Next Article in Journal
Transcriptomic Analysis of the Response of Susceptible and Resistant Bitter Melon (Momordica charantia L.) to Powdery Mildew Infection Revealing Complex Resistance via Multiple Signaling Pathways
Previous Article in Journal
A High-Throughput Small-Angle X-ray Scattering Assay to Determine the Conformational Change of Plasminogen
Previous Article in Special Issue
Chemical Diversity of Mo5S5 Clusters with Pyrazole: Synthesis, Redox and UV-vis-NIR Absorption Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of Novel Nano-Sized Imine Complexes Using Coriandrum sativum Extract: Structural Elucidation, Non-Isothermal Kinetic Study, Theoretical Investigation and Pharmaceutical Applications

by
Shimaa Hosny
1,*,
Randa F. Abd El-Baki
1,
Zeinab H. Abd El-Wahab
2,
Gamal A. Gouda
3,
Mohammed S. Saddik
4,
Ateyatallah Aljuhani
5 and
Ahmed M. Abu-Dief
5,6,*
1
Department of Chemistry, Faculty of Science, New Valley University, Alkharga 72511, Egypt
2
Department of Chemistry, Faculty of Science (Girl’s), Al-Azhar University, Cairo 11754, Egypt
3
Department of Chemistry, Faculty of Science, Al-Azhar University, Assiut Branch, Assiut 71524, Egypt
4
Department of Pharmaceutics and Clinical Pharmacy, Faculty of Pharmacy, Sohag University, P.O. Box 82524, Sohag 82524, Egypt
5
Chemistry Department, College of Science, Taibah University, Madinah 42353, Saudi Arabia
6
Chemistry Department, Faculty of Science, Sohag University, Sohag 82524, Egypt
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(18), 14259; https://doi.org/10.3390/ijms241814259
Submission received: 11 August 2023 / Revised: 7 September 2023 / Accepted: 11 September 2023 / Published: 19 September 2023
(This article belongs to the Special Issue The Design, Synthesis and Study of Metal Complexes)

Abstract

:
A new Schiff base (H2L) generated from sulfamethazine (SMT), as well as its novel micro- and nanocomplexes with Ni(II) and Cd(II) metal ions, have been synthesized. The proposed structures of all isolated solid compounds were identified with physicochemical, spectral, and thermal techniques. Molar conductance studies confirmed that the metal complexes are not electrolytic. The molecular geometry located at the central metal ion was found to be square planar for the NiL2 and tetrahedral for the CdL2 complexes. The kinetic and thermal parameters were obtained using the Coats and Redfern approach. Coriandrum sativum (CS) in ethanol was used to create the eco-friendly Ni and Cd nanocomplexes. The size of the obtained nanoparticles was examined using PXRD and TEM, and found to be in the sub-nano range (3.07–4.61 nm). Furthermore, the TEM micrograph demonstrated a uniform and homogeneous surface morphology. The chemistry of the prepared nanocomplexes was studied using TGA and TEM techniques. The effect of temperature on the prepared nanocomplexes’ size revealed a decrease in size by heating. Furthermore, the nanocomplexes’ antimicrobial and anticancer properties were evaluated. The outcomes demonstrated that the nanocomplexes exhibited better antimicrobial properties. Moreover, the antitumor results showed that after heating, the Ni nanocomplex exhibited a substantial antitumor activity (IC50 = 1.280 g/mL), which was higher than the activity of cis-platin (IC50 = 1.714 g/mL). Finally, molecular-docking studies were performed to understand the evaluated compounds’ ability to bind to methionine adenosyl-transferases (PDB ID: 5A19) in liver cancer and COVID-19 main protease (PDB ID: 6lu7) cell-proteins. The findings reveal that [NiL2]·1.5H2O2 has a higher binding energy of −37.5 kcal/mol with (PDB ID: 5A19) cell protein.

1. Introduction

Schiff bases are formed through the condensation of primary amines and carbonyl compounds to generate imines or azomethine groups [1]. Schiff bases are important in the synthesis of Schiff base complexes because these ligands have the ability to form stable compounds with metal ions [2]. In addition to this, Schiff bases and their complexes have a broad range of biological activities [3,4].
On the other hand, sulfonamides, generally referred to as sulfa medicines, were the first routinely used chemotherapeutics for treating and preventing bacterial infections in humans [5]. Due to their uses, a wide variety of families of organic and inorganic compounds are now being studied. One of the most common compounds is sulfonamide and its N-derivatives [6]. Numerous Schiff bases derived from sulfonamide have been synthesized and used as chelates in the synthesis of strong metal chelates. Sulfamethazine is a sulfa-based antibacterial drug used for treating livestock diseases [7]. Like other sulfonamides, sulfamethazine has been modified through the creation of Schiff bases [8] and metal complexes [9]. The creation of nano-sized composites results in the synthesis of new compounds with distinctive physical, chemical, and biological characteristics [10].
Because of the huge number of surface atoms relative to atoms, nanoparticles (1–100 nm) generated by nanotechnological processes exhibit various novel physicochemical features in comparison to macro-sized structures [11,12]. Many studies on the creation of nanoparticles, which have a wide range of uses, have lately been carried out, and significant findings have been obtained. Although the physical and chemical technologies employed in nanoparticle synthesis enable the manufacture of nanoparticles of any size in a short amount of time, their toxicity is high [13]. In order to obtain harmless types of nanoparticles, it is necessary to employ newer and more practical techniques
The green synthesis of nano-metal-complexes offers an alternative, effective, low-cost, and eco-friendly way to create nanoparticles with defined geometries [14,15]. Researchers have succeeded in creating different nanoparticles for this purpose through the use of plant extracts and microorganisms [16]. The manufacture of nanoparticles using plants has not received much attention, according to a review of the relevant scientific literature. The production of nanoparticles from plant extracts is faster than the microbial synthesis, and the resulting nanoparticles have a more stable structure [17].
In this article, we have synthesized two novel micro- and nanocomposites with a ligand derived from sulfamethazine and 2-hydroxy-1-naphthaldehyde. Numerous methods, including physicochemical and analytical tools, were applied to describe the produced Schiff base and its micro- and nanocomposites. The chemistry of the generated nano-composites was studied using TGA and TEM techniques. Also, computational investigations, such as DFT and molecular docking, were carried out on the prepared compounds. The kinetics and thermodynamic parameters for some thermal decomposition steps have been critically studied. Moreover, the in vitro antibacterial, antifungal, and anticancer activities of the prepared compounds were examined.

2. Results and Discussion

2.1. Physical, Microanalytical, and Molar Conductance Data

According to the complex’s molar conductance measurement, which was carried out in 10−3 M DMF solvent and was between 0.7 and 12 ohm−1 cm2 mol−1, the complexes under examination have a non-electrolytic nature. The microanalytical data revealed that all metal chelates were synthesized with a 1:2 (M:L) stoichiometry (Table 1).

2.2. 1H-NMR

The 1H NMR spectrum of the newly prepared H2L ligand was recorded in DMSO-d6 (Figure S1). The peaks appear at δ 2.21, δ 9.64, and δ 2.51 ppm, which may be assigned to the CH3, OH, and DMSO-d6 [18] protons. The proton of azomethine was detected at δ 8.58 ppm [18], confirming the production of the Schiff base ligand. A signal at δ 10.61 ppm is due to the proton of the NH group [19]. The signals recorded between δ 7.24 and δ 7.91 ppm are characteristic of the aromatic protons [20].

2.3. Infrared Spectra

2.3.1. IR Spectroscopy of Metal Complexes

The chelation mode between the ligand and the metal ions may be established by comparing the complex IR spectra to those of the free ligand (Figure S2). The finding proves that the ligand behaves in a bidentate manner. The characteristic vibrational bands were identified at 3200, 3457, 1384, and 1156 cm−1 for the free H2L ligand, assigned to υ(OH), υ(NH), υas(SO2), and υsy(SO2) groups, respectively. The (HC=N) vibration of the ligand at 1633 cm−1 indicates the synthesis of the H2L ligand; however, in the complexes of Ni(II) and Cd(II), this peak migrated to 1602 and 1624 cm−1, due to its coordination [21]. In the Ni(II) and Cd(II) complexes, the phenolic (O-H) of the H2L vanished, demonstrating proton transfer during complex formation. The persistence of the υ(NH) at 3457, υas(SO2) at 1384 cm−1, and υsy(SO2) at 1156 cm−1 in the spectra of the complexes and their free ligand rules out the coordination through these groups [22]. Finally, the peaks observed for the metal chelates at 544–549 and 566–585 cm−1 were assigned to (metal–nitrogen) and (metal–oxygen), respectively. Moreover, the band at 970 in the spectrum of the Ni(II) chelate could be ascribed to the (H2O) [23] (Table S1).

2.3.2. IR Spectroscopy of Nano-Complexes

The binding modes of the Ni and Cd nanocomposites produced in CS/EtOH media (Figure S3) are revealed by the FTIR spectra. This is corroborated by a shift in the peak location of the nanocomposites as compared to the free ligand.

2.4. Mass Spectra

Mass spectrometry has been used successfully to confirm the H2L Schiff base’s molecular ion peaks and its metal complexes. The observed mass spectrum of the H2L ligand indicates a molecular ion peak (M+) at 432.50, which strongly supports the proposed formula. The fragment pattern of the H2L ligand provides an impression of the possible deterioration, with a sequence of peaks corresponding to different fragments. The suggested fragmentation pattern of the H2L is given in a supplementary data file (Scheme S1). The MS of Ni(II) and Cd(II) complexes showed molecular ion peaks at 948.73 and 975.79, respectively, which suited the molecular weights of these complexes. They also showed that Ni and Cd isotope peaks were present at m/z 60 and 114, respectively (Figure 1).

2.5. XRD and Morphological Studies

2.5.1. XRD of Nanocomplexes

The obtained Ni and Cd nanocomplexes in CS/EtOH media were examined by the powder XRD method (Figure 2). XRD patterns suggested that the prepared nanocomplexes lie between amorphous and crystalline characters. As a result, it was found that nanocomplexes were unsuitable for single-crystal XRD. The peak broadening indicated that the prepared compounds are nanoscale in size [24]. The XRD pattern shows that the crystal size can be calculated according to the Debye–Scherrer formula [18,25]. The average crystal sizes of Ni and Cd nanocomplexes are 33.1 and 36.4 nm, respectively.

2.5.2. TEM Analysis

The morphology and size of Ni and Cd nanoparticles created in Coriandrum sativum (CS) media were studied, as well as the impact of heat on the prepared nanocomplexes after heating them at 200 °C for two hours using the TEM technique (Figure 3). The images demonstrate particle morphologies to be homogeneous and close together, which is evidence that there are identical matrices. The mean particle size of Ni and Cd nanocomplexes before heating was about 24.55 and 33.89 nm, respectively. However, after heating, the Ni and Cd nanocomposites at 200 °C were 4.61 and 3.07 nm (sub-nano), respectively, indicating that heating reduces the particle size of the formed nanocomplexes. These findings accord well with the estimated crystallite size values from the XRD technique.

2.6. Thermal Behavior

2.6.1. TG-DTG of Complexes

The TG-DTG thermograms of Ni(II) and Cd(II) chelates are shown in Figure 4. The TG curve of the Ni(II) chelate exposed four decomposition steps. The first step from 52 to 111 °C with an estimated mass loss of 2.32% (calc. 2.84%) corresponds to the removal of the non-coordinated H2O molecules. The second step, from 265 to 321 °C with a mass loss of 20.93% (calc. 22.58%), corresponds to the removal of 2C6H8N2. The third decomposition step occurred at temperatures ranging from 404 to 566 °C with a mass loss of 21.70% (calc. 22.79%), corresponding to the loss of SO2 and 2C6H6. At a higher temperature of 727–767 °C, a mass loss of 30.01% (calc. 30.18%) is attributed to the elimination of 2C10H7O. The TG curve of the Cd(II) chelate exposed a three-step decomposition. The Cd(II) complex showed no mass loss up to 271 °C, confirming the absence of coordinated water and the complex’s high thermal stability. The mass losses in temperature from 271 to 318 °C with a mass loss of 21.69% (calc. 21.96%) may be attributed to the decomposition of 2C6H8N2 followed by the loss of 2C6H6, SO2 and 2HCN with a mass loss of 26.19% (calc. 27.71%) within the temperature from 420 to 481 °C. After this decomposition, the mass loss of the third decomposition stage was within the range from 772 to 811 °C with a mass loss of 30.42% (calc. 29.14%), corresponding to the removal of the rest of 2C10H7O. In Ni(II) and Cd(II) chelates, metal oxides were left as a residue.

2.6.2. TG-DTG of Nanocomplexes

The thermogram of nickel and cadmium nanodomain metal chelates prepared in Coriandrum sativum (CS) extract in 20% ethanol aims to discover the chemistry of nanodomain metal chelates by evaluating the size of the nanodomain metal chelates at each step of thermal heating to ascertain how the heat affects the size of the nanodomain metal chelates (Figure 5). The thermograms of Ni and Cd nanodomain were studied before and after heating the nanodomain chelates at 200 °C for 2 h. The mean particle size of Ni and Cd nano-chelates before heating was about 10.11–21.39 and 12.90–31.96 nm, respectively. However, after heating, the Ni and Cd nano-chelates at 200 °C were 2.05–39.6 nm (sub-nano) and 1.08–2.78 nm (sub-nano), respectively, indicating that heating reduces the particle size of the formed nano-chelates (Table 2 and Table 3).

2.6.3. Kinetic Calculations

In the present work, the kinetic and thermodynamic parameters, the energy of activation (Ea), the activation enthalpy (ΔH*), the entropy of activation (ΔS*), the Gibbs energy change (ΔG*), the order of the reaction (n), the correlation coefficient (r), and the pre-exponential factor (Z) for the non-isothermal decomposition of the prepared metal complexes were determined by the integral method proposed by Coats and Redfern [26]. The obtained data are listed in Table 4. Using the following equations, the activation enthalpy (ΔH*), the activation entropy (ΔS*), and the free energy (ΔG*) were determined.
ΔS* = 2.303 [log (Zh/KT)]R
ΔH* = E − RT
ΔG* = ΔH − TS ΔS
The high Ea values illustrate that the H2L is powerfully bonded to the Ni(II) and Cd(II) ions. The existence of negative ΔS* values in NiL2 (first, second, and third steps) and CdL2 (first step) demonstrates that the activated complexes are more ordered and the reactants are slower [27]. The positive activation enthalpy ΔH* data refer to the endothermic decomposition process. The positive values of ΔG* mean that the decomposition reaction is not spontaneous [28]. Figure 6 shows the plots of ln[1(1 − α)1 − n/(1 − n)T2] vs. 1/T for different models of Ni(II) and Cd(II) chelates, respectively, demonstrating that all models exhibit a linear trend with a good correlation coefficient.

2.7. UV-Visible Absorption Study

The electronic spectra of the prepared compounds were noted in DMF (10−4 M) at room temperature. The free ligand demonstrated that absorption bands appear at 386, 445, and 468 nm due to the π→π* transition of the aromatic ring, n→π* of azomethine (HC=N), and n→π* of the phenolic group, respectively. Additionally, the band appearing at 539 nm may be due the charge-transfer (CT). The Ni(II) chelates exhibit three distinct bands at 365, 408, and 542 nm, attributed, respectively, to the transitions n→π* of azomethine (HC=N), 1A1g1B1g, and 1A1g1A2g for the square planar Ni(II) chelate [29]. The electronic spectrum of the Cd(II) chelate demonstrated bands due to n→π* transition of azomethine (HC=N), which appeared at 456 nm. The second band is observed at 525 nm, which may be due to the CT transition [18,30]. The electronic spectrum of green-synthesized Ni and Cd nanocomplexes displayed a band related to the surface plasmon resonance (SPR) [31], proving the formation of nanocomposites. The Ni nanocomplexes exhibit bands at 363, 399, and 535 nm, attributed, respectively, to the transitions n→π* of imine (C=N), 1A1g1B1g, and 1A1g1A2g for the square planar Ni nanocomplex. The spectra of the Cd nanocomplex show absorption bands at 406 nm which can be assigned to the imine (C=N) group. Finally, there is an absorption band at 513 nm, which may refer to a charge-transfer band caused by a change in electron distribution between the metal and a ligand. The spectra of Ni(II) and Cd(II) chelates and their nanocomplexes are shown in the supporting information (Figure S4), and the electronic spectral data are shown in Table 5.

2.8. Molecular Modeling

2.8.1. Geometry Optimization of the H2L Ligand

The most appropriate structural configuration of the H2L ligand is shown in Figure 7. The charges from the natural bond orbital (NBO) analysis are (−0.693), (−0.953), (−0.924), (−0.531), (−0.528), (−0.891), (−0.427), and (2.357) for O1, O2, O3, N1, N2, N3, N4, and S, respectively. This supports the coordination of metal ions with O3 and N4 in a bidentate manner.

2.8.2. Optimization of [NiL2]·1.5H2O

The appropriate structural configuration of the [NiL2]·1.5H2O (Figure 8A) was demonstrated to support the ligand and metal ion interaction preferences. The nickel atom is four-coordinated in a square planar geometry in which the atoms O1, N1, O4, and N5 deviate from the plane by 0.008° degrees (Table 6). The bond lengths between N1-----O1 and N5-----O4 in the ligand (3.715 Å) are reduced in the [NiL2]·1.5H2O to 2.690, as a result of coordination. Also, the charges of coordinating atoms were, Ni (+0.654), O1 (−0.616), O4 (−0.616), N1 (−0.674), and N5 (−0.674).

2.8.3. Optimization of [CdL2]

The [CdL2] structural configuration (Figure 8B) was shown to support the ligand and metal ion interaction preferences. The cadmium atom is four-coordinated in a tetrahedral geometry (Table 6). The bond lengths between N1-----O1 and N5-----O4 in the ligand (3.715 Å) are reduced in the [CdL2] to 2.950 Å and 2.948 Å for N5-----O4, as a result of coordination. Also, the charges of coordinating atoms were Cd (+1.145), O1 (−0.695), O4 (−0.699), N1 (−0.684), and N5 (−0.687).

2.8.4. Molecular Orbitals (MO) and MEP Maps

The computed energies of formation, HOMO (eV), LUMO (eV), as well as other chemical parameters connected to HOMO and LUMO energies, were determined for the H2L ligand and its metal chelates (Table 7). As reported elsewhere, HOMO energy measures the electron donation character, while LUMO energy measures the electron affinity [32]. Higher stability compared to free ligand is shown by more negative metal chelate formation energy values [33]. The chelating agent’s computed (Eg) is greater than that of the Ni(II) and Cd(II) chelates, demonstrating its affinity for binding to the Ni(II) and Cd(II) ions [34]. The extraordinarily low dipole moment value of the Ni(II) complex may be caused by the two ligands canceling each other out due to their trans-orientation (Figure 9 and Figure 10). On the other hand, MEP maps (Figure 11) were created to distinguish the electrostatic potential of positive, negative, and neutral zones using the colors blue, red, and green, respectively. The MEP contains positive zones around hydrogen atoms and negative zones over electronegative atoms (oxygen and nitrogen).

2.9. Antitumor Activity

The anticancer activity of the ligand H2L and its Ni and Cd nano-chelates generated in CS/EtOH (before and after) heating them was examined against the HepG-2 human cancer cell line. Cis-platin (IC50 = 1.714 μg/mL) was used as the standard. Compounds with IC50 values of less than 5.00, 5.00–10.00, and 10.00–25.00 μg/mL are categorized as strong, moderate, and weak antitumor agents, respectively [35]. The growth of the HepG-2 cancer cells was generally inhibited by the investigated compounds in a concentration-dependent manner. Moreover, the examined nano metal chelates were more cytotoxic than the comparable free ligand. This could be attributed to the metal’s redox-active center [36]. Interestingly, the nano-sized Ni complex and Cd complex, after heating, displayed strong cytotoxicity compared with other compounds, whereas the Ni nanocomplex after heating exhibited antitumor activity (IC50 = 1.280 g/mL), which was higher than the activity of cis-platin (IC50 = 1.714 g/mL), and the Cd nanocomplex (before heating) displayed moderate antitumor activity (Figure 12 and Figure 13).

2.10. In Vitro Antimicrobial Results

The newly synthesized Schiff base ligand (H2L) and its Ni and Cd nanocomplexes (before and after) heating were investigated for their inhibitory effects on the growth of Bacillus cereus (G+ve), E. coli (G−ve), Micrococcus luteus (G+ve), Pseudomonas aeruginosa (G−ve), Serratia marcescens (G−ve), and Staphylococcus aureus (G+ve) bacteria, and Aspergillus flavus, Candida albicans, Fusarium oxysporum, Geotrichum candidum, Scopulariopsis brevicaulis, and Trichophyton rubrum fungi. The antimicrobial activity was tested using the disk diffusion method; the clear zone of inhibition around each disk was measured (in mm) and compared to the known sensitive drugs: chloramphenicol (CHL) as an antibacterial drug and clotrimazole (CLO) as an antifungal drug. The findings suggest that the nanocomplexes exhibit greater activity as compared to the H2L ligand under the same experimental conditions. This would indicate that the chelation could improve the ability of a nanocomplex to cross a cell membrane, which can be described by Tweedy’s chelation theory [37]. From the data in Table 8, the nano-range complexes after heating displayed higher influences on the tested bacteria compared with other compounds, indicating that the decrease in particle size caused by heating increases the activities. In case of antifungal activity, Cd nanocomplexes after heating showed good antifungal results against A. flavus and F. oxysporum within the inhibition zone, at 28 and 20 mm, which are greater than those of the standard clotrimazole (24 mm) and (18 mm), respectively (Figure 14 and Figure 15).
The growth of bacterial pathogens on each concentration was checked to determine the minimum concentration that inhibits the growth of the organism. It is evident from Table 9 that the MIC value for the Ni nanocomplex after heating was 0.625 mg/mL for B. cereus, E. coli (G−ve), and Serratia marcescens (G−ve). Furthermore, the Ni nanocomplex after heating demonstrates antifungal activity against the tested fungi, but Trichophyton rubrum appeared as the most sensitive fungus among all fungi involved in this study (Table 9).

2.11. Molecular Docking Studies

Molecular docking is a potent method for analyzing the biological activity of target molecules and producing distinguishing structural features for the creation of novel therapeutics. Moreover, it is a common computational technique for determining binding sites with appropriate conformations and estimating binding affinity. To predict the possible binding modes at the active pockets, all prepared compounds were docked with methionine adenosyl-transferases (PDB ID: 5A19) and COVID-19 main protease viral protein (PDB ID: 6lu7) (Figure 16 and Figure 17). The results demonstrated that the ligand’s dominant interaction force was the H-acceptor, although other binding interactions, such as the H donor, also took place in complexes. A stronger interaction between the investigated compounds and receptors reflects more negative energy. As a result, the interaction ability towards the receptor was arranged as follows: in the case of (PDB ID: 5A19), NiL2 > CdL2 > H2L, whereas in the case of (PDB ID: 6lu7), CdL2 > NiL2 > H2L (Table 10 and Table 11).

3. Materials and Methods

3.1. Materials and Equipment

In-depth information about the chemicals, tools, and processes used for the structural verification and applications can be found in Section S1.

3.2. Procedure of Schiff Base Synthesis

The tested H2L ligand was prepared by adding a solution of (2.7833 g, 10 mmol) Sulfamethazine sodium salt to a solution of 2-hydroxy-1-naphthaldehyde (1.7218 g, 10 mmol) in 20 mL EtOH. The reaction mixture was left under reflux for 4 h at 70 °C. The resulting yellow solid was then mixed with dilute hydrochloric acid (5%) to form a neutral solution. The yellow precipitate was recrystallized from EtOH and then dried in a vacuum (Scheme 1).

3.3. Procedure of Metal Complex Synthesis

The metal chelates were prepared by refluxing a mixture of the H2L ligand in 20 mL hot ethanol with a solution of NiCl2·6H2O or CdCl2·2.5H2O in a 1:1 molar ratio. The reaction mixtures were kept under reflux for 4–6 h at 70 °C, filtered, washed with ethanol, and dried at room temperature (Scheme 1).

3.4. Procedure of Nanocomplex Synthesis

Nano domain compounds of Ni(II) and Cd(II) were created in a 1:1 molar ratio by adding Coriandrum sativum (CS) media extract in 20% ethanol, as explained previously in our publications [32,38].

3.5. DFT Studies

The Schiff base ligand and its Ni(II) and Cd(II) chelates were optimized with the Gaussian 09 program [39], by implementing the LANL2DZ basis set for Ni and Cd metals, while the 6-311++G basis set was used for the rest of the atoms [40].

3.6. Cytotoxicity Assay

The antitumor property of the examined compounds was measured on a microplate reader (Sunrise, Tecan, Inc., East Lyme, CT, USA) using 490 nm filters against the Hepatocellular carcinoma cell line, HepG-2 cells (ATCC No. HB-8064), and compared to the cis-platin (cis-diamminedichloroplatinium) as reference drug. The experiments were conducted in the tissue culture section of Al-Azhar University’s Regional Institute for Mycology and Biotechnology in Cairo, Egypt. The IC50 values were calculated using the graphed prism version [41].

3.7. Anti-Pathogenic Activities

The prepared compounds were screened against Bacillus cereus (G+ve), E. coli (G−ve), Micrococcus luteus (G+ve), Pseudomonas aeruginosa (G−ve), Serratia marcescens (G−ve), and Staphylococcus aureus (G+ve) bacteria, and Aspergillus flavus, Candida albicans, Fusarium oxysporum, Geotrichum candidum, Scopulariopsis brevicaulis, and Trichophyton rubrum fungi [42].

3.8. Molecular Docking (MD)

Molecular docking studies were performed using the MOE2019 software. The output file of the Gaussian 09 software was used to convert the structures of the H2L ligand and its complexes into a PDB format file. The crystal structures of the target proteins for (PDB ID: 5A19) in liver cancer [43] and COVID-19 main protease (PDB ID: 6lu7) [44] were downloaded from the Protein Data Bank (https://www.rcsb.org accessed on 1 January 2020). For protein preparation, the bounded water molecules, the co-crystallized ligand, and the other cofactors were removed. In order to assign the charges and the parameters, an MMFF94x force field was used. The MOE site finder was used to produce the most probable protein binding sites. The energy of the protein molecules and the prepared compounds was reduced utilizing the energy minimization algorithm of the Molecular Operating Environment (MOE2019 software). The binding affinity between the protein and the generated compounds was determined through analyzing the binding energies of the top ten poses. By comparing the values of the binding-free energy and hydrogen bond lengths, the best scoring binding affinity could be determined. The estimated hydrogen bond lengths should not exceed 3.5 Å.

4. Conclusions

In this study, two types of nano-sized complexes have been synthesized that are characterized by various physicochemical and spectral analyses. The ligand H2L acts as a bidentate chelate by coordinating to the metal ions through NO donor atoms, forming square planar and tetrahedral complexes. Furthermore, the prepared complexes’ kinetic and thermodynamic properties are evaluated. The particles of the complexes under investigation were sub-nanoscale, as revealed by the XRD and TEM results. The findings reveal a decrease in the size of nano metal chelates after heating. The DFT calculations of the prepared compounds confirm the experimental results. The antimicrobial activities of the ligand, H2L, and its nano metal chelates showed that the metal ion in the nanocomposites enhanced the antimicrobial activities in comparison to the free ligand. Also, all compounds under examination demonstrated a reduction in HEPG-2 cell growth, and heating the nanocomplexes increased their cytotoxic effects. The Ni nanocomplex exhibited substantial antitumor activity (IC50 = 1.280 g/mL), which was higher than the activity of cis-platin (IC50 = 1.714 g/mL). Additionally, a molecular docking study suggests that [NiL2]·1.5H2O and CdL2 have the most activity with (PDB ID: 5A19) and (PDB ID: 6lu7) cell-proteins, respectively. It is worth mentioning that there is a limitation for using the investigated compounds as drugs due to their lack of stability. Also, there is a need for further investigation of the toxicity of the prepared complexes towards normal cancer cells.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241814259/s1.

Author Contributions

Conceptualization, S.H., R.F.A.E.-B., A.M.A.-D. and G.A.G.; methodology, S.H., R.F.A.E.-B., Z.H.A.E.-W., G.A.G. and A.M.A.-D.; software S.H., R.F.A.E.-B. and A.M.A.-D.; validation, S.H., R.F.A.E.-B., Z.H.A.E.-W., G.A.G. and A.M.A.-D.; formal analysis, S.H., R.F.A.E.-B., Z.H.A.E.-W., G.A.G. and A.M.A.-D.; investigation, S.H., R.F.A.E.-B., Z.H.A.E.-W., G.A.G., A.M.A.-D. and A.A.; resources, S.H., R.F.A.E.-B., A.A., G.A.G., M.S.S. and A.M.A.-D.; data curation, S.H., R.F.A.E.-B., A.M.A.-D., M.S.S. and G.A.G.; writing—original draft preparation, S.H., R.F.A.E.-B., A.M.A.-D. and G.A.G.; writing—review and editing, S.H., R.F.A.E.-B., A.M.A.-D. and G.A.G.; visualization, S.H., R.F.A.E.-B., A.A., G.A.G. and A.M.A.-D.; supervision, S.H., R.F.A.E.-B., A.M.A.-D. and G.A.G.; project administration, S.H., R.F.A.E.-B., A.M.A.-D. and G.A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by [Ministry of Education in Saudi Arabia] with grant number [445-9-198].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw/processed data generated in this work are available upon request from the corresponding author.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research and Innovation, Ministry of Education in Saudi Arabia, for funding this research workthrough the project number 445-9-198.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brodowska, K.; Lodyga-Chruscinska, E. Schiff bases—Interesting range of applications in various fields of science. ChemInform 2015, 46, 129–134. [Google Scholar] [CrossRef]
  2. Xavier, A.; Srividhya, N. Synthesis and study of Schiff base ligands. IOSR J. Appl. Chem. 2014, 7, 6–15. [Google Scholar] [CrossRef]
  3. Sharma, K.; Agarwal, S.; Gupta, S. Antifungal, antibacterial and antifertility activities of biologically active macrocyclic complexes of Tin (II). Int. J. Chem. Tech. Res. 2013, 5, 456–463. [Google Scholar]
  4. Ramírez-Jiménez, A.; Gómez, E.; Hernández, S. Penta-and heptacoordinated tin (IV) compounds derived from pyridine Schiff bases and 2-pyridine carboxylate: Synthesis and structural characterization. J. Organomet. Chem. 2009, 694, 2965–2975. [Google Scholar] [CrossRef]
  5. Yun, M.-K.; Wu, Y.; Li, Z.; Zhao, Y.; Waddell, M.B.; Ferreira, A.M.; Lee, R.E.; Bashford, D.; White, S.W. Catalysis and sulfa drug resistance in dihydropteroate synthase. Science 2012, 335, 1110–1114. [Google Scholar] [CrossRef] [PubMed]
  6. Valarmathy, G.; Subbalakshmi, R.; Sabarika, B.; Nisha, C. Schiff bases derived from 4-amino-N-substituted benzenesulfonamide: Syn thesis, spectral characterisation and MIC evaluation. Bull. Chem. Soc. Ethiop. 2021, 35, 435–448. [Google Scholar] [CrossRef]
  7. de Zayas-Blanco, F.; García-Falcón, M.; Simal-Gándara, J. Determination of sulfamethazine in milk by solid phase extraction and liquid chromatographic separation with ultraviolet detection. Food Control 2004, 15, 375–378. [Google Scholar] [CrossRef]
  8. Mansour, A.M.; Ghani, N.T.A. Hydrogen-bond effect, spectroscopic and molecular structure investigation of sulfamethazine Schiff-base: Experimental and quantum chemical calculations. J. Mol. Struct. 2013, 1040, 226–237. [Google Scholar] [CrossRef]
  9. Mansour, A.M. Coordination behavior of sulfamethazine drug towards Ru (III) and Pt (II) ions: Synthesis, spectral, DFT, magnetic, electrochemical and biological activity studies. Inorg. Chim. Acta 2013, 394, 436–445. [Google Scholar] [CrossRef]
  10. Zare, N.; Zabardasti, A. A new nano-sized mononuclear Cu (II) complex with N, N-donor Schiff base ligands: Sonochemical syn-thesis, characterization, molecular modeling and biological activity. Appl. Organomet. Chem. 2019, 33, e4687. [Google Scholar] [CrossRef]
  11. Ahmadov, I.; Bandaliyeva, A.; Nasibova, A.; Hasanova, F.; Khalilov, R. The synthesis of the silver nanodrugs in the medicinal plant baikal skullcap (Scutellaria baicalensis georgi) and their antioxidant, antibacterial activity. Adv. Biol. Earth Sci. 2020, 5, 103–118. [Google Scholar]
  12. Mohammed, H.A.; Amin, M.A.; Zayed, G.; Hassan, Y.; El-Mokhtar, M.; Saddik, M.S. In vitro and in vivo synergistic wound healing and anti-methicillin-resistant Staphylococcus aureus (MRSA) evaluation of liquorice-decorated silver nanoparticles. J. Antibiot. 2023, 76, 291–300. [Google Scholar] [CrossRef] [PubMed]
  13. Nasibova, A. Generation of nanoparticles in biological systems and their application prospects. Adv. Biol. Earth Sci. 2023, 8, 140–146. [Google Scholar]
  14. Jafarova, A.; Ramazanli, V. Antibacterial characteristics of Ag nanoparticle extracted from Olive leaf. Adv. Biol. Earth Sci. 2020, 5, 218. [Google Scholar]
  15. Aljabali, A.A.; Akkam, Y.; Al Zoubi, M.S.; Al-Batayneh, K.M.; Al-Trad, B.; Abo Alrob, O.; Alkilany, A.M.; Benamara, M.; Evans, D.J. Synthesis of gold nanoparticles using leaf extract of Ziziphus zizyphus and their antimicrobial activity. Nanomaterials 2018, 8, 174. [Google Scholar] [CrossRef] [PubMed]
  16. Singh, P.; Kim, Y.-J.; Zhang, D.; Yang, D.-C. Biological synthesis of nanoparticles from plants and microorganisms. Trends Bio-Technol. 2016, 34, 588–599. [Google Scholar] [CrossRef] [PubMed]
  17. Philip, D.; Unni, C. Extracellular biosynthesis of gold and silver nanoparticles using Krishna tulsi (Ocimum sanctum) leaf. Phys. E Low-Dimens. Syst. Nanostruct. 2011, 43, 1318–1322. [Google Scholar] [CrossRef]
  18. Hosny, S.; Gouda, G.A.; Abu-El-Wafa, S.M. Novel nano copper complexes of a new Schiff Base: Green synthesis, a new series of solid Cr (II), Co (II), Cu (II), Pd (II) and Cd (II) chelates, characterization, DFT, DNA, antitumor and molecular docking studies. Appl. Organomet. Chem. 2022, 36, e6627. [Google Scholar] [CrossRef]
  19. Tailor, S.M.; Patel, U.H. Synthesis, spectroscopic characterization, antimicrobial activity and crystal structure of silver and copper complexes of sulfamethazine. J. Coord. Chem. 2015, 68, 2192–2207. [Google Scholar] [CrossRef]
  20. Fazli, M.; Akbarzadeh-T, N.; Beitollahi, H.; Dušek, M.; Eigner, V. New Schiff base ligand N-(2-hydroxy-1-naphthylidene)-2-methyl aniline and its nano-sized copper (II) complex: Synthesis, characterization, crystal structure and application as an electrochemical sensor of 2-phenylphenol in the presence of 4-chlorophenol. J. Mater. Sci. Mater. Electron. 2021, 32, 25118–25136. [Google Scholar] [CrossRef]
  21. Gaber, G.A.; Hosny, S.; Mohamed, L.Z. Experimental and theoretical studies of 2-cyano-N-(4-morpholinobenzyldine) acetohy-drazide as corrosion inhibitor for galvanized steel and 304 stainless steel in 1M H2SO4 solution. Int. J. Electrochem. Sci. 2021, 16, 211214. [Google Scholar] [CrossRef]
  22. Al-Saidi, H.M.; Gouda, G.A.; Abdel-Hakim, M.; Alsenani, N.I.; Alfarsi, A.; Mahross, M.H.; Farghaly, O.; Hosny, S. Synthesis and Characterization of Ni (II), Cu (II), Zn (II) and Azo Dye Based on 1, 10-o-Phenanthroline Binary Complexes: Corrosion Inhibition Properties and Computational Studies. Int. J. Electrochem. Sci. 2022, 17, 2. [Google Scholar] [CrossRef]
  23. Hosseinian, A.; Jabbari, S.; Rahimipour, H.R.; Mahjoub, A.R. Synthesis and characterization of nano-scale of a new azido Co (II) complex as single and nano-scale crystals: Bithiazole precursor for the preparation of Co3O4 nano-structures. J. Mol. Struct. 2012, 1028, 215–221. [Google Scholar] [CrossRef]
  24. Rana, A.; Yadav, K.; Jagadevan, S. A comprehensive review on green synthesis of nature-inspired metal nanoparticles: Mechanism, application and toxicity. J. Clean. Prod. 2020, 272, 122880. [Google Scholar] [CrossRef]
  25. Coats, A.W.; Redfern, J. Kinetic parameters from thermogravimetric data. Nature 1964, 201, 68–69. [Google Scholar] [CrossRef]
  26. Harris, G. Kinetics and Mechanism. By AA Frost and RG Pearson. Inorg. Chem. 1963, 2, 432. [Google Scholar] [CrossRef]
  27. Yusuff, K.M.; Sreekala, R. Thermal and spectral studies of 1-benzyl-2-phenylbenzimidazole complexes of cobalt (II). Thermochim. Acta 1990, 159, 357–368. [Google Scholar] [CrossRef]
  28. Amindzhanov, A.; Manonov, K.; Kabirov, N.; Abdelrahman, G.A.H. Copper (II) complexation with 1-methyl-2-mercaptoimidazole in 7 M HCl. Russ. J. Inorg. Chem. 2016, 61, 81–85. [Google Scholar] [CrossRef]
  29. Kumari, K.; Kumar, S.; Singh, K.N.; Drew, M.G.; Singh, N. Synthesis and characterization of new square planar heteroleptic cationic complexes [Ni (ii) β-oxodithioester-dppe]+; their use as a catalyst for Chan–Lam coupling. New J. Chem. 2020, 44, 12143–12153. [Google Scholar] [CrossRef]
  30. Golcu, A.; Tumer, M.; Demirelli, H.; Wheatley, R.A. Cd (II) and Cu (II) complexes of polydentate Schiff base ligands: Synthesis, characterization, properties and biological activity. Inorg. Chim. Acta 2005, 358, 1785–1797. [Google Scholar] [CrossRef]
  31. Ghosh, S.K.; Pal, T. Interparticle coupling effect on the surface plasmon resonance of gold nanoparticles: From theory to applications. Chem. Rev. 2007, 107, 4797–4862. [Google Scholar] [CrossRef] [PubMed]
  32. Hosny, S.; Ragab, M.S.; Abd El-Baki, R.F. Synthesis of a new sulfadimidine Schiff base and their nano complexes as potential anti-COVID-19 and anti-cancer activity. Sci. Rep. 2023, 13, 1502. [Google Scholar] [CrossRef] [PubMed]
  33. Rozas, I.; Alkorta, I.; Elguero, J. Unusual Hydrogen Bonds: H⋯π Interactions. J. Phys. Chem. A 1997, 101, 9457–9463. [Google Scholar] [CrossRef]
  34. Yousef, T.; El-Reash, G.A.; El Morshedy, R. Quantum chemical calculations, experimental investigations and DNA studies on (E)-2-((3-hydroxynaphthalen-2-yl) methylene)-N-(pyridin-2-yl) hydrazinecarbothioamide and its Mn (II), Ni (II), Cu (II), Zn (II) and Cd (II) complexes. Polyhedron 2012, 45, 71–85. [Google Scholar] [CrossRef]
  35. Shier, W.T. Mammalian Cell Culture on $5 a Day: A Laboratory Manual of Low Cost Methods; University of the Philippines: Los Banos, CA, USA, 1991; Volume 64, pp. 9–16. [Google Scholar]
  36. Aljohani, E.T.; Shehata, M.R.; Abu-Dief, A.M. Design, synthesis, structural inspection of Pd2+, VO2+, Mn2+, and Zn2+ chelates incorporating ferrocenyl thiophenol ligand: DNA interaction and pharmaceutical studies. Appl. Organomet. Chem. 2021, 35, e6169. [Google Scholar] [CrossRef]
  37. Abu-Dief, A.M.; Alotaibi, N.H.; Al-Farraj, E.S.; Qasem, H.A.; Alzahrani, S.; Mahfouz, M.K.; Abdou, A. Fabrication, structural elucidation, theoretical, TD-DFT, vibrational calculation and molecular docking studies of some novel adenine imine chelates for biomedical applications. J. Mol. Liq. 2022, 365, 119961. [Google Scholar] [CrossRef]
  38. El-ghamry, M.A.; Nassir, K.M.; Elzawawi, F.M.; Aziz, A.A.A.; Abu-El-Wafa, S.M. Novel nanoparticle-size metal complexes derived from acyclovir. Spectroscopic characterization, thermal analysis, antitumor screening, and DNA cleavage, as well as 3D modeling, docking, and electrical conductivity studies. J. Mol. Struct. 2021, 1235, 130235. [Google Scholar] [CrossRef]
  39. Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. Gaussian 09, Revision D. 01; Gaussian, Inc.: Wallingford, CT, 2013. Google Scholar There is no corresponding record for this ref-erence.(b) Laurent, AD.; Jacquemin, D. Int. J. Quantum Chem. 2013, 113, 2019–2039. [Google Scholar]
  40. Ragab, M.S.; Shehata, M.R.; Shoukry, M.M.; Haukka, M.; Ragheb, M.A. Oxidative DNA cleavage mediated by a new unexpected [Pd (BAPP)][PdCl 4] complex (BAPP = 1, 4-bis (3-aminopropyl) piperazine): Crystal structure, DNA binding and cytotoxic behavior. RSC Adv. 2022, 12, 1871–1884. [Google Scholar] [CrossRef]
  41. Mosmann, T. Rapid colorimetric assay for cellular growth and survival: Application to proliferation and cytotoxicity assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
  42. Zaki, R.M.; Kamal El-Dean, A.M.; Radwan, S.M.; Sayed, A.S. Synthesis and Antimicrobial Activity of Novel Piperidinyl Tetra-hydrothieno [2, 3-c] isoquinolines and Related Heterocycles. ACS Omega 2019, 5, 252–264. [Google Scholar] [CrossRef]
  43. Murray, B.; Antonyuk, S.V.; Marina, A.; Lu, S.C.; Mato, J.M.; Hasnain, S.S.; Rojas, A.L. Crystallography captures catalytic steps in human methionine adenosyltransferase enzymes. Proc. Natl. Acad. Sci. USA 2016, 113, 2104–2109. [Google Scholar] [CrossRef]
  44. Jin, Z.; Du, X.; Xu, Y.; Deng, Y.; Liu, M.; Zhao, Y.; Zhang, B.; Li, X.; Zhang, L.; Peng, C. Structure of Mpro from SARS-CoV-2 and discovery of its inhibitors. Nature 2020, 582, 289–293. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mass spectra diagrams of (1) H2L ligand; (2) Ni(II) complex; and (3) Cd(II) complex.
Figure 1. Mass spectra diagrams of (1) H2L ligand; (2) Ni(II) complex; and (3) Cd(II) complex.
Ijms 24 14259 g001
Figure 2. The XRD patterns of (1) Ni and (2) Cd nanocomplexes in CS media.
Figure 2. The XRD patterns of (1) Ni and (2) Cd nanocomplexes in CS media.
Ijms 24 14259 g002
Figure 3. TEM images and histogram of particle size distribution for Ni and Cd nanocomplexes before heating (a,b) and after heating at 200 °C (c,d).
Figure 3. TEM images and histogram of particle size distribution for Ni and Cd nanocomplexes before heating (a,b) and after heating at 200 °C (c,d).
Ijms 24 14259 g003
Figure 4. TG and DTG curves of the [NiL2]·1.5H2O (a) and CdL2 (b) complexes.
Figure 4. TG and DTG curves of the [NiL2]·1.5H2O (a) and CdL2 (b) complexes.
Ijms 24 14259 g004
Figure 5. TG-DTG curves of Ni nanocomplex (a,b) and Cd nanocomplex (c,d) before and after heating, respectively.
Figure 5. TG-DTG curves of Ni nanocomplex (a,b) and Cd nanocomplex (c,d) before and after heating, respectively.
Ijms 24 14259 g005
Figure 6. Coats–Redfern curves of Ni(II) complex (a) and Cd(II) complex (b).
Figure 6. Coats–Redfern curves of Ni(II) complex (a) and Cd(II) complex (b).
Ijms 24 14259 g006
Figure 7. Geometry-optimized structure, dipole moment, and the charges on active centers of the H2L ligand.
Figure 7. Geometry-optimized structure, dipole moment, and the charges on active centers of the H2L ligand.
Ijms 24 14259 g007
Figure 8. Geometry–optimized structures and the charges on active centers of [NiL2]·1.5H2O (A) and [CdL2] (B).
Figure 8. Geometry–optimized structures and the charges on active centers of [NiL2]·1.5H2O (A) and [CdL2] (B).
Ijms 24 14259 g008
Figure 9. Surface plots of HOMO − LUMO of the H2L ligand.
Figure 9. Surface plots of HOMO − LUMO of the H2L ligand.
Ijms 24 14259 g009
Figure 10. Surface plots of HOMO-LUMO of [NiL2]·1.5H2O and [CdL2].
Figure 10. Surface plots of HOMO-LUMO of [NiL2]·1.5H2O and [CdL2].
Ijms 24 14259 g010
Figure 11. Molecular electrostatic potential (MEP) of H2L, [NiL2]·1.5H2O, and [CdL2].
Figure 11. Molecular electrostatic potential (MEP) of H2L, [NiL2]·1.5H2O, and [CdL2].
Ijms 24 14259 g011
Figure 12. The cell viability of H2L and its nano metal chelates versus cis-platin drug.
Figure 12. The cell viability of H2L and its nano metal chelates versus cis-platin drug.
Ijms 24 14259 g012
Figure 13. In vitro antitumor activity (IC50) of H2L and its nano metal chelates (before and after) heating against HepG-2 cells.
Figure 13. In vitro antitumor activity (IC50) of H2L and its nano metal chelates (before and after) heating against HepG-2 cells.
Ijms 24 14259 g013
Figure 14. Antibacterial activity of H2L and its nanocomplexes.
Figure 14. Antibacterial activity of H2L and its nanocomplexes.
Ijms 24 14259 g014
Figure 15. Antifungal activity of H2L and its nanocomplexes.
Figure 15. Antifungal activity of H2L and its nanocomplexes.
Ijms 24 14259 g015
Figure 16. Two-dimensional and three-dimensional plots of the interactions between H2L, [NiL2]·1.5H2O, and CdL2 with the (PDB ID: 5A19) protein receptor.
Figure 16. Two-dimensional and three-dimensional plots of the interactions between H2L, [NiL2]·1.5H2O, and CdL2 with the (PDB ID: 5A19) protein receptor.
Ijms 24 14259 g016
Figure 17. Two-dimensional and three-dimensional plots of the interactions between H2L, [NiL2]·1.5H2O, and CdL2 with the (PDB ID: 6lu7) protein receptor.
Figure 17. Two-dimensional and three-dimensional plots of the interactions between H2L, [NiL2]·1.5H2O, and CdL2 with the (PDB ID: 6lu7) protein receptor.
Ijms 24 14259 g017
Scheme 1. General route for the synthesis of the Schiff base ligand (H2L) and its metal complexes.
Scheme 1. General route for the synthesis of the Schiff base ligand (H2L) and its metal complexes.
Ijms 24 14259 sch001
Table 1. Analytical and physical data of the compounds studied.
Table 1. Analytical and physical data of the compounds studied.
CompoundChemical
Formula (M.Wt.)
ColorM.P. °Cm
(ohm−1 cm2 mol−1)
Analysis: Found (calc.), %
%C%H%N%S
H2LC23H20N4O3S (432.50)Yellow1850.763.69 (63.87)4.48 (4.66)12.73 (12.95)7.52 (7.41)
[NiL2]·1.5H2OC46H41N8NiO7.5S2 (948.70)Black>300658.37 (58.24)4.45 (4.36)11.63 (11.81)6.82 (6.76)
CdL2C46H38CdN8O6S2 (975.39)Reddish brown>3001256.89 (56.64)3.75 (3.93)11.33 (11.49)6.68 (6.57)
Table 2. Thermoanalytical results and the particle size of Ni and Cd nanocomplexes before heating.
Table 2. Thermoanalytical results and the particle size of Ni and Cd nanocomplexes before heating.
NanocompositeMediaTemperature
Range (°C)
Found (%)Particle Size (nm)
Ni nanocomplexC.S48–110
110–260
260–618
618–1000
12.89
20.25
29.40
16.14
21.39
17.06
12.05
10.11
Cd nanocomplex22–220
220–524
524–1000
5.68
29.51
42.70
31.96
22.53
12.90
Table 3. Thermoanalytical results and the particle size of Ni and Cd nanocomplexes after heating.
Table 3. Thermoanalytical results and the particle size of Ni and Cd nanocomplexes after heating.
NanocompositeMediaTemperature
Range (°C)
Found (%)Particle Size (nm)
Ni nanocomplexC.S29–204
204–550
550–1000
14.52
34.10
20.89
3.94
2.60
2.05
Cd nanocomplex29–176
176–566
566–1000
9.53
28.90
44.94
2.78
1.97
1.08
Table 4. The kinetic and thermodynamic data of the thermal decomposition of complexes.
Table 4. The kinetic and thermodynamic data of the thermal decomposition of complexes.
CompoundStepr2nZ, (min−1)E*S*H*G*
[NiL2]·1.5H2O10.9953136.2 × 10434.68−153.1731.7286.18
20.999612.4 × 1013157.00−13.34152.00159.15
30.995725.2 × 108137.12−85.74130.82195.72
40.982221.2 × 10521008.19741.34999.70244.38
[CdL2]10.978321.6 × 103118.2−245.43113.44255.04
20.997927.7 × 101655.8270.4449.77−1.43
30.999323.0 × 10551142.16806.71133.30274.66
r2 = correlation value of regression, n = order of the decomposition reaction, Z = pre-exponential factor in min−1; E*, ∆H*, and ∆G* are in kJ mol−1, ∆S* in JKmol−1.
Table 5. UV-visible spectral data of H2L and its micro- and nanocomplexes.
Table 5. UV-visible spectral data of H2L and its micro- and nanocomplexes.
CompoundSpectral Bands (nm)Wavenumber (cm−1)εmax
(L.mol−1 cm−1)
Band AssignmentGeometry
H2L386
445
468
539
25,906
22,471
21,367
18,552
12,610
9940
9270
1169
π→π*
n→π*
n→π*
CT
---------
[NiL2]·1.5H2O365
408
542
27,397
24,509
18,450
6491
5086
1132
n→π*
1A1g→1B1g
1A1g→1A2g
Square planar
CdL2456
525
21,929
19,047
5138
3276
n→π*
CT
Tetrahedral
Ni nanocomplex363
394
535
27,548
25,380
18,691
11,050
8202
1782
n→π*
1A1g→1B1g
1A1g→1A2g
Square planar
Cd nanocomplex406
513
24,630
19,493
11,440
6713
n→π*
CT
Tetrahedral
Table 6. Important optimized bond lengths (Å) and bond angles (o) of the prepared [NiL2]·1.5H2O and [CdL2] complexes.
Table 6. Important optimized bond lengths (Å) and bond angles (o) of the prepared [NiL2]·1.5H2O and [CdL2] complexes.
[NiL2]·1.5H2O
BondBond Length (Å)
Complex
BondBond Length (Å)
H2L Complex
Ni-O11.855N1-----O13.7152.69064
Ni-O41.855N5-----O43.7152.69061
Ni-N11.92523
Ni-N51.92523
AngleAngle (°)
Complex
AngleAngle (°)
Complex
N1-Ni-O190.71470O1-Ni-N589.28729
N1-Ni-N5179.99223O1-Ni-O4179.99581
N1-Ni-O489.28483O4-Ni-N590.71317
O1-N1-O4-N50.008 *
[CdL2]
BondBond length (Å)
Complex
BondBond length (Å)
H2L Complex
Cd-O12.153N1-----O13.7152.950
Cd-O42.154N5-----O43.7152.948
Cd-N12.27750
Cd-N52.28187
AngleAngle (°)
Complex
AngleAngle (°)
Complex
N1-Cd-O183.45095O1-Cd-N5109.65028
N1-Cd-N5138.91524O1-Cd-O4142.85925
N1-Cd-O4110.01703O4-Cd-N583.24996
* dihedral angle.
Table 7. Computed chemical parameters of the ligand and its complexes.
Table 7. Computed chemical parameters of the ligand and its complexes.
PropertyH2L[NiL2]·1.5H2O[CdL2]
Total energy E (a.u.)−1731.338−3630.46−3509.19
HOMO (eV)−5.8499−5.8847−6.0575
LUMO (eV)−1.6191−2.4558−2.6503
Eg = ELUMO − EHOMO (eV)4.23083.42893.4072
Dipole moment (Debye)6.13040.0025837.9638
Ionization potentialI = −EHOMO5.84995.88476.0575
Electron affinityA = −ELUMO1.61912.45582.6503
Electronegativityχ = (I +A)/23.73454.170254.3539
Chemical hardnessη = (I − A)/22.11541.71441.7036
Chemical softnessS = 1/2η0.23640.85720.8518
Chemical potentialμ = −χ−3.7345−4.1702−4.3539
Electrophilicityω = μ2/2η3.29645.07185.5636
Table 8. Antibacterial and antifungal activity (inhibition zone in mm) of chemical compounds.
Table 8. Antibacterial and antifungal activity (inhibition zone in mm) of chemical compounds.
BacteriaCompound
H2LNi Nanocomplex
B. H. *
Cd Nanocomplex
B. H. *
Ni Nanocomplex
A. H. *
Cd Nanocomplex
A. H. *
CHL
Bacillus cereus (G+ve)6.15 ± 0.1115.25 ± 0.1410.15 ± 0.0818.25 ± 0.0914.40 ± 0.0722.25 ± 0.13
E. coli (G−ve)7.20 ± 0.0816.18 ± 0.1514.25. ± 0.0520.15 ± 0.1614.50 ± 0.0520.15 ± 0.04
Micrococcus luteus (G+ve)9.10 ± 0.1216.55 ± 0.1314.60 ± 0.0318.75 ± 0.0914.75 ± 0.1120.35 ± 0.15
Pseudomonas aeruginosa (G−ve)4.15 ± 0.1512.35 ± 0.0812.35 ± 0.1214.45 ± 0.0514.45 ± 0.0818.15 ± 0.05
Serratia marcescens (G−ve)3.25 ± 0.0416.20 ± 0.0714.25 ± 0.0316.50 ± 0.1315.75 ± 0.1120.55 ± 0.14
Staphylococcus aureus (G+ve)5.25 ± 0.0616.15 ± 0.0912.10 ± 0.0416.70 ± 0.1214.30 ± 0.1318.45 ± 0.06
FungiCompound
H2LNi Nanocomplex
B. H. *
Cd Nanocomplex
B. H. *
Ni Nanocomplex
A. H. *
Cd Nanocomplex
A. H. *
CLO
Aspergillus flavus12.25 ± 0.1515.13 ± 0.0722.50 ± 0.1617.45 ± 0.1428.50 ± 0.0924.20 ± 0.14
Candida albicans7.55 ± 0.0910.60 ± 0.0513.45 ± 0.1514.50 ± 0.1816.75 ± 0.0722.35 ± 0.15
Fusarium oxysporum6.65 ± 0.128.25 ± 0.1310.55 ± 0.109.45 ± 0.1220.65 ± 0.0618.75 ± 0.16
Geotrichum candidum8.45 ± 0.1310.40 ± 0.1912.40 ± 0.1014.45 ± 0.0716.75 ± 0.0928.25 ± 0.18
Scopulariopsis brevicaulis10.30 ± 0.0711.55 ± 0.0910.50 ± 0.0513.45 ± 0.1412.40 ± 0.1022.45 ± 0.15
Trichophyton rubrum12.25 ± 0.0914.65 ± 0.1218.35 ± 0.0816.75 ± 0.0634.55 ± 0.1142.75 ± 0.09
B. H. * = before heating; A. H. * = after heating.
Table 9. Antibacterial and antifungal activity (inhibition zone and MIC) of chemical compounds (mg/mL).
Table 9. Antibacterial and antifungal activity (inhibition zone and MIC) of chemical compounds (mg/mL).
BacteriaCompound
H2LNi Nanocomplex
B. H. *
Cd Nanocomplex
B. H. *
Ni Nanocomplex
A. H. *
Cd Nanocomplex
A. H. *
CHL
Bacillus cereus (G+ve)3.15 (3.25)8.45 (1.25)8.25 (2.5)8.50 (0.625)8.15 (2.5)10.25 (0.078)
E. coli (G-ve)5.20 (4.25)10.75 (1.25)7.60 (2.5)8.25 (0.625)10.25 (2.5)10.15 (0.156)
Micrococcus luteus (G+ve)7.45 (3.75)10.25 (2.5)8.50 (5)14.45 (2.5)8.45 (1.25)10.50 (0.625)
Pseudomonas aeruginosa (G-ve)2.15 (2.50)8.50 (5)8.75 (5)10.55 (1.25)8.15 (2.5)10.75 (0.156)
Serratia marcescens (G-ve)2.25 (3.25)6.40 (1.25)8.40 (2.5)6.35 (0.625)10.30 (2.5)10.25 (0.625)
Staphylococcus aureus (G+ve)3.25 ± (4.75)6.25 (0.625)6.50 (2.5)8.15 (2.5)6.15 (1.25)10.15 (0.08)
FungiCompound
H2LNi Nanocomplex
B. H. *
Cd Nanocomplex
B. H. *
Ni Nanocomplex
A. H. *
Cd Nanocomplex
A. H. *
CLO
Aspergillus flavus8.15 (17)10.45 (12.75)18.15 (20)11.15 (12)8.45 (5)10.45 (0.039)
Candida albicans6.20 (16)8.45 (14.50)14.20 (20)9.40 (13)8.55 (2.5)8.60 (0.312)
Fusarium oxysporum4.85 (21)10.55 (20)10.75 (20)9.11 (13.40)17.35 (20)12.25 (0.312)
Geotrichum candidum5.45 (18)7.45 (3.5)7.25 (21)8.35 (2.5)9.50 (14)10.15 (0.156)
Scopulariopsis brevicaulis7.40 (19)9.50 (21)10.50 (20)16.75 (20)8.35 (5)8.75 (0.078)
Trichophyton rubrum10.25 (14)11.35 (4.5)27.24 (20)18.25 (2.5)10.15 (5)22.30 (0.025)
B. H. * = before heating; A. H. * = after heating; numbers outside parentheses reveal the inhibition zone in (mm); numbers in parentheses illustrate the MIC in (mg/mL) of the tested compounds.
Table 10. Docking interaction data for the ligand (H2L) and its complexes towards liver cancer protein (PDB ID: 5A19).
Table 10. Docking interaction data for the ligand (H2L) and its complexes towards liver cancer protein (PDB ID: 5A19).
ReceptorInteractionDistance (Å) *E (kcal/mol)
H2L
O 12NZ LYS 289H-acceptor2.82 (1.93)−26.0
[NiL2]·1.5H2O
O 52NZ LYS 307H-acceptor2.90 (1.91)−29.4
O 53NZ LYS 307H-acceptor3.06 (2.39)−3.0
O 52SG CYS 149H-donor3.48 (3.00)−2.9
6-ringCA THR 146pi-H4.00−1.3
6-ringCA LYS 307pi-H4.39−0.9
CdL2
O 20NE ARG 373H-acceptor3.65 (2.77)−0.7
O 20NH2 ARG 373H-acceptor3.51 (2.60)−2.5
O 52NZ LYS 303H-acceptor2.91 (2.00)−24.8
* The lengths of H-bonds are in brackets.
Table 11. Docking interaction data for the ligand (H2L) and its complexes towards COVID-19 main protease viral protein (PDB ID: 6lu7).
Table 11. Docking interaction data for the ligand (H2L) and its complexes towards COVID-19 main protease viral protein (PDB ID: 6lu7).
ReceptorInteractionDistance (Å) *E (kcal/mol)
H2L
O 12SD MET 165H-donor3.27 (2.07)−6.9
6-ringNE2 GLN 189pi-H4.67−1.1
6-ring5-ring HIS 41pi-pi3.98−0.1
[NiL2]·1.5H2O
O 21CB SER 46H-acceptor3.49 (2.51)−0.8
O 52SG CYS 145H-acceptor3.66 (2.36)−2.0
O 53NE2 HIS 163H-acceptor3.36 (2.61)−1.2
O 53CA MET 165H-acceptor3.25 (2.44)−0.9
6-ringCA GLU 47pi-H4.13−1.5
6-ringCE MET 49pi-H3.74−0.5
CdL2
O 20N GLN 127H-acceptor3.49 (2.52)−2.9
O 52NZ LYS 5H-acceptor2.94(1.99)−26.4
6-ringCA MET 6pi-H4.24−0.9
6-ringN ALA 7pi-H3.62−3.1
6-ringNZ LYS 137pi-cation4.42−0.5
* The lengths of H-bonds are in brackets.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hosny, S.; El-Baki, R.F.A.; El-Wahab, Z.H.A.; Gouda, G.A.; Saddik, M.S.; Aljuhani, A.; Abu-Dief, A.M. Development of Novel Nano-Sized Imine Complexes Using Coriandrum sativum Extract: Structural Elucidation, Non-Isothermal Kinetic Study, Theoretical Investigation and Pharmaceutical Applications. Int. J. Mol. Sci. 2023, 24, 14259. https://doi.org/10.3390/ijms241814259

AMA Style

Hosny S, El-Baki RFA, El-Wahab ZHA, Gouda GA, Saddik MS, Aljuhani A, Abu-Dief AM. Development of Novel Nano-Sized Imine Complexes Using Coriandrum sativum Extract: Structural Elucidation, Non-Isothermal Kinetic Study, Theoretical Investigation and Pharmaceutical Applications. International Journal of Molecular Sciences. 2023; 24(18):14259. https://doi.org/10.3390/ijms241814259

Chicago/Turabian Style

Hosny, Shimaa, Randa F. Abd El-Baki, Zeinab H. Abd El-Wahab, Gamal A. Gouda, Mohammed S. Saddik, Ateyatallah Aljuhani, and Ahmed M. Abu-Dief. 2023. "Development of Novel Nano-Sized Imine Complexes Using Coriandrum sativum Extract: Structural Elucidation, Non-Isothermal Kinetic Study, Theoretical Investigation and Pharmaceutical Applications" International Journal of Molecular Sciences 24, no. 18: 14259. https://doi.org/10.3390/ijms241814259

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop