 35.294 18.829 0.1585 0.107 0.3
2.78 1.111 0.358 –0.042 33.069 18.572 0.158 0.107 0.4
2.70 1.351 0.362 –0.056 30.712 18.499 0.153 0.112 0.5
2.63 1.579 0.365 –0.047 30.095 17.945 0.159 0.106 0.6
2.56 1.795 0.366 –0.05 26.730 17.170 0.158 0.107 0.7
2.50 2.000 0.37 –0.048 24.689 17.477 0.161 0.104 0.8
2.44 2.195 0.371 –0.05 24.474 17.715 0.1605 0.105 0.9
2.381 2.381 0.365 –0.046 22.385 17.271 0.1595 0.106 1
2.33 2.558 0.374 –0.033 20.052 15.228 0.1705 0.095 1.1
2.27 2.727 0.378 –0.042 19.078 15.814 0.168 0.097 1.2
2.17 3.043 0.397 –0.029 12.721 15.312 0.184 0.081 1.4
2.08 3.333 0.402 –0.014 10.285 12.994 0.194 0.071 1.6
2.00 3.600 0.404 –0.016 9.1415 11.326 0.194 0.071 1.8
1.92 3.846 0.41 –0.006 9.9837 10.948 0.2047 0.060 2
1.85 4.074 0.41 –0.007 9.5524 7.9592 0.2015 0.064 2.2
1.79 4.286 0.405 –0.016 7.2598 9.9597 0.1945 0.071 2.4
1.72 4.483 0.407 –0.011 6.1943 8.6621 0.198 0.067 2.6
1.67 4.667 0.405 0.006 4.9333 8.9457 0.2055 0.060 2.8
1.61 4.839 0.402 –0.004 3.8367E 8.3597 0.199 0.066 3

Further, on plotting (j) values versus (ipa) anode peak current in the (Cu+ to Cu2+) peak, the break on the curve (j = 1.04 ≈ 1) predicts (1:1) stoichiometry as shown in Figure 14, which coincided with the results from the elemental analysis calculations.

Figure 14.

Figure 14

Plot of j vs ipa.

The stability constant was achieved (βj = 28.18 × 103, log βj = 4.45), and the Gibbs free energy was obtained (ΔG = −24.80 kJ/mol) based on the eq 25,82 predicting a favorable (1:1) stable and spontaneous interaction between Cu2+ and o-H2BMP species in the solution environment.

3.10.2. 25

3.11. Biological Studies

3.11.1. Antioxidant Studies

The antioxidant capacity of the (o-H2BMP) ligand and its complexes were measured using DPPH and ABTS radical scavenging methods, and usually, l-ascorbic acid (Vitamin C) is utilized as a reference.83 The percentage of inhibition was obtained from eq 26

3.11.1. 26

The concentrations 370.37 and 250 μM were used to obtain the optimum inhibition effect in both the DPPH and ABTS methods, respectively. The free ligand and complexes displayed DPPH antioxidant activity in the sequence: (5.10%; Cd(II) complex) < (23.40%; o-H2BMP) < (43.90%; Cu(II) complex) < (87.31%; VO(II) complex), while displayed ABTS antioxidant activity in the sequence: (33.66%; Cu(II) complex) < (92.98%; Cd(II) complex) < (94.93%; o-H2BMP) < (95.09%; VO(II) complex) as illustrated in Figure 15. The IC50 values were calculated for the samples that gave (I %) of more than (50%) by using the serial dilutions (370, 185, 92.5, 46.25, and 23.125 μM for DPPH) and (250, 125, 62.5, 31.25, and 15.625 μM for ABTS) and then plotting the values of the ordered concentrations versus inhibition percent via the GraphPad Prism 9 program, as shown in Figures 12S and 13S and listed in Tables 7, 9S, and 10S.

Figure 15.

Figure 15

Inhibition effect (I %) values of the DPPH and ABTS methods.

Table 7. Results of (I %) and IC50 by DPPH and ABTS Methods.
  DPPH method
ABTS method
no. I % IC50 (μM) I % IC50 (μM)
1 23.4 94.9 23.09
2 87.3 21.8 95.1 6.3
3 43.9 33.7
4 5.1 93.0 58.6
Vit C 94.3 27.8 100 3.6

3.11.2. Microbial Studies

The antimicrobial activity index of (o-H2BMP) ligand and metal complexes compared to gentamycin and clotrimazole standards can be evaluated84 by using eq 27

3.11.2. 27

The screening, as summarized in Table 8, reflects the following results: (i) the antimicrobial activity for the ligand revealed remarkable values in the range between (65.0–71.4%) compared to the standards against all microbial strains except S. typhi (ii) copper complex showed moderate values between (39.4–60.0%) against all strains excluding S. typhi (iii) cadmium complex gave values close to 40% with Gram-positive bacteria and fungus whereas inactive toward Gram-negative bacteria (iv) oxovanadate complex inhibited B. subtilis with 54.2% and S. typhi with 56.2% only.

Table 8. Results of the Inhibitory Zone and the % Antimicrobial Activity for o-H2BMP and Its Metal Complexes.
  bacterial strains
 
  Gram-positive bacteria
Gram-negative bacteria
fungus
no. S. aureus B. subtilis E. coli S. typhi Candida
1 15 mm, (71.4%) 17 mm, (70.8%) 13 mm, (65%) 23 mm, (69.7%)
2 13 mm, (54.2%) 9 mm, (56.25%)
3 10 mm, (47.6%) 10 mm, (41.7%) 12 mm, (60%) 13 mm, (39.4%)
4 10 mm, (47.6%) 11 mm, (45.8%) 14 mm, (42.4%)
gentamicin 21 mm, (100%) 24 mm, (100%) 20 mm, (100%) 16 mm, (100%)
clotrimazole 33 mm, (100%)

3.11.3. Cell Viability

The synthesized ligand, metal complexes, and the corresponding free metal salts were assessed via the MTT assay against MDA-MB-231 and MCF-7 breast cancer cells on the use of cis-platin as a positive control. This method depends on the conversion of the yellow tetrazolium bromide (MTT) by mitochondrial succinate dehydrogenase in viable cells to a purple formazan derivative, which can be measured spectrophotometrically, and the percent of cell viability was calculated using eq 28(8587)

3.11.3. 28

In order to determine the IC50, the cells were incubated with serial dilutions (50, 25, 12.5, 6.25, 3.125, and 1.56 μM) of the most active compounds which gave cell viability less than 50%, and plotted the ordered concentrations of each compound (conc.) against the percentage of cell viability (% cell viability) as described in Figures 16, 14S, and 15S and listed in Tables 9, 11S, and 12S.

Figure 16.

Figure 16

Percentage cell viability of MDA-MB-231 and MCF-7 cell lines.

Table 9. Results of % Cell Viability and IC50 of MDA-MB-231 and MCF-7 Cell Lines.
  MDA-MB-231
MCF-7
no. % cell viability IC50 (μM) % cell viability IC50 (μM)
o-H2BMP 56.3 85.2
[VO(o-H2BMP)(SO4)]·4H2O 0 4.5 10.9 6.8
[Cu(o-BMP)(H2O)2]·2H2O 86.6 71.7
[Cd(o-BMP)(H2O)2]·2H2O 0 10.5 0 7.2
VOSO4 0.6 4.3 1.2 5.9
CuCl2 8.1 10.9 3.4 7.1
CdCl2 5.7 8.1 0.1 2.5
cisplatin 0 7.2 7.5 31.4

According to the results, which reflect the ability to reduce the growth of (MDA-MB-231) and (MCF-7) tumor cells, the tested compounds can be discussed as follows: (i) low toxic compounds within high cell viability percentages (>50%) appear on both the free ligand (56.4% against MDA-MB-231 and 85.2% against MCF-7) and the Cu(II) complex (86.6% against MDA-MB-231 and 71.1% against MCF-7), (ii) all metal salts VOSO4, CuCl2, and CdCl2 can be considered as high toxic compounds accompanied by low cell viability values against both breast cancer cell lines, (iii) the most toxic compounds with no substantial cell viability is Cd(II) complex predominantly, (iv) VO(II) complex is the most compatible compound with cis-platin where it gave a noticeable cell viable percent (10.9%) against (MCF-7) tumor cells without any viability against (MDA-MB-231) cells. Further, the largest IC50 values were found for the Cd(II) complex and CuCl2, which were highly effective against MDA-MB-231 (10.5 and 10.9 μM) and MCF-7 (7.3 and 7.2 μM). However, against MDA-MB-231, the smallest IC50 values were achieved with the VO(II) complex (4.5 μM) and VOSO4 (4.4 μM), while against MCF-7, the smallest IC50 values were achieved with CdCl2 (2.5 μM) and VOSO4 (5.9 μM).

As a consequence of the above remarks, the metal complexes and their corresponding metal salts showed better cytotoxic potency and lower cell viability than the free ligand and comparable potency to that of cis-platin. Specifically, one of the VO(II) complexes from the series presented great cytotoxic activity against all cancer cell lines that were tested in this study, with IC50 values in the range of 4.5–6.8 μM.

3.12. Molecular Docking

The main factor responsible for breast cancer development is estrogen. In many breast cancers, alpha estrogen receptor (ERα) activation increases cancer proliferation, while beta estrogen receptor (ERβ) activation inhibits it.88,89

In this research, the molecular docking of the novel (o-H2BMP) ligand and its complexes as human ERα inhibitors within protein or enzyme receptors (6CBZ, 1FDW, 2WTT, 4GL7, 1A53, and 1X7R) were studied. The results of the binding energy scores (S), bond energy, total energy, types of bonds, types of interactions, and root-mean-square deviation (rmsd) collected in Table 10 showed a high degree of selectivity and potency toward binding to the active sites of the proteins being studied. Molecular docking studies identify various types of interactions, such as hydrophobic, electrostatic, and H-bonding.90,91

Table 10. Most Efficient Docking Results of the (o-H2BMP) Ligand and Its Complexes.

no. enzyme (PDB code) S rmsd no. of bonds total energy (kcal/mol)
1 1A53 –7.7177 1.399 2 –10.9
1 2WTT –5.7477 1.8404 4 –9.2
1 6CBZ –5.3872 1.181 1 –8.8
2 4GL7 –7.8116 1.2756 8 –27
2 6CBZ –6.0583 1.7843 3 –16.2
2 2WTT –5.6126 1.4855 5 –12
2 1A53 –6.1172 2.6850 7 –12.1
2 1X7R –6.811 1.6811 6 –10.3
3 1A53 –6.4602 2.684 9 –29.6
3 2WTT –5.3431 2.6696 2 –11.1
3 1X7R –4.6324 0.9098 4 –10.7
4 1A53 –5.7249 1.3188 7 –27.7
4 1X7R –3.7207 1.4272 5 –16.5
4 1FDW –5.3399 2.4339 3 –12.5

The most efficient docking results represented in Table 10 revealed the following observations: (i) compound 1 predicted the highest total energy value (−10.9 kcal/mol) toward the protein receptors pocket (1A53) where this free ligand interacted with (LYS53) and (ARG182) amino acids via ligand active centers (O18 and O19) producing two hydrogen bonds as shown in Figures 17 and 18, also (ii) compound 2 predicted the largest total energy value (−27 kcal/mol) where this complex interacted through seven hydrogen bonds and one ionic bond with six amino acids (MET303, GLY436, ARG115, ARG435, TRP141, and ARG145) as illustrated in Figures 19 and 20, as well (iii) compound 3 predicted the best total energy value (−29.6 kcal/mol) toward the protein receptors pocket (1A53) where this complex interacted by three conventional hydrogen bonds, one ionic, one π-cation, and four π-H interactions, with six amino acids (SER211, LYS53, ARG182, ASN180, GLY212, and SER234) as clarified in Figures 21 and 22, finally (iv) compound 4 predicted the highest total energy value (−27.7 kcal/mol) toward the protein receptors pocket (1A53) where this complex interacted by three conventional hydrogen bonds, one ionic, one π-cation, and two π-H interactions, with four amino acids (SER211, LYS53, ARG182, and SER234) as represented in Figures 23 and 24. More details about the molecular docking study are mentioned in Supporting Information, Tables 13S–18S and Figures 16S–27S.

Figure 17.

Figure 17

2D interaction between (o-H2BMP) and 1A53.

Figure 18.

Figure 18

3D interaction between (o-H2BMP) and 1A53.

Figure 19.

Figure 19

2D interaction between [VO(o-H2BMP)(SO4)]·4H2O and 4GL7.

Figure 20.

Figure 20

3D interaction between [VO(o-H2BMP)(SO4)]·4H2O and 4GL7.

Figure 21.

Figure 21

2D interaction between [Cu(o-BMP)(H2O)2]·2H2O and 1A53.

Figure 22.

Figure 22

3D interaction between [Cu(o-BMP)(H2O)2]·2H2O and 1A53.

Figure 23.

Figure 23

2D interaction between [Cd(o-BMP)(H2O)2]·2H2O and 1A53.

Figure 24.

Figure 24

3D interaction between [Cd(o-BMP)(H2O)2]·2H2O and 1A53.

According to the docking results, it can be concluded that the complexes of [VO(o-H2BMP)(SO4)]·4H2O and [Cd(o-BMP)(H2O)2]·2H2O are efficient as anticancer, which are so closely matched with the experimental cytotoxicity screening on both cell lines (MDA-MB-231) and (MCF-7). The MEPs give us useful information about the potent sites concerning the interactions with the residues present in amino acids of protein receptors via hydrogen bonds. Where the H-acceptor interaction in molecular docking coincides with the red regions in MEP surface maps enhancing nucleophilic attack, whereas the H-donor interaction corresponds to the blue regions in MEP surface maps enhancing electrophilic attack.

3.13. Structure–Activity Relationship Studies

Many parameters like the dipole moment, EHOMO, ELUMO, and Egap, which depend on the theoretical data of DFT calculations, can be linked to the antimicrobial and antioxidant activity of compounds under study through the structure–activity relationship (SAR)92 as dipole moments are very useful in determining penetration rates through cell membranes and excretion rates. The studies of SAR predicted an inverse relationship between the dipole moment and the activity of compounds against the examined microorganism. With a reduced dipole moment, the compound becomes more lipophilic, which allows better penetration through the lipid layer of microorganisms, thus damaging them quickly. Based on the outcome data, the VO(II) complex exhibited a higher dipole moment than other complexes and caused the lowest antimicrobial activity. Whereas the free ligand, which contains donor atoms such as oxygen, nitrogen, and sulfur atoms that can donate electrons to biological receptors, has a low dipole moment, so the free ligand is more active as the bactericidal and fungicidal agent. In general, the lower results as antimicrobial agents are ascribed to their inability to diffuse through the microorganism lipid layer due to the dipole moment increasing.93,94

The EHOMO and ELUMO are closely concerned with the free radical scavenging activity of the antioxidant agents. In concept, nucleophiles and electrophiles share radical scavenging activities endorsed by the relative energy influence of the HOMO/LUMO orbitals, where the substance has a low Egap, loses electrons more easily, and thus, it can participate in reactions. All investigated compounds that have low Egap values have been classified as good electron-releasing species with potent antioxidant activity.95,96 Particularly, the VO(II) complex showed the greatest antioxidant efficiency toward both DPPH (I = 87.3%, IC50 = 21.8 μM) and ABTS (I = 95.1%, IC50 = 6.3 μM).

3.14. Molecular Property and Drug-Likeness Prediction

The hydrazone (o-H2BMP) ligand and its complexes were examined to check their compatibility with the Lipinski rule of five,9799 according to which an orally active drug has no more than one violation of the following criteria: hydrogen bond donors (HBD: the total number of nitrogen–hydrogen and oxygen–hydrogen bonds) ≤ 5, hydrogen bond acceptors (HBA: all nitrogen or oxygen atoms) ≤ 10, an octanol–water partition coefficient (log P) ≤ 5, and molecular weight (m. wt) ≤ 500. Molecular weight is very influential for drugs that interact with drug receptors/DNA. As the molecular weight increases, the compounds become denser. As well, log P values play a major role in promoting inactive membrane division and activating permeability, in the opposite direction of their influence on solubility. Finally, polar surface atoms (PSA) such as nitrogen, oxygen, and sulfur are significant parameters to evaluate drug transport properties. Based on MolSoft software, a model for estimating the drug-likeness score of the investigated compounds and their molecular properties was developed and summarized in Table 11. The values of HBA and HBD for most compounds obey the applied rule, so the prepared compounds could implement through the cellular membrane and acting as drugs.

Table 11. Calculated Parameters for Drug-Likeness Based on the Lipinski Rulea.

no. m. wt log P HBA HBD PSA m. volume
1 353.41 1.76 7 2 96.34 301.24
2 516.41 –2.4 12 2 154.22 383.6
3 450.97 –5.9 9 4 109.74 341.03
4 499.83 –5.9 9 4 109.74 341.03
a

Notes: HBA and HBD are hydrogen bond acceptors and hydrogen bond donors, respectively. The calculated m.wt of compounds does not include crystallized water molecules.

4. Conclusions

A series of Cu(II), VO(II), and Cd(II) metal complexes were prepared using a novel hydrazone ligand (o-H2BMP). These synthesized compounds were structurally characterized through several spectroscopic techniques, which revealed that the ligand behaved either neutral bidentate with VO(II) or binegative tetradentate with Cu(II) and Cd(II). Electronic spectra as well as magnetic measurements suggested the octahedral configurations for Cu(II) and Cd(II); while the square pyramidal around the VO(II) ion. PXRD patterns referred that all compounds had crystal phase characteristics and crystal size within the nanoscale range. Moreover, the Qualx calculations related to the X-ray diffraction patterns showed that both the free ligand and the vanadium complex have a triclinic crystal system, while the copper and cadmium complexes have a monoclinic crystal system. As well, the compounds are computationally simulated using Gaussian 09 view via B3LYP functional and LANL2DZ/6-31+G(d,p) mixed basis set to predict the optimized structures. The DFT calculations reveal that all isolated compounds are chemically and biologically active. In addition, the electrochemical measurements performed for Cu(II) in absence/presence of ligand showed the formation of a (1:1) stable complex through a spontaneous process. Finally, the outcome data from biological activities for the all-isolated solid compounds were screened practically through antioxidant, antimicrobial, and cytotoxicity studies and screened theoretically using MOE docking, SAR, and drug-likeness prediction. The antioxidant activity of the VO(II) complex is more potent than Vit C in the DPPH method. Also, the antimicrobial activity values for the o-H2BMP ligand and its complexes (100 mL) revealed good activity compared to the gentamicin standard against B. subtilis. Finally, the IC50 values of VO(II) and Cd(II) complexes showed high performance in MDA-MB-231 and MCF-7, respectively.

Acknowledgments

This work was supported by the Chemistry Department, Faculty of Education and Applied Sciences—Arhab, Sana’a University, Yemen, and the Department of Chemistry, Faculty of Science, Mansoura University, Egypt.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.2c07592.

  • Biological screening method, IR correlation data, 1H NMR data, TGA curves, DFT calculations, molecular docking, and other supporting data (PDF)

Author Contributions

The manuscript was written through the contributions of all authors. All authors have given approval to the final version of the manuscript.

The authors declare no competing financial interest.

Supplementary Material

ao2c07592_si_001.pdf (4.2MB, pdf)

References

  1. Nur Mohammad E. U.; Uddin S.; Babar I. H.; Hossain S.; Bitu N. A.; Khan M. N.; Asraf A.; Hossen F.; Kudrat-E-Zahan Md. Exploring Schiff Base Chemistry-An Overview. Int. J. Chem. Pharm. Sci. 2021, 9, 18–31. [Google Scholar]
  2. Nassar M. Y.; Ahmed I. S.; Dessouki H. A.; Ali S. S. Synthesis and Characterization of Some Schiff Base Complexes Derived from 2, 5-Dihydroxyacetophenone with Transition Metal Ions and Their Biological Activity. J. Basic Environ. Sci. 2018, 5, 60–71. [Google Scholar]
  3. Silva Y. F.; Riga B. A.; Deflon V. M.; Souza J. R.; Silva L. H. F.; Machado A. E. H.; Maia P. I. S.; Valdemiro P C.-J.; Goi B. E. Organometallic-Mediated Radical Polymerization Using Well-Defined Schiff Base Cobalt (II) Complexes. J. Coord. Chem. 2018, 71, 3776–3789. 10.1080/00958972.2018.1527322. [DOI] [Google Scholar]
  4. Abubakar M. J.; Husaini M.; Nuhu A. H. Synthesis and Characterization of Schiff Base of 3-[(2-Hydroxy-Phenylimino)-Methyl]-6-Methoxy-Quinolin-2-Ol and Its Metal Complexes and Their Evaluation for Antibacterial and Antifungal Activity. Alger. J. Eng. Technol. 2020, 2, 29–36. [Google Scholar]
  5. Dalia S. A.; Afsan F.; Hossain M. S.; Khan M. N.; Zakaria C.; Zahan M. K. E.; Ali M. A Short Review on Chemistry of Schiff Base Metal Complexes and Their Catalytic Application. Int. J. Chem. Stud. 2018, 6, 2859–2866. [Google Scholar]
  6. Bekheit M. M.; El-Shobaky A. R.; Gad Allah M. T. Synthesis, Spectroscopic Characterization and Antimicrobial Studies of Some Metal Complexes with 2-Acetylpyridine Phenoxyacetyl Hydrazone (HAPPA). Arab. J. Chem. 2017, 10, S3064–S3072. 10.1016/j.arabjc.2013.11.048. [DOI] [Google Scholar]
  7. Hall I. H.; Peaty N. J.; Henry J. R.; Easmon J.; Heinisch G.; Pürstinger G. Investigations on the Mechanism of Action of the Novel Antitumor Agents 2-Benzothiazolyl, 2-Benzoxazolyl, and 2-Benzimidazolyl Hydrazones Derived from 2-Acetylpyridine. Arch. Pharm. 1999, 332, 115–123. . [DOI] [PubMed] [Google Scholar]
  8. Al Zoubi W.; Al-Hamdani A. A. S.; Ahmed S. D.; Ko Y. G. A New Azo-Schiff Base: Synthesis, Characterization, Biological Activity and Theoretical Studies of Its Complexes. Appl. Organomet. Chem. 2018, 32, e3895 10.1002/aoc.3895. [DOI] [Google Scholar]
  9. Bansod A.; Bhaskar R.; Ladole C.; Salunkhe N.; Thakare K.; Aswar A. Mononuclear Pyrazine-2-Carbohydrazone Metal Complexes: Synthesis, Structural Assessment, Thermal, Biological, and Electrical Conductivity Studies. Eur. J. Chem. 2022, 13, 126–134. 10.5155/eurjchem.13.1.126-134.2186. [DOI] [Google Scholar]
  10. Filipović N.; Todorović T.; Marković R.; Marinković A.; Tufegdžić S. T.; Godevac D.; Andelković K. Synthesis, Characterization and Biological Activities of N-Heteroaromatic Hydrazones and Their Complexes with Pd(II), Pt(II) and Cd(II). Transit. Met. Chem. 2010, 35, 765–772. 10.1007/s11243-010-9391-9. [DOI] [Google Scholar]
  11. Abd El-Hady M. N.; Zaky R. R.; Ibrahim K. M.; Gomaa E. A. (E)-3-(2-(Furan-Ylmethylene)Hydrazinyl)-3-Oxo-N-(Thiazol-2yl)Propanamide Complexes: Synthesis, Characterization and Antimicrobial Studies. J. Mol. Struct. 2012, 1016, 169–180. 10.1016/j.molstruc.2012.02.006. [DOI] [Google Scholar]
  12. Zaky R. R.; Yousef T. A. Spectral, Magnetic, Thermal, Molecular Modelling, ESR Studies and Antimicrobial Activity of (E)-3-(2-(2-Hydroxybenzylidene) Hydrazinyl)-3-Oxo- n(Thiazole-2-Yl)Propanamide Complexes. J. Mol. Struct. 2011, 1002, 76–85. 10.1016/j.molstruc.2011.06.050. [DOI] [Google Scholar]
  13. Abd El-Hady M. N.; Gomaa E. A.; Zaky R. R.; Gomaa A. I. Synthesis, Characterization, Computational Simulation, Cyclic Voltammetry and Biological Studies on Cu(II), Hg(II) and Mn(II) Complexes of 3-(3,5-Dimethylpyrazol-1-Yl)-3-Oxopropionitrile. J. Mol. Liq. 2020, 305, 112794. 10.1016/j.molliq.2020.112794. [DOI] [Google Scholar]
  14. Abd El-Hady M. N.; Gomaa E. A.; Al-Harazie A. G. Cyclic Voltammetry of Bulk and Nano CdCl2 with Ceftazidime Drug and Some DFT Calculations. J. Mol. Liq. 2019, 276, 970–985. 10.1016/j.molliq.2018.10.125. [DOI] [Google Scholar]
  15. Zaky R. R.; Yousef T. A.; Ibrahim K. M. Co (II), Cd (II), Hg (II) and U (VI) O2 Complexes of o-Hydroxyacetophenone [N-(3-Hydroxy-2-Naphthoyl)] Hydrazone: Physicochemical Study, Thermal Studies and Antimicrobial Activity. Spectrochim. Acta, Part A 2012, 97, 683–694. 10.1016/j.saa.2012.05.086. [DOI] [PubMed] [Google Scholar]
  16. Priya M. K.; Revathi B. K.; Renuka V.; Sathya S.; Asirvatham P. S. Molecular Structure, Spectroscopic (FT-IR, FT-Raman, 13C and 1H NMR) Analysis, HOMO-LUMO Energies, Mulliken, MEP and Thermal Properties of New Chalcone Derivative by DFT Calculation. Mater. Today Proc. 2019, 8, 37–46. 10.1016/j.matpr.2019.02.078. [DOI] [Google Scholar]
  17. Kalarani R.; Sankarganesh M.; Kumar G. V.; Kalanithi M. Synthesis, spectral, DFT calculation, sensor, antimicrobial and DNA binding studies of Co(II), Cu(II) and Zn(II) metal complexes with 2-amino benzimidazole Schiff base. J. Mol. Struct. 2020, 1206, 127725. 10.1016/j.molstruc.2020.127725. [DOI] [Google Scholar]
  18. Zhou X.; Pan Q. J.; Xia B. H.; Li M. X.; Zhang H. X.; Tung A. C. DFT and TD-DFT Calculations on the Electronic Structures and Spectroscopic Properties of Cyclometalated Platinum(II) Complexes. J. Phys. Chem. A 2007, 111, 5465–5472. 10.1021/jp064044r. [DOI] [PubMed] [Google Scholar]
  19. Moreira D. C.ABTS Decolorization Assay—In Vitro Antioxidant Capacity; Protocols.Io, 2019; Vol. 10–15. [Google Scholar]
  20. Zhou J.; Diao X.; Wang T.; Chen G.; Lin Q.; Yang X.; Xu J. Phylogenetic Diversity and Antioxidant Activities of Culturable Fungal Endophytes Associated with the Mangrove Species Rhizophora Stylosa and R. Mucronata in the South China Sea. PLoS One 2018, 13, e01973599 10.1371/journal.pone.0197359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Moon J.-K.; Shibamoto T. Antioxidant Assays for Plant and Food Components. J. Agric. Food Chem. 2009, 57, 1655–1666. 10.1021/jf803537k. [DOI] [PubMed] [Google Scholar]
  22. Mannaa A. H.; Zaky R. R.; Gomaa E. A.; Abd El-Hady M. N. Bivalent Transition Metal Complexes of Pyridine-2 , 6-Dicarbohydrazide : Structural Characterization , Cyclic Voltammetry and Biological Studies. J. Mol. Struct. 2022, 1269, 133852. 10.1016/j.molstruc.2022.133852. [DOI] [Google Scholar]
  23. Stylianakis I.; Kolocouris A.; Kolocouris N.; Fytas G.; Foscolos G. B.; Padalko E.; Neyts J.; De Clercq E. Spiro [Pyrrolidine-2, 2′-Adamantanes]: Synthesis, Anti-Influenza Virus Activity and Conformational Properties. Bioorg. Med. Chem. Lett. 2003, 13, 1699–1703. 10.1016/s0960-894x(03)00231-2. [DOI] [PubMed] [Google Scholar]
  24. Devi J.; Yadav M.; Jindal D. K.; Kumar D.; Poornachandra Y. Synthesis, Spectroscopic Characterization, Biological Screening and in Vitro Cytotoxic Studies of 4-Methyl-3-Thiosemicarbazone Derived Schiff Bases and Their Co (II), Ni (II), Cu (II) and Zn (II) Complexes. Appl. Organomet. Chem. 2019, 33, e5154 10.1002/aoc.5154. [DOI] [Google Scholar]
  25. Gutiérrez L.; Stepien G.; Gutiérrez L.; Pérez-Hernández M.; Pardo J.; Pardo J.; Grazú V.; de la Fuente J. M.. 1.09 - Nanotechnology in Drug Discovery and Development. In Comprehensive Medicinal Chemistry III; Chackalamannil S., Rotella D., Ward S. E., Eds.; Elsevier: Oxford, 2017; pp 264–295. [Google Scholar]
  26. Mitra I.; Mukherjee S.; Reddy B V. P.; Dasgupta S.; Bose K J. C.; Mukherjee S.; Linert W.; Moi S. C. Benzimidazole Based Pt(II) Complexes with Better Normal Cell Viability than Cisplatin: Synthesis, Substitution Behavior, Cytotoxicity, DNA Binding and DFT Study. RSC Adv. 2016, 6, 76600–76613. 10.1039/c6ra17788c. [DOI] [Google Scholar]
  27. Dhahagani K.; Mathan Kumar S.; Chakkaravarthi G.; Anitha K.; Rajesh J.; Ramu A.; Rajagopal G. Synthesis and Spectral Characterization of Schiff Base Complexes of Cu(II), Co(II), Zn(II) and VO(IV) Containing 4-(4-Aminophenyl)Morpholine Derivatives: Antimicrobial Evaluation and Anticancer Studies. Spectrochim. Acta, Part A 2014, 117, 87–94. 10.1016/j.saa.2013.07.101. [DOI] [PubMed] [Google Scholar]
  28. Denizot F.; Lang R. Rapid colorimetric assay for cell growth and survival. J. Immunol. Methods 1986, 89, 271–277. 10.1016/0022-1759(86)90368-6. [DOI] [PubMed] [Google Scholar]
  29. Hammoud M. M.; Nageeb A. S.; Morsi M. A.; Gomaa E. A.; Elmaaty A. A.; Al-Karmalawy A. A. Design, synthesis, biological evaluation, and SAR studies of novel cyclopentaquinoline derivatives as DNA intercalators, topoisomerase II inhibitors, and apoptotic inducers. New J. Chem. 2022, 46, 11422–11436. 10.1039/d2nj01646j. [DOI] [Google Scholar]
  30. Sibuh B. Z.; Khanna S.; Taneja P.; Sarkar P.; Taneja N. K. Molecular Docking, Synthesis and Anticancer Activity of Thiosemicarbazone Derivatives against MCF-7 Human Breast Cancer Cell Line. Life Sci. 2021, 273, 119305. 10.1016/j.lfs.2021.119305. [DOI] [PubMed] [Google Scholar]
  31. Abu-Melha K. S.; El-Metwally N. M. Spectral and Thermal Studies for Some Transition Metal Complexes of Bis(Benzylthiocarbohydrazone) Focusing on EPR Study for Cu(II) and VO2+. Spectrochim. Acta, Part A 2008, 70, 277–283. 10.1016/j.saa.2007.07.058. [DOI] [PubMed] [Google Scholar]
  32. Paul K. W.; Borda M. J.; Kubicki J. D.; Sparks D. L. Effect of Dehydration on Sulfate Coordination and Speciation at the Fe-(Hydr)Oxide-Water Interface: A Molecular Orbital/Density Functional Theory and Fourier Transform Infrared Spectroscopic Investigation. Langmuir 2005, 21, 11071–11078. 10.1021/la050648v. [DOI] [PubMed] [Google Scholar]
  33. Ndwandwe S.; Tshibangu P.; Dikio E. D. Synthesis of Carbon Nanospheres from Vanadium β-Diketon Catalyst. Int. J. Electrochem. Sci. 2011, 6, 749–760. [Google Scholar]
  34. Salagram M. Infrared spectrum of VO2+ entity in K2C2O4·H2O crystals. Phys. Status Solidi A 1988, 105, 161–164. 10.1002/pssa.2211050260. [DOI] [Google Scholar]
  35. Zaky R. R.; Ibrahim K. M.; Gabr I. M. Bivalent Transition Metal Complexes of O-Hydroxyacetophenone [N-(3-Hydroxy-2-Naphthoyl)] Hydrazone: Spectroscopic, Antibacterial, Antifungal Activity and Thermogravimetric Studies. Spectrochim. Acta, Part A 2011, 81, 28–34. 10.1016/j.saa.2011.05.028. [DOI] [PubMed] [Google Scholar]
  36. 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, e6910 10.1002/aoc.6910. [DOI] [Google Scholar]
  37. Al-Hazmi G. A.; Abou-Melha K. S.; El-Metwaly N. M.; Saleh K. A. Synthesis of Novel VO(II)-Perimidine Complexes: Spectral, Computational, and Antitumor Studies. Bioinorg. Chem. Appl. 2018, 2018, 1–22. 10.1155/2018/7176040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Hassoon A. A.; Al-Radadi N. S.; Nawar N.; Mostafa M. M. New Square-Pyramidal Oxovanadium (IV) Complexes Derived from Polydentate Ligand (L&amp;lt;sup&amp;gt;1&amp;lt;/sup&amp;gt;). Open J. Inorg. Chem. 2016, 06, 23–65. 10.4236/ojic.2016.61003. [DOI] [Google Scholar]
  39. Abou-Hussen A. A.; El-Metwally N. M.; Saad E. M.; El-Asmy A. A. Spectral, Magnetic, Thermal and Electrochemical Studies on Phthaloyl Bis(Thiosemicarbazide) Complexes. J. Coord. Chem. 2005, 58, 1735–1749. 10.1080/00958970500262270. [DOI] [Google Scholar]
  40. Singh V. P. S. Synthesis, electronic and ESR spectral studies on copper(II) nitrate complexes with some acylhydrazines and hydrazones. Spectrochim. Acta, Part A 2008, 71, 17–22. 10.1016/j.saa.2007.11.004. [DOI] [PubMed] [Google Scholar]
  41. Manikandan P.; Muthukumaran R.; Thomas K. R. J.; Varghese B.; Chandramouli G. V. R.; Manoharan P. T. Copper(II) Azide Complexes of Aliphatic and Aromatic Amine Based Tridentate Ligands: Novel Structure, Spectroscopy, and Magnetic Properties. Inorg. Chem. 2001, 40, 2378–2389. 10.1021/ic0009223. [DOI] [PubMed] [Google Scholar]
  42. Verquin G.; Fontaine G.; Abi-Aad E.; Zhilinskaya E.; Aboukaïs A.; Bernier J. L. EPR Study of Copper(II) Complexes of Hydroxysalen Derivatives in Order to Be Used in the DNA Cleavage. J. Photochem. Photobiol., B 2007, 86, 272–278. 10.1016/j.jphotobiol.2006.12.003. [DOI] [PubMed] [Google Scholar]
  43. Mustafi D.; Galtseva E. V.; Krzystek J.; Brunel L. C.; Makinen M. W. High-Frequency Electron Paramagnetic Resonance Studies of VO 2 + in Low-Temperature Glasses. J. Phys. Chem. A 1999, 103, 11279–11286. 10.1021/jp991287t. [DOI] [Google Scholar]
  44. Arber J. M.; Sharpe P. H. G.; Joly H. A.; Morton J. R.; Preston K. F. The ESR/Alanine Dosimeter-Power Dependence of the X-Band Spectrum. Int. J. Radiat. Appl. Instrum., Part A 1991, 42, 665–668. 10.1016/0883-2889(91)90037-2. [DOI] [Google Scholar]
  45. Hathaway B. J.; Billing D. E. The Electronic Properties and Stereochemistry of Mono-Nuclear Complexes of the Copper(II) Ion. Coord. Chem. Rev. 1970, 5, 143–207. 10.1016/s0010-8545(00)80135-6. [DOI] [Google Scholar]
  46. Fetoh A.; Mohammed M. A.; Youssef M. M.; Abu El-Reash G. M. Characterization, Cyclic Voltammetry and Biological Studies of Divalent Co, Ni and Cu Complexes of Water-Soluble, Bioactive and Photoactive Thiosemicarbazone Salt. J. Mol. Liq. 2019, 287, 110958. 10.1016/j.molliq.2019.110958. [DOI] [Google Scholar]
  47. El-Metwally N. M.; El-Shazly R. M.; Gabr I. M.; El-Asmy A. A. Physical and Spectroscopic Studies on Novel Vanadyl Complexes of Some Substituted Thiosemicarbazides. Spectrochim. Acta, Part A 2005, 61, 1113–1119. 10.1016/j.saa.2004.06.027. [DOI] [PubMed] [Google Scholar]
  48. McGarvey B. R. The Isotropic Hyperfine Interaction. J. Phys. Chem. 1967, 71, 51–66. 10.1021/j100860a007. [DOI] [Google Scholar]
  49. Onay H.; Yerli Y.; Öztürk R. Synthesis and EPR Studies of Vanadyl Tetrakis(Selenodiazole)Porphyrazine. Transit. Met. Chem. 2009, 34, 163–166. 10.1007/s11243-008-9172-x. [DOI] [Google Scholar]
  50. Karabulut B.; ilkin I.; Tapramaz R. EPR and Optical Absorption Studies of VO2+Doped Trisodium Citrate Dihydrate Single Crystals. Z. Naturforsch., A: Phys. Sci. 2005, 60, 95–100. 10.1515/zna-2005-1-216. [DOI] [Google Scholar]
  51. Kalkhoran S.; Benowitz N. L.; Rigotti N. A. Reprint of: Prevention and Treatment of Tobacco Use. J. Am. Coll. Cardiol. 2018, 72, 2964–2979. 10.1016/j.jacc.2018.10.020. [DOI] [PubMed] [Google Scholar]
  52. Li L.; Zhang F.; Zaia J.; Linhardt R. J. Top-down Approach for the Direct Characterization of Low Molecular Weight Heparins Using LC-FT-MS. Anal. Chem. 2012, 84, 8822–8829. 10.1021/ac302232c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Mazzarino M.; De La Torre X.; Botrè F.; Gray N.; Cowan D. A Rapid Screening LC-MS/MS Method Based on Conventional HPLC Pumps for the Analysis of Lowmolecular Weight Xenobiotics: Application to Doping Control Analysis. Drug Test. Anal. 2010, 2, 311–322. 10.1002/dta.148. [DOI] [PubMed] [Google Scholar]
  54. Fetoh A.; Mohammed M. A.; Youssef M. M.; Abu El-Reash G. M. Characterization, Cyclic Voltammetry and Biological Studies of Divalent Co, Ni and Cu Complexes of Water-Soluble, Bioactive and Photoactive Thiosemicarbazone Salt. J. Mol. Liq. 2019, 287, 110958. 10.1016/j.molliq.2019.110958. [DOI] [Google Scholar]
  55. Fetoh A.; Mohammed M. A.; Youssef M. M.; Abu El-Reash G. M. Synthesis, Characterization, Cyclic Voltammetry and Biological Studies of Zn (II), Cd (II), Hg (II) and UO 22+ Complexes of Thiosemicarbazone Salt. Appl. Organomet. Chem. 2019, 33, e4787 10.1002/aoc.4787. [DOI] [Google Scholar]
  56. Coats A. W.; Redfern J. P. Kinetic Parameters from Thermogravimetric Data. Nature 1964, 201, 68–69. 10.1038/201068a0. [DOI] [Google Scholar]
  57. Horowitz H. H.; Metzger G. A New Analysis of Thermogravimetric Traces. Anal. Chem. 1963, 35, 1464–1468. 10.1021/ac60203a013. [DOI] [Google Scholar]
  58. Hatakeyama T.; Quinn F. X.. Thermal Analysis: Fundamentals and Applications to Polymer Science; Wiley, 1999. [Google Scholar]
  59. Frost A. A.; Pearson R. G.. Kinetics and Mechanism; John Wiley and Sons. Inc.: New York, 1961; p 19. [Google Scholar]
  60. Kandil S. S.; El-Hefnawy G. B.; Baker E. A. Thermal and Spectral Studies of 5-(Phenylazo)-2-Thiohydantoin and 5-(2-Hydroxyphenylazo)-2-Thiohydantoin Complexes of Cobalt (II), Nickel (II) and Copper (II). Thermochim. Acta 2004, 414, 105–113. 10.1016/j.tca.2003.11.021. [DOI] [Google Scholar]
  61. 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, 35–41. 10.1016/s0040-6031(98)00548-6. [DOI] [Google Scholar]
  62. 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. 10.1016/0040-6031(90)80121-e. [DOI] [Google Scholar]
  63. Huang H.; Leung D. Y. C.; Ye D. Effect of Reduction Treatment on Structural Properties of TiO2 Supported Pt Nanoparticles and Their Catalytic Activity for Formaldehyde Oxidation. J. Mater. Chem. 2011, 21, 9647–9652. 10.1039/c1jm10413f. [DOI] [Google Scholar]
  64. 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. 10.1016/j.molstruc.2022.134139. [DOI] [Google Scholar]
  65. Ma S.; Zhang J.; Liu Y.; Qian J.; Xu B.; Tian W. Direct Observation of the Symmetrical and Asymmetrical Protonation States in Molecular Crystals. J. Phys. Chem. Lett. 2017, 8, 3068–3072. 10.1021/acs.jpclett.7b01454. [DOI] [PubMed] [Google Scholar]
  66. Wu Z. Y.; Li Y. T.; Xu D. J. Diaqua(2,2′-Diamino-4,4′-Bi-1,3-Thiazole)Oxosulfatovanadium(IV) Tetrahydrate. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2005, 61, 463–465. 10.1107/s0108270105028982. [DOI] [PubMed] [Google Scholar]
  67. Puttreddy R.; Von Essen C.; Peuronen A.; Lahtinen M.; Rissanen K. Halogen Bonds in 2,5-Dihalopyridine-Copper(Ii) Chloride Complexes. CrystEngComm 2018, 20, 1954–1959. 10.1039/c8ce00209f. [DOI] [Google Scholar]
  68. Kukovec B. M.; Popović Z.; Pavlović G. A One-Dimensional CdII Coordination Polymer: Catena-Poly[Cadmium(II)-Bis-(μ-6-Methyl-Picolinato)]. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2007, 63, 389–391. 10.1107/s0108270107034646. [DOI] [PubMed] [Google Scholar]
  69. Rezaee R.; Montazer M.; Mianehro A.; Mahmoudirad M. Single-Step Synthesis and Characterization of Zr-MOF onto Wool Fabric: Preparation of Antibacterial Wound Dressing with High Absorption Capacity. Fibers Polym. 2022, 23, 404–412. 10.1007/s12221-021-0211-y. [DOI] [Google Scholar]
  70. Tamer Ö.; Avcı D.; Dege N.; Atalay Y. S. Synthesis, crystal structure, photophysical properties, density functional theory calculations and molecular docking studies on Cd(II) complex of 4,4′-dimethyl-2,2′-dipyridyl. J. Mol. Struct. 2020, 1202, 127288. 10.1016/j.molstruc.2019.127288. [DOI] [Google Scholar]
  71. Al-Fahemi J. H.; Abdallah M.; Gad E. A. M.; Jahdaly B. A. A. L. Experimental and Theoretical Approach Studies for Melatonin Drug as Safely Corrosion Inhibitors for Carbon Steel Using DFT. J. Mol. Liq. 2016, 222, 1157–1163. 10.1016/j.molliq.2016.07.085. [DOI] [Google Scholar]
  72. Frisch M.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson Ga.; et al. Gaussian 09, Revision D. 01; Gaussian, Inc.: Wallingford CT, 2009.
  73. Addison A. W.; Rao T. N.; Reedijk J.; van Rijn J.; Verschoor G. C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc., Dalton Trans. 1984, 1349–1356. 10.1039/dt9840001349. [DOI] [Google Scholar]
  74. Fedorova O. A.; Shepel N. E.; Tokarev S. D.; Lukovskaya E. V.; Sotnikova Y. A.; Moiseeva A. A.; D’Aléo A.; Fages F.; Maurel F.; Fedorov Y. V. Intramolecular Electron Transfer in Cu(Ii) Complexes with Aryl-Imidazo-1,10-Phenanthroline Derivatives: Experimental and Quantum Chemical Calculation Studies. New J. Chem. 2019, 43, 2817–2827. 10.1039/c8nj05697h. [DOI] [Google Scholar]
  75. Altürk S.; Avcı D.; Başoğlu A.; Tamer Ö.; Atalay Y.; Dege N. Copper(II) Complex with 6-Methylpyridine-2-Carboxyclic Acid: Experimental and Computational Study on the XRD, FT-IR and UV–Vis Spectra, Refractive Index, Band Gap and NLO Parameters. Spectrochim. Acta, Part A 2018, 190, 220–230. 10.1016/j.saa.2017.09.041. [DOI] [PubMed] [Google Scholar]
  76. 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, e10355 10.1016/j.heliyon.2022.e10355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Chaudhary M. K.; Srivastava A.; Singh K. K.; Tandon P.; Joshi B. D. Computational Evaluation on Molecular Stability, Reactivity, and Drug Potential of Frovatriptan from DFT and Molecular Docking Approach. Comput. Theor. Chem. 2020, 1191, 113031. 10.1016/j.comptc.2020.113031. [DOI] [Google Scholar]
  78. Toubi Y.; Abrigach F.; Radi S.; Souna F.; Hakkou A.; Alsayari A.; Bin Muhsinah A.; Mabkhot Y. N. Synthesis, Antimicrobial Screening, Homology Modeling, and Molecular Docking Studies of a New Series of Schiff Base Derivatives as Prospective Fungal Inhibitor Candidates. Molecules 2019, 24, 3250. 10.3390/molecules24183250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Bayoumy A. M.; Ibrahim M.; Omar A. Mapping Molecular Electrostatic Potential (MESP) for Fulleropyrrolidine and Its Derivatives. Opt. Quant. Electron. 2020, 52, 346–413. 10.1007/s11082-020-02467-6. [DOI] [Google Scholar]
  80. El-Shereafy S. E.; Gomaa E. A.; Yousif A. M.; Abou El-Yazed A. S. Electrochemical and Thermodynamic Estimations of the Interaction Parameters for Bulk and Nano-Silver Nitrate (NSN) with Cefdinir Drug Using a Glassy Carbon Electrode. Iran J. Mater. Sci. Eng. 2017, 14, 48–57. 10.22068/ijmse.14.4.48. [DOI] [Google Scholar]
  81. Hoogerheide J. G.; Popov A. I. Study of Monensin Complexes with Monovalent Metal Lons in Anhydrous Methanol Solutions. J. Solution Chem. 1978, 7, 357–372. 10.1007/bf00662896. [DOI] [Google Scholar]
  82. Gomaa E. A.; Morsi M. A.; Negm A. E.; Sherif Y. A. Cyclic Voltammetry of Bulk and Nano Manganese Sulfate with Doxorubicin Using Glassy Carbon Electrode. Int. J. Nano Dimens. 2017, 8, 89–96. 10.22034/IJND.2017.24380. [DOI] [Google Scholar]
  83. Chen Z.; Bertin R.; Froldi G. EC50 Estimation of Antioxidant Activity in DPPH* Assay Using Several Statistical Programs. Food Chem. 2013, 138, 414–420. 10.1016/j.foodchem.2012.11.001. [DOI] [PubMed] [Google Scholar]
  84. Abdelghany A. M.; Menazea A. A.; Ismail A. M. Synthesis, Characterization and Antimicrobial Activity of Chitosan/Polyvinyl Alcohol Blend Doped with Hibiscus Sabdariffa L. Extract. J. Mol. Struct. 2019, 1197, 603–609. 10.1016/j.molstruc.2019.07.089. [DOI] [Google Scholar]
  85. Elimam D. M.; Elgazar A. A.; El-Senduny F. F.; El-Domany R. A.; Badria F. A.; Eldehna W. M. Natural Inspired Piperine-Based Ureas and Amides as Novel Antitumor Agents towards Breast Cancer. J. Enzyme Inhib. Med. Chem. 2022, 37, 39–50. 10.1080/14756366.2021.1988944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. El-Senduny F. F.; Shabana S. M.; Rosel D.; Brabek J.; Althagafi I.; Angeloni G.; Manolikakes G.; Shaaban S. Urea-Functionalized Organoselenium Compounds as Promising Anti-HepG2 and Apoptosis-Inducing Agents. Future Med. Chem. 2021, 13, 1655–1677. 10.4155/fmc-2021-0114. [DOI] [PubMed] [Google Scholar]
  87. Abdellatif K. R. A.; Belal A.; El-Saadi M. T.; Amin N. H.; Said E. G.; Hemeda L. R. Design, Synthesis, Molecular Docking and Antiproliferative Activity of Some Novel Benzothiazole Derivatives Targeting EGFR/HER2 and TS. Bioorg. Chem. 2020, 101, 103976. 10.1016/j.bioorg.2020.103976. [DOI] [PubMed] [Google Scholar]
  88. Kareem F. A. K.; Al-Hujaj H. H.; Jassem A. M.; Al-Masoudi N. A. A Click Synthesis, Molecular Docking, Cytotoxicity on Breast Cancer (MDA-MB 231) and Anti-HIV Activities of New 1,4-Disubstituted-1,2,3-Triazole Thymine Derivatives. Russ. J. Bioorg. Chem. 2020, 46, 360–370. 10.1134/s1068162020030024. [DOI] [Google Scholar]
  89. Chang E. C.; Frasor J.; Komm B.; Katzenellenbogen B. S. Impact of Estrogen Receptor β on Gene Networks Regulated by Estrogen Receptor α in Breast Cancer Cells. Endocrinology 2006, 147, 4831–4842. 10.1210/en.2006-0563. [DOI] [PubMed] [Google Scholar]
  90. Dhumad A. M.; Jassem A. M.; Alharis R. A.; Almashal F. A. Design, Cytotoxic Effects on Breast Cancer Cell Line (MDA-MB 231), and Molecular Docking of Some Maleimide-Benzenesulfonamide Derivatives. J. Indian Chem. Soc. 2021, 98, 100055. 10.1016/j.jics.2021.100055. [DOI] [Google Scholar]
  91. Trott O.; Olson A. J. AutoDock Vina: Improving the Speed and Accuracy of Docking with a New Scoring Function, Efficient Optimization, and Multithreading. J. Comput. Chem. 2010, 31, 455–461. 10.1002/jcc.21334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Carcelli M.; Mazza P.; Pelizzi C.; Zani F. Antimicrobial and Genotoxic Activity of 2,6-Diacetylpyridine Bis(Acylhydrazones) and Their Complexes with Some First Transition Series Metal Ions. X-Ray Crystal Structure of a Dinuclear Copper(II) Complex. J. Inorg. Biochem. 1995, 57, 43–62. 10.1016/0162-0134(94)00004-t. [DOI] [PubMed] [Google Scholar]
  93. Farhadi F.; Khameneh B.; Iranshahi M.; Iranshahy M. Antibacterial Activity of Flavonoids and Their Structure–Activity Relationship: An Update Review. Phyther. Res. 2019, 33, 13–40. 10.1002/ptr.6208. [DOI] [PubMed] [Google Scholar]
  94. Aljahdali M.; El-Sherif A. A. Synthesis, Characterization, Molecular Modeling and Biological Activity of Mixed Ligand Complexes of Cu (II), Ni (II) and Co (II) Based on 1, 10-Phenanthroline and Novel Thiosemicarbazone. Inorg. Chim. Acta. 2013, 407, 58–68. 10.1016/j.ica.2013.06.040. [DOI] [Google Scholar]
  95. Gacche R. N.; Jadhav S. G. Antioxidant Activities and Cytotoxicity of Selected Coumarin Derivatives: Preliminary Results of a Structure-Activity Relationship Study Using Computational Tools. J. Exp. Clin. Med. 2012, 4, 165–169. 10.1016/j.jecm.2012.04.007. [DOI] [Google Scholar]
  96. Parrilha G. L.; Da Silva J. G.; Gouveia L. F.; Gasparoto A. K.; Dias R. P.; Rocha W. R.; Santos D. A.; Speziali N. L.; Beraldo H. Pyridine-Derived Thiosemicarbazones and Their Tin(IV) Complexes with Antifungal Activity against Candida Spp. Eur. J. Med. Chem. 2011, 46, 1473–1482. 10.1016/j.ejmech.2011.01.041. [DOI] [PubMed] [Google Scholar]
  97. Heidarpoor Saremi L.; Ebrahimi A.; Lagzian M. Identification of New Potential Cyclooxygenase-2 Inhibitors: Insight from High Throughput Virtual Screening of 18 Million Compounds Combined with Molecular Dynamic Simulation and Quantum Mechanics. J. Biomol. Struct. Dyn. 2021, 39, 1717–1734. 10.1080/07391102.2020.1737574. [DOI] [PubMed] [Google Scholar]
  98. Zaky R.; Fekri A. Solid State Ball Milling as a Green Approach to Prepare Cu(II) Complexes: Structural, Spectral, DFT, and DNA Studies. New J. Chem. 2017, 41, 4555–4563. 10.1039/c7nj00840f. [DOI] [Google Scholar]
  99. Fekri A.; Zaky R. Solvent-Free Synthesis and Computational Studies of Transition Metal Complexes of the Aceto- and Thioaceto-Acetanilide Derivatives. J. Organomet. Chem. 2016, 818, 15–27. 10.1016/j.jorganchem.2016.05.015. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao2c07592_si_001.pdf (4.2MB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES