5.2
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original Article
2025
:18;
1622024
doi:
10.25259/AJC_162_2024

Development of some 1,2,3-triazole-based complexes incorporating hydroxy and acetamide groups: Synthesis, characterization, DFT calculation, and biological screening supported by molecular docking approach

Department of Chemistry, College of Science, Taibah University, P.O. Box 344, Madinah, Saudi Arabia

* Corresponding author: E-mail address: hqasem@taibahu.edu.sa (H. Qasem)

Licence
This is an open-access article distributed under the terms of the Creative Commons Attribution-Non Commercial-Share Alike 4.0 License, which allows others to remix, transform, and build upon the work non-commercially, as long as the author is credited and the new creations are licensed under the identical terms.

Abstract

A versatile and effective ligand, 1-p-Tolyl-1H-1,2,3-triazol -4-yl-methanol (TTM) was been thoroughly characterized. Four novel coordination compounds derived from the TTM ligand were synthesized, and their structures were elucidated using a variety of spectroscopic and analytical techniques, including ultraviolet–visible (UV-Vis), infrared (IR), nuclear magnetic resonance (NMR) spectroscopy (1H and 13C), and thermal analysis. In these complexes, metal coordination occurred via the nitrogen atom of the triazole ring and the terminal hydroxyl group. The thermal degradation behavior of the synthesized TTM complexes was investigated, revealing that they exhibit endothermic properties, as suggested by the calculated kinetic parameters for their various degradation stages. The structures proposed through density functional theory (DFT) calculations aligned well with the experimental findings, which confirmed the formation of complexes between the TTM and [Zn(II), Ni(II), Cu(II), and Co(II)] in a 2:1 ratio (L:M). In addition, the newly designed triazole-based complexes have an octahedral geometry, as evidenced by the correlation of the resulting analytical, chemical, and physical data. Antibacterial assays conducted against several bacterial and fungal strains demonstrated that the TTMCu complex displayed the highest activity, closely matching that of the reference standard. Additionally, the newly synthesized complexes were tested against various cell lines using the MTT assay, indicating potential anticancer properties as reflected in the percentage growth inhibition (IC50) values. Considering the IC50 values, the order of effectiveness was found as follows: TTM < TTMCo < TTMZn < TTMNi < TTMCu, across the MCF-7 (breast cancer), Hep-G2 (hepatocellular carcinoma), and HCT-116 (colon cancer) cell lines. Such activity emphasizes their potential as anticancer agents. Promising results were further supported by pharmacophore modeling and Swiss-ADME calculations. The in silico data highlighted a significant enhancement in activity upon complexation with metal ions, contrasting with the lower activity observed for the free ligand, which aligns with the experimental results. Our findings revealed that triazole-based copper complex (TTMCu) could be employed as a promising antimicrobial and anticancer drug.

Keywords

1,2,3-Triazole ligand
Antimicrobial activity
Anticancer activity
Complexion of 1,2,3-triazole
DFT
Docking

1. Introduction

Five-membered nitrogen-containing heterocycles have received a great deal of attention in recent years due to their structural diversity and possible medical applications [1]. Numerous nitrogen-based 1,2,3-triazole cores have been identified as intriguing pharmacophores capable of interacting with specific biological targets as antibacterial [2], anticancer [3-6], anti-COVID [7], antiparasitic [8], antituberculosis [9], and antiviral [10] agents. Molecular hybridization is an effective way of developing novel ligands or prototypes. It involves discovering and combining certain pharmacophoric moieties derived from the molecular framework of two or more compounds with bioactive properties [11]. Molecular hybridization was well known as a crucial approach in drug discovery since it facilitated an oversight of undesired side effects as well as the development of prospective dual-acting medications that combine the impacts of several drugs [12]. Triazole conjugates have become known as intriguing targets for novel drug discovery. The triazole appendages can be connected using a variety of synthetic approaches, including click chemistry and other coupling processes, to produce a library of diversified molecule hybrids for pharmacological screening [13]. The structural diversity of triazole conjugates allows for optimization of their pharmacological and physicochemical properties, thus enhancing their therapeutic effectiveness [14]. Recently, the synthetic importance of such 1,2,3-triazole scaffolds has gained great interest, especially since the discovery of the powerful and efficient Cu-catalyzed click alkyne-azide cycloaddition approach (CuAAC) developed by Sharpless and his team [15-17], now widely employed in drug discovery. Chemistry by clicking is a simple and effective method for linking two molecular building blocks under moderate and adjustable conditions, producing high yields with minimal purification. Click chemistry’s versatility has made it a useful tool in many domains, including analytical chemistry, materials science, chemical biology, and drug development [18,19]. In addition, the 1,2,3-triazole core has been recognized as a preferred coordinating architecture due to its remarkable structural characteristics [20]. The presence of three adjacent sp2-nitrogen atoms in their framework provides these triazole nuclei with distinct and particular electro-donating properties, potentially serving as an adaptive bridge ligand between multiple metal centers [21]. Since the first investigation in 2008 [22] on the synthesis of metal complexes based 1,2,3-triazole ligands, there has been an increasing interest in the coordination of this class of ligands with transition metals, which has revealed exciting applications in various areas [23,24]. Furthermore, metal complexes tethering triazole ligands have effectively established them as one of the most biologically active materials, having antimicrobial, anticancer, and antitubercular properties [25-29].

In fact, some metal-based drugs have been known to be more promising than their organic scaffold analogs, most likely due to their unique electrical and stereochemical properties, which might lead to new modes of action [30,31]. Therefore, the chelation of triazole-based ligands with metal ions, including copper and zinc as metallodrugs, has been found to improve their pharmacological properties, offering a platform for the development of novel therapeutic drugs [32-35]. Based on the biological significance of the 1,2,3-triazole scaffold and the chemical properties of transition metals, we anticipated the combination of these two entities in a single molecular framework as a continuation of our interest in the development of novel bioactive metal-based drugs [36-41]. Herein, we report the synthesis and spectral characterization of a new triazole-based ligand namely 1-p-Tolyl-1H-1,2,3-triazol-4-yl-methanol (TTM). The coordination of TTM with different metal salts led to the formation of four novel 1,2,3-triazole coordination complexes (TTMCo, TTMNi, TTMCu, and TTMZn). The newly designed metal complexes-based 1,2,3-triazole were characterized using various spectral and thermal techniques. Theoretical optimization of the structures of the TTM-based complexes has been investigated using DFT calculation. The Gaussian 09W software was used to do all theoretical computations under the Becke’s 3parameter (B3) along with gradient-corrected correlation of LeeeYangeParr’s (LYP) (B3LYP) model [42,43] at two different base sets: 6-311G (d, p) for C, N, O, and H atoms plus LANL2DZ for metal atoms [44,45]. We opted to use DFT for optimizing the geometry because our ligand and its complexes lack high-quality crystal data. Furthermore, alongside the experiments, the DFT-based study was expanded to look at the electronic structure, molecular quantum chemical descriptors, and infrared (IR) [46,47]. While UV-Vis spectra have been calculated using the time-dependent density functional theory (TD-DFT), the molecular geometries of the ligand and its complexes were carefully optimized at their stationary point of ground state energy levels, which is used for all other calculations with the same functional and bases sets [48,49]. Over the past 20 years, DFT has emerged as a potent method in the physical, chemical, and material sciences for examining the electronic structure of many-body systems, including their geometries, vibrational frequencies, electronic populations, energies, and other associated molecular characteristics [50]. DFT offers a basis for comprehending how the electrons in the materials behave according to their density. When there are no experimental requirements, it may be used alone [51-56]. It can also be used to forecast the qualities before doing the experiment, which can save money and time. However, it may also be used together with the experiment to validate the results obtained in the lab, enabling the prediction and validation of experimentally observable quantities. It has uses in a number of contemporary applications [57-59].

Finally, the TTM-based ligand and its metal complexes were assessed for their antimicrobial, anticancer and antioxidant activities to evaluate their biological performance. The prediction of anti-breast cancer, and anti-microbial activities of the title compounds was based on a quantum topic of view (in silico) through assessing MOE-docking studies against 3HB5, 6DRS, and 4JIT target proteins.

2. Materials and Methods

2.1. Employed instruments

The instruments employed in the current investigation have been elaborated in the supporting information.

2.2. Materials

All reagents and chemicals used in the study have been elaborated in the supporting information

2.3. Synthesis and characterization of 1-(p-tolyl)-1H-1,2,3-triazol-4-yl)methanol

A solution of 1-azido-4-methylbenzene (1) (10 mmol) and propargyl alcohol (2) (12 mmol) in dimethyl sulfoxide (DMSO) (10 mL) was added with stirring to an aqueous solution (10 mL) containing CuSO4.5H2O (0.10 g) and sodium ascorbate (0.15 g). The stirring was continued for 16 hrs and the reaction’s completion was monitored by thin layer chromatography (TLC). Then, ice cold water was added to the reaction mixture. The resulting precipitate was collected by filtration, washed with a saturated ammonium chloride solution, and recrystallized from ethanol/DMF to produce (1-(p-tolyl)-1H-1,2,3-triazol-4-yl) methanol (3) in 92% yield. Mp: 125-127 °C. IR (KBr, υ): 3215 (OH), 3076 (Ar-CH), 2926 (Al-CH), 1591 (N=N), 1519 (C=C) and 1244 cm-1 (C-O). 1H NMR: δH = 8.62 (s, 1H, 1,2,3-triazole-H5), 7.79 (2H, d, J = 8 Hz, Ar-H), 7.39 (2H, d, J = 8 Hz, Ar-H), 5.32 (1H, t, J = 4 Hz, OH), 4.62 (2H, d, J = 4 Hz, CH2O), 2.37 (3H, s, CH3). 13C NMR: δC = 149.49, 138.52, 134.96, 130.64, 121.29, 120.26 (Ar-C); 55.47 (CH2O); 20.98 (CH3). Found: C, 63.40; H, 5.90; N, 22.16; Calcd. For C, 63.48; H, 5.86; N, 22.21

2.4. Preparation of TTM complexes

Half molar ratio of acetate transition metal (Cu2+, Zn2+, Ni2+ & Co2+) was added to a solution of the synthesized (TTM) ligand (10 mmol, 1.89 g) and equivalent molar ratio of diisopropylethylamine in a 1:1 mixture of solvents (chloroform and ethanol). After heating under reflux for 2 hrs, the resulting solid was collected by filtration, then washed with cold water, followed by ethanol and dried overnight to synthesise the desired TTM complexes.

2.4.1. TTMCo complex (Diacetato Cobalt(II) (1-(4-methylphenyl)-1H-1,2,3-triazol-4-yl)methanol):

Color : Light pink; Yield = 86 %; m.p. > 300°C; IR (KBr, υ): 3215 (OH), 3074 (Ar-CH), 2926 (Al-CH), 1562 (N=N), 1519 (C=C), 1413 (asyCOO-), 1359 (syCOO-), 1242 (C-O), 541 (Co-O) and 459 cm-1 (Co-N); Analysis for [Co(TTM)(CH3COO)2].2H2O; C24H32CoN6O8, 591.48: Calculated: C, 48.73; H, 5.45; N, 14.21; Co, 9.96% Found: C, 48.49; H, 5.25; N, 14.52; Co, 9.65 %. Λ m* in DMF (ohm-1cm2 mol-1): 12.15; µeff (B.M.); 3.38.

2.4.2. TTMNi complex (Diacetato Nickel(II) (1-(4-methylphenyl)-1H-1,2,3-triazol-4-yl)methanol):

Color: Turquoise; Yield = 84 %; m.p. > 300°C; IR (KBr, υ): 3205 (OH), 3089 (Ar-CH), 2926 (Al-CH), 1566 (N=N), 1517 (C=C), 1411 (asyCOO-), 1359 (syCOO-), 1251 (C-O), 549 (Ni-O) and 443 cm-1 (Ni-N); Analysis for [Ni(TTM)(CH3COO)2]. H2O; C24H30NiN6O7, 573.22: Calculated: C, 50.29; H, 5.28; N, 14.66; Ni, 10.24 % Found: C, 49.95; H, 5.15; N, 14.82; Ni, 10.55 %. Λ m* in DMF (ohm-1cm2 mol-1): 9.45; µeff (B.M.); 2.25.

2.4.3. TTMCu complex (Diacetato Copper(II) (1-(4-methylphenyl)-1H-1,2,3-triazol-4-yl)methanol):

Color : Cyan; Yield = 81 %; m.p. > 300 °C; IR (KBr, υ): 3211 (OH), 3091 (Ar-CH), 2924 (Al-CH), 1579 (N=N), 1516 (C=C), 1388 (asyCOO-), 1334 (syCOO-), 1257 (C-O), 563 (Cu-O) and 474 cm-1 (Cu-N); Analysis for [Cu(TTM)(CH3COO)2]. 2H2O; C24H32CuN6O8, 596.09: Calculated: C, 48.36; H, 5.41; N, 14.66; Cu, 10.66 % Found: C, 48.55; H, 5.25; N, 14.85; Cu, 10.35 %. Λ m* in DMF (ohm-1cm2 mol-1): 13.15; µeff (B.M.); 1.75.

2.4.4. TTMZn complex (Diacetato Zinc(II) (1-(4-methylphenyl)-1H-1,2,3-triazol-4-yl)methanol):

Color: light beige; Yield = 80 %; m.p. > 300 °C; IR (KBr, υ): 3205 (OH), 3118 (Ar-CH), 2926 (Al-CH), 1548 (N=N), 1517 (C=C), 1440 (asyCOO-), 1381 (syCOO-), 1244 (C-O), 621 (Zn-O) and 462 cm-1 (Zn-N), Analysis for [Zn(TTM)(CH3COO)2]. H2O; C24H30ZnN6O7, 578.92: Calculated: C, 49.71; H, 5.21; N, 14.49; Zn, 11.28 % Found: C, 49.57; H, 5.05; N, 14.67; Zn, 11.43 %. Λ m* in DMF (ohm-1cm2 mol-1): 8.45; µeff (B.M.); diamagnetic.

2.5. Stoichiometric determination of the prepared complexes

The equilibrium constants of the solutions were estimated using molar ratios and continuous variation methods [60]. After stirring, the metal salts and TTM ligand solutions were allowed to reach equilibrium. The specific Λmax absorbance for each compound was recorded. For further testing, 10 mL aliquots of stock solutions were diluted into 10 volumetric flasks of varying sizes. According to the method outlined in the previous study [61]. The initial concentration of both metal ions and TTM ligands was set at 10 mM.

2.6. Estimations of the metal chelates’ apparent formation constants

The formation constants (Kf) for the newly synthesized compounds were evaluated using spectrophotometric methods, as detailed in Eq. (1) [62]. To calculate the change in free energy (ΔG) for the chelates at 25°C, the equation ΔG = -RT ln Kf was utilized. In this equation, Kf signifies the formation constant, The ideal gas constant is R, and the temperature is T in Kelvin (Eq. 1).

(1)
k f = A A M 4 C 2 ( 1 A / A m ) 3

2.7. Kinetic investigation of prepared complexes

The integral method and the Coats–Redfern technique were used to analyze the kinetics of heat breakdown and dehydration of the aforementioned metal chelates, as detailed in Eq. (2). This approach [63-65] was utilized to graphically analyze kinetic parameters such as the activation energy (E) and the frequency factor (A). Additionally, the Coats–Redfern relation was employed to derive thermodynamic parameters, including the enthalpy of activation (ΔH), entropy of activation (ΔS), and the change in free energy (ΔG) during the decomposition steps, based on thermal gravimetric analysis (TGA) data collected from the complexes.

(2)
L o g   [ L o g ( W W w ) T 2 ] = L o g   [ A R E *   ( 1 2 R T E *   ) ]   E * 2.303 R T

In this context, W represents the mass loss up to temperature T, R is the universal gas constant, and ϕ denotes the heating rate. W∞ is the total mass loss after the decomposition process. When the left side of the equation is plotted against 1/T, it yields a straight line because 1−2RT/E* ≈ 1. Using the slope and intercept of this graph, one can calculate E* and determine the Arrhenius constant (A). These equations were employed to calculate the thermodynamic parameters, including the free energy change of activation (ΔG*), enthalpy of activation (ΔH*), and entropy of activation (ΔS*) as detailed in Eqs. (3, 4).

(3)
Δ H * = E * RT

(4)
Δ S * = 2 . 3 0 3R log A h K B T , Δ G * = Δ H * T Δ S

Where (KB) and (h) are constants of Boltzmann and Plank, respectively

2.8. Spectrophotometric studies

The absorption spectra of the free TTM ligand and its metal complexes at 1 × 10-4 M and 1 × 10-3 M were obtained. The 200–700 nm wavelength range was used to scan the spectra.

2.9. DFT calculation

The entire DFT calculations in the current work have been completed under Becke’s 3parameter (B3) along with gradient-corrected correlation of LeeeYangeParr’s (LYP), B3LYP model [42,43] at two base sets; 6-311G (d, p) for C, N, O, and H atoms as well LANL2DZ [44,66] for metal atoms using Gaussian09W package [45]. This DFT model was successful in geometry optimization with the prediction of electronic structure and related properties of the compounds with accurate results and a high level of reliability [67]. In addition to geometry optimization, DFT was also adopted to study the molecular electrostatic potential (MEP) and energy of frontier molecular orbitals (FMOs) (i.e., highest occupied molecular orbital (HOMO), lowest unoccupied molecular orbital (LUMO) and energy gap). Moreover, the quantum chemical parameters like hardness ( η), electronegativity ( χ), softness ( σ), electrophilicity ( ω), electron affinity ( A), chemical potential ( CP), and ionization potential ( I) were also calculated using the same DFT model. Besides, the TD-DFT was also employed to predict UV–Vis and IR spectra for all title compounds. Consequently, all the computations in this study will be conducted using the B3LYP approach.

2.10. Molecular docking

The molecular operation environment (MOE-2015) has been employed to achieve docking simulation [68]. The crystal structure of the target protein for breast cancer therapy, (Pdb ID: 3HB5, homo sapiens) [69], target protein for antifungal compounds, (Pdb ID: 6DRS, aspergillus flavus) and target protein for antibacterial compounds, (Pdb ID: 4JIT, escherichia coli) [70], were obtained from PDB database (http://www.rcsb.org/pdb/). Before the simulation, all crystallized ligands, water molecules and ions were removed from the protein receptors to afford a pristine formation. Besides, MOE facilities have been employed to check and prepare the protein structures for precise docking results. Moreover, the active site residues of the target proteins have been allocated using the site finder tool. The structure of the studied compounds (i.e., TTM ligand and complexes) was constructed by MOE-2015 and then subjected to protonate, partial charge checkup, and energy minimization with the same software. Afterwards, the structures were saved in .mdb format. The docking simulation was executed by MOE-2015 through the triangle matcher methodology with rigid receptor refinement to predict the binding energy and interactions of the investigated compounds at the active site pocket of the target proteins. The docking results are reported in results and discussion section.

2.11. Bioactivity

2.11.1. Antimicrobial activity

The disc diffusion technique was employed to evaluate the antibacterial and antifungal activities in vitro, as described in references [71-74]. The study included bacterial strains such as M. luteus (ATCC 4698), E. coli (ATCC 25922), and S. marcescens (ATCC 13880), as well as fungal species, including G. candidum (ATCC 204307), A. flavus (ATCC 9643), and F. oxysporum (ATCC 48112). Compounds were dissolved in DMSO to create a 0.001 mol stock solution. Nutrient agar medium was prepared, cooled to 47°C, and inoculated with the test microorganisms. The medium contained 1% peptone, 0.2% beef extract, 0.4% yeast extract, 1% NaCl, and 3% agar-agar. Once solidified, 5 mm diameter holes were punched using a sterile cork borer. The TTM ligand and its metal complexes were dissolved in DMSO at a concentration of 1.0 × 10−3 M and placed into the prepared Petri plates. These plates were then incubated at 37°C for 20 hrs to allow bacterial growth. The diameter of the inhibition zones was measured in mm to assess activity. After 24 hrs at 37°C, the plates were examined for inhibition zones, and the diameters were measured again. The final results were averaged from three readings of antimicrobial activity [75]. The mean ± SD was computed to statistically analyse the data collected over these three occasions.

2.11.2. Anti-cancer activity

The MTT assay, as previously described [76,77], was conducted to assess the viability of cancer cells in response to the synthesized compounds. HCT-116 colon cancer cells, Hep-G2 liver cancer cells, and MCF-7 breast cancer cells were plated at a density of 5 × 103 cells per well in 96-well plates. After a 24-hrs incubation period, the cells were treated with different concentrations of the compounds, ranging from 0 to 10 μg/mL, for 48 hrs. Following this exposure, the medium containing the compounds was removed, and 10 μL of MTT solution (5 mg/mL in phosphate-buffered saline) was added to each well. The plates were incubated for three hrs at 37°C to allow the formation of purple formazan crystals. To dissolve these crystals, 100 μL of DMSO was added to each well, and the OD was measured at 570 nm. Five control wells were included, and each concentration was tested in triplicate, with vinblastine as the standard drug.

The absorbance readings were expressed as a percentage of cell viability, with control wells representing 100% viability, as detailed in Eq. (5). Dose-response curves were generated using Origin software, and IC₅₀ values were calculated to determine the concentration needed to inhibit 50% of cell viability, as detailed in Eq. (6). Each concentration was tested three times, and the mean values ± SD were recorded.

(5)
Survival&nbsp;fraction = O D ( t r e a t e d   c e l l s ) O D c o n t r o l   c e l l s

The percentage growth inhibition (IC50) was computed using (Eq. 6):

(6)
IC 50 ( % ) = Control OD Compound OD Control OD × 100

2.11.3. Antioxidant activity

The radical scavenging ability of the substances against 2,2-diphenyl-1-picrylhydrazyl (DPPH) was evaluated using Blois’ method [78,73]. DPPH is a stable free radical with a distinct absorption band at 517 nm due to its unpaired electron. This absorption diminishes, and decolorization occurs when the radical is neutralized by a scavenger. Different concentrations of the test complexes were prepared and diluted to 100 µL with methanol. Approximately 5 mL of a 0.1 mM DPPH methanolic solution was added to each sample and standard (Drolox), followed by vigorous shaking. A negative control was made by mixing 100 µL of methanol with 5 mL of the 0.1 mM DPPH solution. The samples were left to stand for 20 mins at 27°C, and the absorbance was measured at 517 nm against a methanol blank.

3. Results and Discussion

3.1. Identification of the prepared TTM ligand

The click synthetic protocol adopted for the synthesis of the targeted ligand-based 1,2,3-triazole scaffold has been illustrated in Scheme 1. The precursor azide 1 used in the present study was synthesized through acid-catalyzed (HCl) diazotization of p-toluidine in the presence of sodium nitrite, followed by azidolysis with sodium azide according to the reported procedure [79]. In addition, room temperature 1,3-dipolar cycloaddition ligation of propargyl alcohol 2 with p-tolyl azide building block 1, catalyzed by CuSO4 and Na ascorbate in a solvent mixture of DMSO/H2O, gave (1-(p-tolyl)-1H-1,2,3-triazol-4-yl)methanol (3) in 92% yield.

Synthesis of (1-(p-tolyl)-1H-1,2,3-triazol-4-yl)methanol (TTM) ligand and its metal complexes.
Scheme 1.
Synthesis of (1-(p-tolyl)-1H-1,2,3-triazol-4-yl)methanol (TTM) ligand and its metal complexes.

The formation of the click product 3 has been clearly evidenced by the investigation of its IR, 1H NMR and 13C NMR spectral data (Figures S1-S3). Accordingly, no sp-signals (C≡C, ≡C-H) were recorded in the alkyne area in its IR spectrum, providing strong evidence for the success of the cycloaddition reaction. Moreover, its 1H NMR spectrum also supported the formation of the 1,2,3-triazole moiety through the disappearance of the proton of the terminal alkyne (≡C-H) and the presence of a distinct singlet at δH 8.62 ppm, which is attributed to the 1,2,3-triazole proton, are observed. The presence of the hydroxy group was evidenced by the appearance of a characteristic triplet recorded at δH 5.32 ppm. The spectrum also revealed diagnostic singlet and doublet at δH 2.37 and 4.62 ppm assigned to the CH3 and CH2 protons, respectively. The carbon signals belonging to the CH3 and CH2 groups were recorded in its 13C NMR spectrum at δC 20.98 and 55.47, respectively. All Sp2-carbons of the 1,2,3-triazole and the p-tolyl rings resonated in the aromatic regions at δC 149.49-120.26 ppm.

Figure S1

Figure S2

Figure S3

1H NMR analysis of TTMZn complex (Figure S4) revealed a shift in the position of the OH signal from 5.32 to 5.35 ppm, indicating the coordination of oxygen atoms with Zn(II) ion. Additionally, the signal recorded at 1.89 was assigned to the methyl acetate protons (Table 1).

Figure S4
Table 1. IR and 1H NMR spectral data of TTM ligand and its metals chelates.
Comp. IR
1H NMR
νOH νN=N νasOAc νsOAc νM-O νM-N Triazole-H5 -OH -CH2- -OAc
TTM 3215 1591 - - - - 8.62 5.32 4.62 -
TTMZn 3205 1547 1440 1381 621 462 8.61 5.35 4.63 1.89
TTMNi 3205 1566 1411 1359 549 443 - - - -
TTMCu 3211 1579 1388 1334 563 474 - - - -
TTMCo 3215 1562 1413 1359 541 459 - - - -

3.2. Structural inspection of the prepared complexes

3.2.1. Ft-IR, of TTM ligand and its complexes

The analysis of IR spectra yields valuable information on the binding modes surrounding the metal atom by comparing the spectra of the metal chelates with the TTM ligand. The vibration wavenumbers of the most characteristic function groups of the ligand and their chelates have been presented in Table 1.

The TTM ligand exhibits a distinct band at 1591 cm-1, which is attributed to the azo bond (-N=N-) [80]. This peak is associated with lower wavenumber findings (1579–1548 cm-1), suggesting a coordinating role of the azo group [81,82]. Further analysis of the complexes reveals the presence of new peaks in the ranges of 443–474 cm-1 and 541–621 cm-1, which confirms the chelation of the metal ions to the N2 and O2 atoms of the ligand [83,84]. Additionally, the complexes show the presence of the (CH3COO) group, whose mono-dentate nature is demonstrated by the additional absorption peaks observed at 1242–1257 cm-1. The chelated TTM ligand’s υ(O–H) stretch is thought to be in the range of 3442–3419 cm-1. Moreover, the vibrations of the (OAc)υas and (OAc)υs groups have been found at 1388–1440 cm-1 and 1334–1381 cm-1, respectively, in all the investigated complexes.

3.2.2. C, H, N percent and conductivity measurement in the compounds under investigation

The expected chemical structure of the metal chelates served as the foundation for the inferences made from the elemental analysis (C, H, and N) of the complexes. The elemental analysis results indicated that the TTM ligand is coordinated to the metal ions in a 2:1 ligand-to-metal ratio. The metal chelates were found to have high melting (decomposition) points and were stable when exposed to air. This suggests the formation of thermally stable metal-ligand complexes. Furthermore, the molar conductance data for the prepared TTM metal chelates revealed that they exhibit a non-electrolyte nature. The specific molar conductance values were as follows: TTMCo (12.15), TTMNi (9.45), TTMCu (13.15), and TTMZn (8.45).

3.2.3. Electronic spectra and magnetic moment measurements for the compounds under inspection

The ligand and its complexes with Zn(II), Ni(II), Cu(II), and Co(II) were analyzed by electronic spectroscopy in DMF at room temperature. The results have been presented in Table S1 and Figure 1. The band at 260 nm in the TTM ligand is attributed to the intraligand π-π* transition. The absorption profile of the TTMCo complex shows three peaks: at 260 nm (π→π* transition), 523 nm (LMCT transition), and 580 nm (d→d transition). The TTMCu complex exhibits bands at 261 nm (π→π* transition) and 705 nm (d→d transition). In the TTMNi complex, three peaks are observed at 260 nm (π→π* transition), 390 nm (LMCT transition), and 640 nm (d→d transition). The TTMZn complex shows two peaks at 260 nm and 330 nm, corresponding to π→π* and LMCT transitions, respectively [85]. The magnetic moment values of the metal complexes were measured at ambient temperature. The magnetic moment of the Co(II) complex was measured to be 3.38 BM, suggesting an octahedral coordination environment for the Co(II) metal center [40]. The Ni(II) metal chelate had a magnetic moment value of 2.25 BM, consistent with two uncoupled electrons. The magnetic moment of the TTMCu complex was found to be 1.75 BM, which is attributed to its electronic configuration (3d9). However, higher magnetic moment values have also been observed for Cu(II) complexes due to spin-orbit coupling [86]. The TTMZn complex showed no magnetic moment, indicating the diamagnetic nature of the Zn(II) metal center.

Table S1
Molecular electronic transitions of (a) 1x10-4 M and (b) 1x10-3 M for the prepared compounds.
Figure 1.
Molecular electronic transitions of (a) 1x10-4 M and (b) 1x10-3 M for the prepared compounds.

3.2.4. Thermal analysis

Understanding thermal analysis may help one better learn about complexes’ thermal stability as well as the makeup of the solvent or water molecules that make them up. It facilitates the distinction between lattice water and coordinated water molecules and helps to put out a basic plan for the compounds’ thermal breakdown. TGA was conducted on certain metal complexes under a nitrogen atmosphere, with temperatures ranging from 40 to 1000°C and a heating rate of 10°C/min [87]. The thermal behavior of the Zn(II), Ni(II), Cu(II), and Co(II) complexes is influenced by the specific metal ion present. The thermal decomposition of these complexes occurred in multiple stages, and the observed mass losses align with the suggested formulas derived from analytical data (Table 2). For the TTMCo complex, the first stage involved a 6.12% mass loss (calculated 6.09%) between 45-130°C, attributed to the loss of 2 water molecules. The second stage showed a 19.92% mass loss (calculated 19.97%) between 135-285°C, corresponding to the loss of a C4H6O4 moiety. In the third stage, a 33.19% mass loss (calculated 33.17%) was observed, associated with the loss of a C6H8N6O2 moiety. Finally, a 30.78% mass loss (calculated 30.81%) occurred between 290-370°C, resulting from the decomposition of a C14H14 moiety, leaving a 6.96% Co residue at 520°C. The three other complexes, TTMCu, TTMNi, and TTMZn, exhibited similar thermal decomposition patterns, with the first stage involving the loss of hydrated water molecules between 45-145°C. The second stage represented the loss of the C4H6O4 moiety between 135-290°C. The subsequent stages involved the degradation of the TTM ligand, culminating in the formation of the respective metal residues.

Table 2 The thermal breakdown processes, mass loss percentages, projected lost segments, and final residue thermo-kinetic activation parameters for each decomposition phase of the produced complexes are provided.
Complexes

Temperature

oC

Fragment loss %
Weight loss %

E*

(kJmol-1)

A

( S-1)

∆H*

(k Jmol-1)

∆G*

(k Jmol-1)

∆S*

( Jmol-1K-1)

Molecular formula M. Wt. Found Calc.

TTMCo

45- 130 2H2O 36.03 6.12 (6.09)

39.50

0.009

35.76 137.19 -281.35
135- 285 C4H6O4 118.09 19.92 (19.97) 33.45 171.78 -286.40
290- 370 C6H8N6O2 196.17 33.19 (33.17) 32.55 208.08 -291.10
375- 520 C14H14 182.26 30.78 (30.81) 31.20 243.91 -295.22
Residue >520 Co 58.93 9.96 (9.97)

TTMNi

45- 135 H2O 18.02 3.17 3.14)) 41.55 0.013 36.49 137.89 -282.17
140- 290 C4H6O4 118.09 20.56 (20.60) 35.78 173.21 -289.55
295- 375 C6H8N6O2 196.17 34.24 (34.22) 33.81 209.54 -293.37
380- 540 C14H14 182.26 31.83 (31.80) 32.15 247.59 -297.40
Residue > 540 Ni 58.69 10.20 (10.24)

TTMCu

45- 137 2H2O 36.03 6.09 6.04)) 35.60 0.007 33.56 134.09 -276.19
142- 288 C4H6O4 118.09 19.78 19.81)) 31.87 168.78 -280.55
293- 385 C6H8N6O2 196.17 32.95 32.91)) 29.90 203.25 -283.25
390- 550 C14H14 182.26 30.71 (30.76) 28.13 241.97 -287.80
Residue >550 Cu 63.55 10.64 (10.66)
TTMZn 45- 145 H2O 18.02 3.14 (3.11) 38.23 146.22 -293.45
146- 300 C4H6O4 118.09 19.82 (19.75) 36.17 182.92 -295.87
301-410 C6H8N6O2 196.17 33.80 (33.83) 44.12 0.08 34.85 223.48 -300.12
411- 620 C14H14 182.26 31.48 31.43)) 32.90 272.33 -303.65
Residue >600 Zn 65.35 11.30 (11.27)

The asterisk refers to the activation state of the materials.

3.2.5. Characterization of the synthesized compounds’ stoichiometry using spectrophotometry:

The stoichiometry of the complexes formed was identified via the spectrophotometric molar ratio and Job’s methods [88]. The molar ratio graph for the prepared complexes (Figure S5) shows a linear component crossing at a molar ratio of approximately 2, suggesting that the metal complexes were formed in a 2:1 ratio (L:M). Additionally, Job’s method indicated maximum absorbance at a ligand mole fraction of 0.62. These findings, shown in Figure 2, confirmed the molar ratio results and indicated the formation of [Zn(II), Ni(II), Cu(II), and Co(II)] complexes in a 2:1 (L:M).

Figure S5
The stoichiometry of the inspected complexes in solutions.
Figure 2.
The stoichiometry of the inspected complexes in solutions.

3.2.6. The synthetic complexes’ apparent formation constants

The formation constants (Kf) and stability constants (pK) were estimated and have been presented in Table 3, with the values ranked as follows: TTMCo > TTMCu > TTMNi > TTMZn. Additionally, the Gibbs free energy (ΔG*) values calculated for the complexes were negative, indicating the spontaneity of the complexation reaction [89].

Table 3. The values of the formation constant (Kf), stability constant (pK), and Gibbs free energy (ΔG*) of the synthesized complexes at a temperature of 298 K.
Complex Type of complex Kf Log Kf ΔG * (kJmol-1)
TTMCo 1:2 8.45×107 7.93 -45.22
TTMCu 1:2 6.95×107 7.84 -44.74
TTMNi 1:2 4.85×107 7.69 -43.85
TTMZn 1:2 2.65×107 7.42 -42.35

Based on the correlation of all collected analytical, chemical, and physical data for the compounds under investigation, it can be concluded that the examined complexes have an octahedral geometry, as seen in Scheme 1.

3.2.7. pH stability compounds under investigation

The anticipated pH stability profiles of the complexes were highly consistent over the pH = 4 - 10 range, as shown in Figure 3, suggesting that the compounds were quite comparable to one another. They are safe for use in several applications because they maintain their stability throughout a large pH range.

The stability behavior of the inspected complexes in media with different pH levels.
Figure 3.
The stability behavior of the inspected complexes in media with different pH levels.

3.2.8. Structure optimization using DFT calculation

The stability of optimized structure for TTM free ligand and its complexes TTMCu, TTMCo, TTMZn and TTMNi, chelates were investigated and confirmed by executing the ‘‘stabile’’ keyword and validated by the absence of imaginary frequencies, which indicated the stationary point of the computed structure at ground state [90,91]. The optimized structures have been given in Figure 4. The optimized structure of TTM free ligand exhibited a coplanar conformation with respect to the phenyl moiety, while TTMCu, TTMCo, TTMZn, and TTMNi complexes showed octahedral conformations, as depicted in Figure 4.

3D optimized structures of (a) TTM, (b) TTMZn, (c) TTMCo, (d) TTMCu, (e) TTMNi compounds.
Figure 4.
3D optimized structures of (a) TTM, (b) TTMZn, (c) TTMCo, (d) TTMCu, (e) TTMNi compounds.

The stability and reactivity of the compounds are mainly linked to the electron acceptor and electron donor orbitals (i.e., highest occupied and lowest unoccupied), which are energetically known as EHOMO and ELUMO or referred as FMOs. Furthermore, the energy gap (EHOMO-ELOMO) is the energetic key to explore the kinetic stability and reactivity of the compounds, through shaping the values of hardness ( η), electronegativity ( χ), softness ( σ), electrophilicity ( ω), electron affinity ( A), chemical potential ( CP) and ionization potential ( I) of the investigated compounds [92]. The surface illustration of HOMO and LUMO orbitals of the title compounds are presented in Figure 5, showing +ve and -ve phases as green and red, respectively. The calculated values for FMOs energy and other parameters are listed in Table 4.

(a-e) Graphical illustration of HOMO-LUMO distributions.
Figure 5.
(a-e) Graphical illustration of HOMO-LUMO distributions.
Table 4. Quantum chemical descriptors of free ligand and its complexes.
Compounds EHOMO (eV) ELUMO (eV) ΔE (eV) I (eV) A (eV) χ (eV) CP (eV) η (eV) Σ (eV-1) Ω (eV)
TTM -6.717 -1.065 5.652 6.717 1.065 3.891 -3.891 2.826 0.354 2.678
TTMZn -7.087 -1.519 5.568 7.087 1.519 4.303 -4.303 2.784 0.359 3.326
TTMCo -3.717 -2.501 1.215 3.717 2.501 3.109 -3.109 0.608 1.646 7.953
TTMCu -3.065 -2.470 0.595 3.065 2.470 2.768 -2.768 0.297 3.362 12.877
TTMNi -6.528 -1.086 5.442 6.528 1.086 3.807 -3.807 2.721 0.368 2.664

Based on the DFT values of energy gap and other parameters, especially, hardness which points to the stability and softness which points to more reactivity of the compounds for which the studied compound values are ranging from 0.354 to 3.362 with reactivity order of TTMCu > TTMCo > TTMNi > TTMZn> TTM (see Table 4). In the literature survey, some researchers [93], computed DFT/B3LYP/6-31g softness for the ligand metallophthalocyanine (M) with its complexes, MZn, MMn, MNi, MCu and MCo in the range from 0.5522 to 1.4612. Recently, [38] reported the softness ranging values from 0.24 to 1.16 for novel divalent chelates of Co(II), Cu(II), Zn(II) and Ni(II) with 1-(4-nitrophenyl)-1H-1,2,3-triazol-4-yl)methanol ligand, respectively, under DFT/B3LYP/LANL2DZ/6-311 g(d,p). Likewise, the DFT/B3LYP-6-311G(d,p) work of Cd, Ni, Co, Zn, and Cu complexes with (E)-N’-(3,5-di-Tert-Butyl-2-Hydroxybenzylidene) Isonicotino Hydrazide ligan reported hardness values between 0.53 to 1.69 [39]. Based on the above, the highest softness and reactivity of our complexes, especially TTMCu, TTMCo, and TTMNi, are clear. This justifies the high docking scores and bioactivity behaviour of these complexes in docking results (see docking part). According to the HSAB law (hard-soft/acid–base) regulatory, the chemical reactivity with softness and hardness of the compounds is considered a vital rule in bio-activity of the compounds, like proteins and cells. Because the soft compounds are well-interacting with a biomedical system in comparison to hard ones [94].

The 3D plot of the MEP is usually linked to the electron density plotted onto the surface of iso-electron density, which can be utilized to characterize the structural and topological features of substrates in a 3D system. MEP specifies the greatest effect of the electrons or nuclei on the molecular system through coloring system fluctuating from maximum electrophilic sites (high positively charged) blue >bluish green (medium positively)> green (less positively) > white (zero potential) > yellow (less negatively) > orang (medium negatively) > red (high negatively charged) at maximum nucleophilic sites. The positive and negative potentials indicate the ability of sites to donate and accept electrons, respectively, through the interactions [92].

MEP representation of the studied molecules has been given in Figure 6. Owing to the richness of electrons on the oxygen and nitrogen atoms, the most negative sites are located around them, so they act as an attractive target for electrophilic attack to construct acceptor H-bonds. On the other side, the highest positive areas are mostly covered by the coordinated hydrogen moiety, which tends to form donor hydrogen bonds in substrate-protein interactions (see docking part). While the bluish green is widely used to paint aromatic rings. This result of DFT-MEP computation closely matches the reported data of similar compounds [36,38,39,93].

(a-e) MEP diagram of the synthesized compounds.
Figure 6.
(a-e) MEP diagram of the synthesized compounds.

The TD-DFT computations of the free ligands and its complexes are depicted in Figure S6, in which the wavelength is extending from 200 to 700 nm to include all possible transitions. The values of computed excitation energy (E), wavelength (Λmax), transitions (major contribution) and oscillator strength (f) of the investigated compounds are listed in Table 5. The free ligand (TTM) possesses Λmax of 251 nm with 2.466 eV of energy value and f of 0.1855 which ascribed to the electronic transitions HOMO>LUMO (45%) and HOMO>LUMO+1 (32%). These values are found to be harmonized with the experimental data and reported values for similar free energy ligands [79]. As expected, all complexes suffered from red shift, which ascribed to the presence of the metal atom. The Λmax for TTMZn complex is pointed at 365 nm having E of 2.607 eV, f of 0.1080, for the electronic transitions HOMO>LUMO (87%). The researchers computed in the literature with similar ligand-Zn complex values of Λmax = 420 E=2.326 eV and f = 0.116, which are in good agreement with the data [38]. In case of TTMCo the value of Λmax reached 290 nm with E of 2.142 eV and f of 0.1070, which attributed to the electronic transitions HOMO>LUMO (48%) and HOMO2 > LUMO (25%). The calculated Λmax of TTMCu complex in the current work is located at 507 nm with 2.160 eV and 0.1644 values of E and f, respectively, which majorly ascribed to the electronic transitions HOMO>LUMO (91%). There are some researchers observed Λmax for similar complexes at 530 and 674 nm [95]. TTMNi complex, has Λmax of 325 nm with E of 2,057 eV, and f of 0.1140, which is attributed to the electronic transitions HOMO> LUMO (61%), which is in the range of experimental values and reported data of similar compounds [36,38,39].

Figure S6
Table 5. Electronic absorption properties; excitation energy (E), wavelength (Λmax), oscillator strength (f) and transitions (major contribution) for the TTM and its complexes.
Compounds E (eV) Λmax (nm) f Major contribution
TTM 2.466 251 0.1855

HOMO>LUMO (45%)

HOMO> LUMO +1 (32%)

TTMZn 2.607 365 0.1080 HOMO>LUMO (87%)
TTMCo 2.142 290 0.1070

HOMO>LUMO (48%)

HOMO2>LUMO (25%)

TTMCu 2.160 507 0.1644 HOMO>LUMO (91%)
TTMNi 2,057 325 0.1140 HOMO> LUMO (61%)

The computed vibrational frequencies in cm-1 units for all studied compounds, have been given in Figure 7. Owing to the different environment between DFT simulation (gas phase) and experiment, it is common to get some variations between vibrational frequency values resulting from the former and latter methods. Generally, the DFT values of vibration modes would be higher compared to the experimental ones. The simulation of vibrational frequencies for all compounds is achieved by the same functional (DFT/B3LYP) 6-311G(d,P) and LANL2DZ.

Computed IR spectra of the synthesized compounds.
Figure 7.
Computed IR spectra of the synthesized compounds.

3.2.9. Molecular docking

The prediction of anti-breast cancer, anti-fungal, and anti-microbial activities of the title compounds was performed by docking studies against 3HB5, 6DRS, and 4JIT target proteins, respectively. These proteins were recognized as a potent target for the stated activities. Table 6 summarizes details of the best conformations for protein-compound docking results, including docking score (S) and bonding interactions. The 3D and 2D plots of the best protein-compound conformations have been depicted in Figures 8 and 9, respectively. Additionally, Figures S7 and S8, in the supplementary material, portray 2D&3D plots for the remaining protein-compound conformations.

Figure S7

Figure S8
Table 6 Docking result for the investigated compounds and original inhibitors against target proteins.
3HB5
Ligand Receptor Interaction Distance E ( kcal/mol) S ( kcal/mol)
L N6 GLY144 H-acceptor 2.53 -1.49 -5.19
6-ring VAL225 π-H 3.45 -0.14
Zn O54 GLY144 H-acceptor 2.24 -2.49 -7.13
O54 TYR155 H-acceptor 2.37 -1.17
5-ring PHE226 π-H 2.95 -1.58
Co O52 VAL188 H-acceptor 2.42 -4.52 -7.26
C45 GLY186 H-donor 2.32 -1.73
C60 VAL188 H-donor 2.17 -2.97
6-ring ILE14 π-H 2.86 -2.35
Cu C25 ASN152 H-donor 2.32 -11.25 -8.45
C33 SER222 H-donor 2.19 -3.46
O48 PRO187 H-acceptor 2.22 -2.31
O52 SER142 H-acceptor 2.29 -1.32
O54 GLY144 H-acceptor 2.62 -0.85
O54 SER142 H-acceptor 1.89 -0.15
6-ring ILE14 π-H 3.98 -1.40
5-ring PHE226 π-H 3.25 -0.27
Ni C25 ASN152 H-donor 2.57 -5.58 -7.92
O54 GLY144 H-acceptor 2.31 -2.10
C45 TYR155 π-H 3.00 -0.13
6-ring ILE14 π-H 2.94 -0.65
E2B O30 SER142 H-acceptor 2.47 -2.15 -9.28
C9 ASN152 H-donor 2.83 -3.46
O28 GLY282 H-donor 2.61 -12.71
6-ring VAL225 π-H 3.20 -1.06
6DRS
Ligand Receptor Interaction Distance E (kcal/mol) S (kcal/mol)
L O24 TYR162 H-donor 1.90 -0.74 -5.47
Zn O54 ARG80 H-acceptor 2.63 -4.96 -7.54
6-ring VAL70 π-H 3.29 -2.38
Co 5-ring LEU77 π-H 3.00 -1.19 -7.23
O24 MET41 H-donor 3.70 -0.63
O55 PHE44 H-acceptor 2.38 -1.09
Cu O54 ARG80 H-acceptor 2.29 -1.29 -7.47
O54 ALA45 H-acceptor 2.33 -2.83
O24 MET41 H-donor 2.41 -0.10
Ni O55 ARG80 H-acceptor 1.77 -6.11 -7.54
O24 MET41 H-donor 3.01 -2.83
6-ring ARG36 π-H 3.07 -1.12
H8A N17 ILE 10 H-donor 2.78 -2.10 -8.39
N18 ASP40 H-donor 2.59 -3.40
C25 PHE44 H-π 3.75 -8.72
6-ring PHE44 π-H 3.90 -1.22
6-ring VAL70 π-H 4.25 -0.83
N21 ARG80 H-acceptor 3.38 -3.80
4JIT
Ligand Receptor Interaction Distance E (kcal/mol) S (kcal/mol)
L N29 GLY94 H-acceptor 2.59 -1.20 -5.18
5-ring ARG69 π-H 2.86 -2.28
5-ring ASP92 π-H 2.94 -0.81
6-ring ASP92 π-H 3.29 -0.46
Zn O54 SER62 H-acceptor 1.94 -2.45 -7.14
C45 ASP92 H-donor 1.63 -0.05
Co O54 LYS115 H-acceptor 2.46 -1.42 -6.73
C48 ASP92 H-donor 1.83 -0.59
Cu O54 ARG69 H-acceptor 2.03 -5.91 -7.43
C60 ASP92 H-donor 2.01 -3.49
6-ring VAL91 π-H 3.20 -1.77
6-ring ASP92 π-H 3.15 -0.14
Ni O54 SER62 H-acceptor 1.73 -3.08 -7.39
6-ring SER61 π-H 2.98 -0.55
6-ring SER62 π-H 3.26 -0.93
3ZF O14 ARG69 H-acceptor 2.75 -21.50 -9.36
O15 ASP92 H-acceptor 2.68 -7.60
O15 THR93 H-acceptor 2.79 -4.90
O13 THR93 H-acceptor 2.83 -7.00
O15 GLY94 H-acceptor 2.74 -5.30
O14 THR96 H-acceptor 2.58 -7.40
6-ring TRP134 π-π 3.72 -2.10
N1 ILE135 H-donor 2.82 -11.50
O8 ILE135 H-acceptor 2.84 -9.60
3D visualization of TTMCu, TTMNi, TTMZn, and original inhibitors docked with target proteins.
Figure 8.
3D visualization of TTMCu, TTMNi, TTMZn, and original inhibitors docked with target proteins.
(a-i) 2D interaction plots of TTMCu and TTMNi complexes and original inhibitors docked at the active site pocket of the proteins showing π-H and hydrogen bond interactions.
Figure 9.
(a-i) 2D interaction plots of TTMCu and TTMNi complexes and original inhibitors docked at the active site pocket of the proteins showing π-H and hydrogen bond interactions.

The outcomes of compound-receptor modelling revealed substantial docking score values with numerous H-bond formations reinforced with pi interactions. Generally, all the synthesized chelates (TTMCo, TTMNi, TTMCu, and TTMZn complexes) exhibited higher results compared to the TTM ligand, more (-ve) S values, indicating the potent inhibitory activity of the chelates against target proteins compared to the TTM ligand.

The docking simulation of synthesized compounds against the vital target of breast cancer 3HB5 exposed the potent inhibitory activity of TTMCu complex with the highest docking score value of (-8.45 kcal/mol) associated with two hydrogen bond donors between C25 and C33 with ASN152 and SER222, respectively, at distances of 2.32 and 2.19Å, in addition, there are four hydrogen bond acceptors formed between O48 with PRO187 at distance 2.22Å and O52 with SER142 with distance of 2.22Å as well as O54 with the both of GLY144 and SER142 residues having distances of 2.62 and 1.89Å, respectively. Furthermore, this conformation is strengthened by two π-H interactions linking ILE14 and PHE226 amino acids with the centroid of phenyl and triazole rings at distances of 3.98 and 3.25Å, respectively, (see Figures 8 and 9).

The docking score of TTMNi complex recorded a value of (-7.92 kcal/mol), as the second potent inhibitor of breast cancer with one hydrogen bond donor linked C25 with ASN152 at 2.57Å length and one hydrogen bond acceptor between O54 and GLY144 at a distance of 2.31Å. Besides, the conformation also exhibited two π-H interactions between C45 with the centroid of the phenyl ring in TYR155 residues at a distance of 3.00Å and ILE14 with the centroid of the TTMNi phenyl ring having an average distance of 2.94Å, as portrayed in Figures 8 and 9.

The complex TTMCo attained a -7.26 kcal/mol value of docking score against the 3HB5 receptor with one hydrogen bond acceptor connected O52 with VAL188 residue at a distance of 2.42Å and two hydrogen bond donor linked C45 and C60 with GLY186 and VAL188 residues at average distances of 2.32 and 2.17Å, respectively. Additionally, the centroid of the TTMCo phenyl ring interacted with ILE14 amino acid through π-H interaction at a distance of 2.86 Å, as depicted in Figures S7 and S8.

On the other side, TTMZn complex and TTM ligand obtained -7.13 and -5.19 kcal/mol value of docking scores, respectively. The TTMZn complex constructed two hydrogen bond acceptors of O54 with GLY144 and TYR155, at distances of 2.24 and 2.37 Å, respectively, as well as one π-H interaction linked PHE226 with the centroid of TTMCo triazole ring at distances of 2.95Å. While the TTM ligand shaped only one hydrogen bond acceptor between N6 and GLY144 with a distance of 2.53 Å, reinforced with one π-H interaction connecting the VAL225 residue with the centroid of the TTM phenyl ring with an average distance of 3.45 Å, as shown in Figures S7 and S8. Therefore, the ranking of the chemicals’ inhibitory effectiveness against breast cancer is as follows: TTMCu > TTMNi > TTMCo > TTMZn > TTM. In the existing literature, researchers docked (E)-N’-(3,5-di-Tert-Butyl-2-Hydroxybenzylidene) Isonicotino Hydrazide ligand and its complexes with Cd, Ni, Co, Zn, and Cu with same 3HB5 protein and found binding energy values in the range of −4.43 to −6.65 kcal/mol, accompanied with H-bond numbers ranging from 1-6 per compound at distances of 3.3 to 3.5 Å, in which the ligand-Cu complex showed the best activity for inhibiting breast cancer [39]. This literature data supports our result for TTMCu and exhibits its potential in comparison.

In the case of antifungal activity (6DRS), the highest docking score belong to TTMZn and TTMNi complexes with the same value of -7,54 kcal/mol. TTMZn complex formed one hydrogen bond acceptor between O54 and ARG80 with 2.63Å bond length and one π-H interaction found between VAL70 residue and centroid of TTMZn phenyl ring at a distance of 3.29Å, as shown in Figures 8 and 9. While the TTMNi complex exhibited one hydrogen bond acceptor at distance of 1.77Å for O55-ARG80 interaction and another one donor between O24 and MET41 residue, with average distance of 3.01 Å, associated with unique π-H interaction linked ARG36 amino acid with centroid of TTMNi phenyl ring at average distance of 3.07Å, as presented in Figures 8 and 9.

Regarding the predicted antifungal activity of TTMCu and TTMCo compounds, the simulation resulted in docking score values of -7.47 and -7.23 kcal/mol for the former and latter, respectively. The O54 atom of TTMCu constructed two hydrogen bond acceptors with ARG80 and ALA45 residues at distances of 2.29 and 2.33 Å, respectively, while the O24 atom exhibited one hydrogen bond donor with MET41 amino acid at an average distance of 2.41 Å. Furthermore, O24 and O55 atoms on TTMCo formed one hydrogen bond donor and one acceptor with MET41 and PHE44 residues with bond distances of 3.70 and 2.38 Å, respectively, as well, one π-H interaction between the LEU77 residue and centroid of the TTMCo triazole ring, having distance of 3.00Å.

The activity of TTM ligand as an antifungal compound came in the last scale with a docking score value of -5.47 kcal/mol. It was associated with a unique hydrogen bond donor that appeared between O24 and the TYR162 residue at a 1.90Å bond length. Therefore, the antifungal activity of the compounds can be arranged as TTMZn = TTMNi > TTMCu > TTMCo > TTM. The literature data for the triazole-thiole ligand and its chelates showed binding energy values in the range of -3.2 to -5.9 kcal/mol [96].

Regarding the anti-bacterial activity of the investigated compounds, the docking simulation exhibited TTMCu as the most potent compound with a docking score value of -7.43 kcal/mol, with one hydrogen bond acceptor belonging to the O54-RG69 interaction, and one donor for the C60-ASP92 interaction, at distances of 2.03 and 2.01Å, respectively. Also, the centroid of one TTMCu phenyl ring shaped two π-H interactions with ASP92 and VAL91 residues at distances of 3.15 and 3.20 Å, respectively. The TTMNi with a docking score value of -7.39 kcal/mol, got the second order as anti-bacterial compound with one hydrogen bond acceptor for the interaction of O54 with SER62 at distance of 1.73Å, reinforced by two π-H interactions connected one TTMCu phenyl ring with SER61 and SER62 residues at average distances of 2.98 and 3.26Å, respectively.

Sequentially, TTMZn and TTMCo, recorded the third and fourth order in the anti-bacterial activity, with -7.14 and -6.73 kcal/mol docking score values, for which each one exhibited one hydrogen bond acceptor and one donor, as listed in Table 6. Finally, the TTM ligand with a docking score value of -5.18 kcal/mol recorded the lowest anti-bacterial activity among all docking results, with one hydrogen bond acceptor and three π-H interactions, as illustrated in Figures S7 and S8. Consequently, the activity of the compounds as anti-bacterial candidates is TTMCu > TTMNi > TTMZn > TTMCo > TTM. There are also some other docking references to literature, evaluated chelates of Co(II), Cu(II), Zn(II) and Ni(II) with 1-(4-nitrophenyl)-1H-1,2,3-triazol-4-yl) methanol as antibacterial and reported the best binding energy values of -7.93 and - 7.74 kcal/mol for ligand-Cu and ligand-Ni, respectively, with 4H-bonds for each compound having distances between from 2.75 Ǻ and 3.56 Ǻ [38].

Moreover, The docking simulations in the current study have been validated through redocking the original inhibitors, cocrystal ligands, 3-{[(9beta,14beta,16alpha,17alpha)-3,17-dihydroxyestra-1,3,5(10)-trien-16-yl]methyl}benzamide (E2B), 3-{[(3R)-7,9-diamino-3-methyl-2,3-dihydrofuro[2,3-f]quinazolin-4-yl]oxy}benzonitrile (H8A) and {2-[(3S)-3-(2-amino-6-oxo-1,6-dihydro-9H-purin-9-yl)pyrrolidin-1-yl]-2-oxoethyl}phosphonic acid (3ZF) with 3HB5, 6DRS and 4JIT, respectively. This validation work revealed a nice fitting of the original inhibitors in the same active site pocket and equivalent H-bonds construction analogous to the contacts that resulted in our TTM ligand and its complexes. Also, the neighboring residues surrounding the binding pocket of both the co-crystal ligand and compounds are found to be parallel, as portrayed in Figures 8 and 9.

Based on the docking simulation, the presence of metals in the scaffold structure enhanced the biological activity, and among all complexes, the TTMCu and TTMNi exhibited promising anti-breast cancer, anti-fungal, and anti-bacterial activities, especially TTMCu, which might be utilized to open new fields as a lead complex for discovery and developing breast cancer drug and therapy.

3.3. Biological activity

3.3.1. In-vitro antimicrobial evaluation

The TTM ligand and its complexes were evaluated for antibacterial activity using the well diffusion approach. Antibacterial activities were tested against M. luteus, E. coli, and S. marcescens, while antifungal activities were assessed against G. candidum, A. flavus, and F. oxysporum at various concentrations. Sabouraud Dextrose Agar and Nutrient Agar were used as the media for antifungal and antibacterial screenings, respectively. Fungal growth inhibition was measured in mm. Ofloxacin and Fluconazole were used as reference medications to evaluate their antibacterial and antifungal properties, respectively (Figures 10 and 11). The antimicrobial screening data confirmed the biocidal activities of all the compounds. Comparisons between the ligand and its metal complexes revealed that the complexes had superior inhibition properties. The TTMCu complex exhibited particularly strong bioactivity, with minimum inhibition concentration (MIC) values of 2.50 µg/mL against S. marcescens and 3.00 mg/mL against G. candidum, the lowest among the tested compounds (Tables 7 and 8). The increased antimicrobial activity of the ligand after complexation can be explained by Overton’s concept and chelation theory [97]. The metal complexes’ lipophilicity increases as the metal ion charge decreases, disrupting cell permeability and hindering cellular processes [98,99]. Once inside, the complex may interfere with cellular processes, leading to microbial cell death. Chelation reduces the polarity of the metal ion due to partial sharing of its positive charge with donor atoms of the ligand, increasing its lipophilicity. Moreover, the antimicrobial activity of metal ions is significantly influenced by their complex structure and charge on the metal ions. Microbial membranes are generally negatively charged due to phospholipids and lipopolysaccharides. Positively charged metal ions (e.g., Co2⁺, Ni2+, Cu2⁺, Zn2⁺) strongly interact with microbial surfaces, leading to membrane disruption, increased permeability, and leakage of cellular contents [100].

Inhibition activity of the compounds under investigation against Microccus Luteus (+ve) bacteria.
Figure 10.
Inhibition activity of the compounds under investigation against Microccus Luteus (+ve) bacteria.
Inhibition activity of the compounds under investigation against Getrichm Candidum fungi.
Figure 11.
Inhibition activity of the compounds under investigation against Getrichm Candidum fungi.
Table 7 Minimum inhibition zone (MIC, µg/mL) for antimicrobial assay of the prepared TTM ligand and its complexes.
Compounds Bacteria
Fungi
E.coli Marcescencei M. Luteus G. candidum A. flavus F. oxysporum
TTM 8.50 7.25 6.25 7.50 8.75 8.25
TTMCo 6.50 5.25 4.50 5.25 6.25 5.50
TTMNi 4.25 3.50 3.00 3.25 4.00 3.50
TTMCu 3.75 3.00 2.50 3.00 3.50 3.25
TTMZn 4.75 4.50 3.75 4.50 5.00 4.75
Ofloxacin 3.50 2.75 2.25
Fluconazole 2.50 3.00 2.75
Comp. 10 [102] 30.16 µg/mL
Comp. 13 [102] 23.71 µg/mL
Table 8 Activity index (%) results for antimicrobial test of the generated TTM ligand and its complexes.
Compounds Activity index (%)
Bacteria
Fungi
E. coli S. Marcescence M. Luteus A. Flavus G. candidum F. oxysporum
TTM 42.21 41.21 43.64 43.00 47.91 45.59
TTMCo 74.30 73.13 85.16 71.98 85.33 81.25
TTMNi 86.68 86.67 95.39 85.27 92.73 93.01
TTMCu 93.43 91.11 97.63 91.30 95.83 96.32
TTMZn 81.99 80.40 91.65 78.26 89.50 88.79

Biological screening showed that inhibition increased with concentration (Table S2) [101]. The antimicrobial abilities of the produced complexes were verified by determining the potency index, as demonstrated in Table 7. Tables S2 and Table 8 show a comparison between the antimicrobial activity of the investigated compounds, with other compounds in the literature showing promising enhancement activity [102,103].

Table S2

3.3.2. In vitro anticancer activity

The promising results from the molecular docking studies of the ligand and its Zn(II), Ni(II), Cu(II), and Co(II) complexes led the researchers to evaluate their cytotoxic potential against various cancer cell lines. Vinblastine was used as a positive control to assess the anticancer properties of the ligands and their complexes. The results showed that all the new complexes exhibited significant cytotoxic activity (Figure 12 and Table S3). The order of effectiveness based on the IC50 values was: TTM < TTMCo < TTMZn < TTMNi < TTMCu, across the MCF-7 (breast cancer), Hep-G2 (hepatocellular carcinoma), and HCT-116 (colon cancer) cell lines. Notably, the TTMCu complex demonstrated the highest cytotoxic potency against MCF-7 cells, with an IC50 of 4.25 μg/mL, which was slightly lower than the positive control, Vinblastine (IC50 = 3.12 μg/mL). This suggests that the TTMCu complex could be a promising therapeutic agent for cancer treatment. The high cytotoxic activity of the TTMCu complex may be attributed to its strong ability to bind DNA through groove and intercalation modes, as well as the potential of the Cu(II) center to generate reactive oxygen species (ROS). These mechanisms could interfere with the transcription of cancer cell DNA and induce apoptosis. The strong correlation between the complexes’ DNA-binding capacity and their cytotoxicity supports this conclusion [104].

Table S3
Cytotoxicity of the compounds under investigation against selected cancer cell lines.
Figure 12.
Cytotoxicity of the compounds under investigation against selected cancer cell lines.

3.3.3. Antioxidant activity

The DPPH radical scavenging assay serves as a prominent and precise technique for assessing the in vitro antioxidant activity of various substances or molecules [105]. This entails examining the reduction in DPPH’s molar absorptivity at a wavelength of 517 nm following the interaction with the experimental substance. The antioxidant effect on DPPH radical scavenging is due to the compounds’ ability to donate hydrogen or act as radical scavengers [106]. The interaction involving the scavenging action of an antioxidant (H-D) with DPPH can be expressed as follows in Eq. (7):

(7)
( DPPH ) + ( H D ) ( Purple ) DPPH H + ( D ) ( Yellow )

As the DPPH radical produces DPPH-H, the presence of antioxidants decreases the amount of stable free radicals in DPPH, leading to a reduction in absorbance. This decrease in absorbance indicates the hydrogen-donating potential and scavenging activity of the antioxidants or compounds [107]. Figure 13 presents the dose-response curve for the DPPH radical scavenging activity of the TTM ligand and its Zn(II), Ni(II), Cu(II), and Co(II) complexes, as compared to Trolox, a standard antioxidant compound. The results show that the Cu(II) complexes exhibit a greater activation of the TTM ligands compared to the free TTM ligands. The IC50 value, which represents the concentration required to scavenge 50% of the DPPH radicals, was observed to be 50% at the screening level (100 µg/mL) for the unbound TTM ligand. However, upon complexation with the metal ions, the IC50 values considerably decreased, ranging from 20 to 32 μg/mL. This indicates that the metal complexation enhanced the DPPH radical scavenging activity of the TTM ligand, with the Cu(II) complex showing the most potent antioxidant properties among the tested compounds.

Free radical suppression of the compounds under investigation.
Figure 13.
Free radical suppression of the compounds under investigation.

3.3.4. Structure-activity relationship (SAR)

The Structure activity relationship (SAR) analysis for the biological reactivity of Co(II), Cu(II), Ni(II), and Zn(II) complexes with 1-p-Tolyl-1H-[1,2,3]triazol-4-yl-methanol ligand considers factors such as coordination geometry, electronic configuration, ligand-metal interactions, and their influence on biological activity [12,108].

Cu-based complexes have shown significant attention in medicinal chemistry, especially with coordination with heterocyclic cores due to their versatile biological activities. This biological activity depends mainly on structural features, the nature of the metal-ligand-metal, the oxidation state of Cu and the molecular stability. Based on the obtained results, Cu-based complexes among all tested complexes exhibited potent bioactivity as anti-microbial, antioxidant and anticancer. Cu-based complexes have shown promising antimicrobial activity against bacteria and fungi. These results agreed with the reported literature that Cu-complexes can interact and damage the microorganism’s cell wall, leading to cell death. Additionally, Cu-complexes can bind to and inhibit the enzymes for microbial survival, such as those involved in DNA replication [109]. Cu-based complexes have shown promising anticancer activity against a panel of cancer cells by induction of apoptosis that can trigger both intrinsic and extrinsic pathways, or by activating the antioxidant capacity by ROS scavenging activity, or by interaction with DNA that can interfere with replication of transcription [110]. Cu-based complexes have shown promising antioxidants by scavenging free radicals, or by mimicking the antioxidant activity like catalase (CAT) and superoxide dismutase (SOD), or by chelating activity [111].

3.3.4.1. Electronic configuration and redox activity

Co(II) can undergo oxidation to Co(III), enabling participation in redox reactions relevant to enzymatic activity. Cu(II) is highly reactive due to redox cycling Cu(II)/Cu(I), which play a role in generating ROS. Zn(II) stabilizes biomolecules and interacts with nucleic acids.

3.3.4.2. Ligand interactions

The TTM ligand enables a triazole ring to act as a nitrogen donor, mimicking biological nucleophiles. Also, a hydroxymethyl group provides additional coordination via oxygen, enhancing stability. Finally, a tolyl group to add hydrophobicity, improving lipophilicity and cell membrane permeability.

These properties influence biological reactivity through enhanced DNA/RNA binding, interaction with enzyme active sites and improved cellular uptake due to lipophilic properties.

3.3.4.3. Stability and selectivity

Cu(II) forms stable complexes with the ligand, allowing enhanced biological activity. It is a selective interaction with biomolecules due to high affinity for nitrogen and oxygen donors. Co(II) has moderate stability and selectivity towards redox-sensitive biological targets.

4. Conclusions

A series of novel metal complexes was synthesized employing a triazole derivative as the coordinating ligand. The complexes were formed with Co(II), Cu(II), Ni(II), and Zn(II) metal ions. Analytical and spectral techniques were employed to elucidate the structural forms of these complexes, which were found to be predominantly octahedral. The synthesized complexes exhibited non-conducting behavior. Additionally, chelation behavior was observed in solution, and 1 M: 2 TTM ratios were found to form in solution, indicating high stability of the complexes. To validate the empirical data and verify the proposed geometries, DFT calculations were performed. Furthermore, computational screening was carried out to evaluate the drug-like properties and potential pharmaceutical applications of the complexes. This included pharmacophore-based drug-like searches and molecular docking studies. The synthesized complexes were then subjected to various biological evaluations, including antifungal, antioxidant, antibacterial, and anticancer assessments. The findings demonstrated that the Cu(II) complex had significant efficacy in inhibiting the proliferation of certain bacteria, surpassing the activity of the free ligand and approaching that of reference drugs. Additionally, the antioxidant and cytotoxic activities of the complexes were found to be promising.

CRediT authorship contribution statement

Hamza A. Qasem: Conceptualization, investigation, methodology, writing – original draft, data curation, supervision, project administration, writing – review & editing.

Declaration of competing interest

The author declares that he has no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Declaration of Generative AI and AI-assisted technologies in the writing process

The authors confirm that there was no use of artificial intelligence (AI)-assisted technology for assisting in the writing or editing of the manuscript and no images were manipulated using AI.

Supplementary data

Supplementary material to this article can be found online at https://dx.doi.org/10.25259/AJC_162_2024.

References

  1. , , , , , , , . Nitrogen-containing heterocyclic drug products approved by the FDA in 2023: Synthesis and biological activity. European Journal of Medicinal Chemistry. 2024;279:116838. https://doi.org/10.1016/j.ejmech.2024.116838
    [Google Scholar]
  2. , . Green ultrasound-assisted three-component click synthesis of novel 1H-1,2,3-triazole carrying benzothiazoles and fluorinated-1,2,4-triazole conjugates and their antimicrobial evaluation. Acta Pharmaceutica (Zagreb, Croatia). 2017;67:309-324. https://doi.org/10.1515/acph-2017-0024
    [Google Scholar]
  3. , , , , , , , . Design, synthesis, ADME prediction and pharmacological evaluation of novel benzimidazole-1,2,3-triazole-sulfonamide hybrids as antimicrobial and antiproliferative agents. Chemistry Central Journal. 2018;12:110. https://doi.org/10.1186/s13065-018-0479-1
    [Google Scholar]
  4. , , , , , , . Design, synthesis, DNA binding, modeling, anticancer studies and DFT calculations of Schiff bases tethering benzothiazole-1,2,3-triazole conjugates. Journal of Molecular Structure. 2021;1225:129148. https://doi.org/10.1016/j.molstruc.2020.129148
    [Google Scholar]
  5. , , , , , . Design, click conventional and microwave syntheses, DNA binding, docking and anticancer studies of benzotriazole-1,2,3-triazole molecular hybrids with different pharmacophores. Journal of Molecular Structure. 2021;1225:129192. https://doi.org/10.1016/j.molstruc.2020.129192
    [Google Scholar]
  6. , , , , , , , , . Design, synthesis and anticancer screening of novel benzothiazole-piperazine-1,2,3-triazole hybrids. Molecules (Basel, Switzerland). 2018;23:2788. https://doi.org/10.3390/molecules23112788
    [Google Scholar]
  7. , . Triazole, imidazole, and thiazole-based compounds as potential agents against coronavirus. Results in Chemistry. 2021;3:100132. https://doi.org/10.1016/j.rechem.2021.100132
    [Google Scholar]
  8. , , , , , , . Expanding the scope of novel 1,2,3-triazole derivatives as new antiparasitic drug candidates. RSC Medicinal Chemistry. 2022;14:122-134. https://doi.org/10.1039/d2md00324d
    [Google Scholar]
  9. , , , , , , , , , , . Novel hybrid 1,2,4- and 1,2,3-triazoles targeting mycobacterium tuberculosis enoyl acyl carrier protein reductase (InhA): Design, synthesis, and molecular docking. International Journal of Molecular Sciences. 2022;23:4706. https://doi.org/10.3390/ijms23094706
    [Google Scholar]
  10. , , , , , , . Exploring of n-phthalimide-linked 1,2,3-triazole analogues with promising -anti-SARS-coV-2 activity: Synthesis, biological screening, and molecular modelling studies. Journal of Enzyme Inhibition and Medicinal Chemistry. 2024;39:2351861. https://doi.org/10.1080/14756366.2024.2351861
    [Google Scholar]
  11. , , . Molecular hybridization: A useful tool in the design of new drug prototypes. Current Medicinal Chemistry. 2007;14:1829-1852. https://doi.org/10.2174/092986707781058805
    [Google Scholar]
  12. , , . Versatile synthetic platform for 1,2,3-triazole chemistry. ACS Omega. 2022;7:36945-36987. https://doi.org/10.1021/acsomega.2c04883
    [Google Scholar]
  13. , , , , , , . 1,2,3‐Triazole in heterocyclic compounds, endowed with biological activity, through 1,3‐Dipolar cycloadditions. European Journal of Organic Chemistry. 2014;2014:3289-3306. https://doi.org/10.1002/ejoc.201301695
    [Google Scholar]
  14. , , , . A review on indole-triazole molecular hybrids as a leading edge in drug discovery: Current landscape and future perspectives. Current Topics in Medicinal Chemistry. 2024;24:1557-1588. https://doi.org/10.2174/0115680266307132240509065351
    [Google Scholar]
  15. , , . Click chemistry, a powerful tool for pharmaceutical sciences. Pharmaceutical Research. 2008;25:2216-2230. https://doi.org/10.1007/s11095-008-9616-1
    [Google Scholar]
  16. , , . An overview of recent advances in biomedical applications of click chemistry. Bioconjugate Chemistry. 2021;32:1455-1471. https://doi.org/10.1021/acs.bioconjchem.1c00247
    [Google Scholar]
  17. , , , , , . Click chemistry and drug delivery: A bird’s-eye view. Acta Pharmaceutica Sinica B. 2023;13:1990-2016. https://doi.org/10.1016/j.apsb.2022.10.015
    [Google Scholar]
  18. , . Recent developments in 1,2,3-triazole-based chemosensors. Dyes and Pigments. 2021;185:108905. https://doi.org/10.1016/j.dyepig.2020.108905
    [Google Scholar]
  19. , , , , , , , . Triazole: A new perspective in medicinal chemistry and material science. Current Organic Chemistry. 2022;26:1691-1702. https://doi.org/10.2174/1385272827666221213145147
    [Google Scholar]
  20. , , . Transition metal complexes of click-derived 1,2,3-triazoles as catalysts in various transformations: An overview and recent developments. Coordination Chemistry Reviews. 2023;493:215317. https://doi.org/10.1016/j.ccr.2023.215317
    [Google Scholar]
  21. , , , . Electronic properties of triazoles Experimental and computational determination of carbocation and radical-stabilizing properties. The Journal of Organic Chemistry. 2017;82:5720-5730. https://doi.org/10.1021/acs.joc.7b00548
    [Google Scholar]
  22. , , . 1,2,3-triazolylidenes as versatile abnormal carbene ligands for late transition metals. Journal of the American Chemical Society. 2008;130:13534-13535. https://doi.org/10.1021/ja805781s
    [Google Scholar]
  23. , . Application of first-row transition metal complexes bearing 1,2,3-triazolylidene ligands in catalysis and beyond. Chemical Society Reviews. 2024;53:6322-6344. https://doi.org/10.1039/d4cs00021h
    [Google Scholar]
  24. , , , , . A review on the efficacy and medicinal applications of metal-based triazole derivatives. Journal of Coordination Chemistry. 2020;73:2838-2877. https://doi.org/10.1080/00958972.2020.1839751
    [Google Scholar]
  25. , , , , , , . 1-(2-picolyl)-substituted 1,2,3-triazole as novel chelating ligand for the preparation of ruthenium complexes with potential anticancer activity. Dalton transactions (Cambridge, England : 2003). 2011;40:5188-5199. https://doi.org/10.1039/c0dt01807d
    [Google Scholar]
  26. , , . Synthesis, structure, antiproliferative activity and molecular docking of divalent and trivalent metal complexes of 4H‐3,5‐diamino‐1,2,4‐triazole and α‐hydroxynaphthaldehyde Schiff base ligand. Applied Organometallic Chemistry. 2018;32 https://doi.org/10.1002/aoc.4583
    [Google Scholar]
  27. , , , . Synthesis, spectroscopic, thermal and molecular modeling studies of Zn2+, Cd2+ and UO22+ complexes of Schiff bases containing triazole moiety Antimicrobial, anticancer, antioxidant and DNA binding studies. Materials Science & Engineering. C, Materials for Biological Applications. 2018;83:78-89. https://doi.org/10.1016/j.msec.2017.11.004
    [Google Scholar]
  28. , , , , , , . 1,2,3-triazole-gold(I)-triethylposphine derivatives active against resistant gram-positive pathogens. Bioorganic & Medicinal Chemistry Letters. 2021;40:127879. https://doi.org/10.1016/j.bmcl.2021.127879
    [Google Scholar]
  29. , , , , , , . 1H‐1,2,3‐triazole‐tethered uracil‐ferrocene and uracil‐ferrocenylchalcone conjugates: Synthesis and antitubercular evaluation. Chemical Biology & Drug Design. 2017;89:856-861. https://doi.org/10.1111/cbdd.12908
    [Google Scholar]
  30. , , . Ruthenium complexes as antimicrobial agents. Chemical Society Reviews. 2015;44:2529-2542. https://doi.org/10.1039/c4cs00343h
    [Google Scholar]
  31. , , . Small organometallic compounds as antibacterial agents. Dalton Transactions (Cambridge, England: 2003). 2012;41:6350-6358. https://doi.org/10.1039/c2dt12460b
    [Google Scholar]
  32. , , , , , . Recent progress in porphyrin-based materials for organic solar cells. Journal of Materials Chemistry A. 2018;6:16769-16797. https://doi.org/10.1039/c8ta06392c
    [Google Scholar]
  33. , , . N-heterocyclic carbene metal complexes in medicinal chemistry. Dalton Transactions (Cambridge, England: 2003). 2013;42:3269-3284. https://doi.org/10.1039/c2dt32617e
    [Google Scholar]
  34. , . Organometallic compounds: An opportunity for chemical biology? ChemBioChem. 2012;13:1232-1252. https://doi.org/10.1002/cbic.201200159
    [Google Scholar]
  35. , , , . A review on the biomedical efficacy of transition metal triazole compounds. Journal of Coordination Chemistry. 2022;75:293-334. https://doi.org/10.1080/00958972.2022.2059359
    [Google Scholar]
  36. , , , , . Synthesis and XRD of neutral NiL complex using unsymmetrical ONNO tetradentate Schiff base: Hirschfeld, spectral, DFT and thermal analysis. Journal of Coordination Chemistry. 2020;73:1280-1291. https://doi.org/10.1080/00958972.2020.1762870
    [Google Scholar]
  37. , , , , , , . Fabrication, structural elucidation, theoretical, TD-DFT, vibrational calculation and molecular docking studies of some novel adenine imine chelates for biomedical applications. Journal of Molecular Liquids. 2022;365:119961. https://doi.org/10.1016/j.molliq.2022.119961
    [Google Scholar]
  38. , , , , , , . Innovation of some novel complexes based on 1‐(4‐nitrophenyl)‐1H‐1,2,3‐triazol‐4‐yl)methanol ligand: Synthesis, structural elucidation DFT calculation and pharmaceutical studies. Applied Organometallic Chemistry. 2024;38 https://doi.org/10.1002/aoc.7486
    [Google Scholar]
  39. , , , , , , , , . Synthesis, structural determination, DFT calculation and biological evaluation supported by molecular docking approach of some new complexes incorporating (E)‐N’‐(3,5‐di‐Tert‐Butyl‐2‐Hydroxybenzylidene) isonicotino hydrazide ligand. Applied Organometallic Chemistry. 2025;39 https://doi.org/10.1002/aoc.7768
    [Google Scholar]
  40. , , , , , , . Tailoring of some novel bis-hydrazone metal chelates, spectral based characterization and DFT calculations for pharmaceutical applications and in-silico treatments for verification. Journal of Molecular Structure. 2022;1264:133263. https://doi.org/10.1016/j.molstruc.2022.133263
    [Google Scholar]
  41. , , , , , , . Development of tripodal imine metal chelates: Synthesis, physicochemical inspection, theoretical studies and biomedical evaluation. Inorganic Chemistry Communications. 2024;162:112248. https://doi.org/10.1016/j.inoche.2024.112248
    [Google Scholar]
  42. , . Inhomogeneous electron gas. Physical Review. 1964;136:B864-B871. https://doi.org/10.1103/physrev.136.b864
    [Google Scholar]
  43. , , . Development of the colle-salvetti correlation-energy formula into a functional of the electron density. Physical Review. B, Condensed matter. 1988;37:785-789. https://doi.org/10.1103/physrevb.37.785
    [Google Scholar]
  44. , . Ab initio effective core potentials for molecular calculations potentials for the transition metal atoms Sc to Hg. The Journal of Chemical Physics. 1985;82:270-283. https://doi.org/10.1063/1.448799
    [Google Scholar]
  45. . Gaussian 09, Revision D. 01/Gaussian.
  46. , , , , . Theoretical investigation for the designing of novel antioxidants. Canadian Journal of Chemistry. 2013;91:126-130. https://doi.org/10.1139/cjc-2012-0356
    [Google Scholar]
  47. , , , , , , , , . Computational study of 2N-atom functionalized corannulene by alkali metals doping: Towards the development of highly efficient nonlinear optical materials. Physica B: Condensed Matter. 2022;640:414041. https://doi.org/10.1016/j.physb.2022.414041
    [Google Scholar]
  48. , , , , , . Unveiling peripheral symmetric acceptors coupling with tetrathienylbenzene core to promote electron transfer dynamics in organic photovoltaics. Scientific Reports. 2024;14:21176. https://doi.org/10.1038/s41598-024-71777-6
    [Google Scholar]
  49. , , , . Theoretical investigation for exploring the antioxidant potential of chlorogenic acid: A density functional theory study. International Journal of Food Properties. 2016;19:745-751. https://doi.org/10.1080/10942912.2015.1042588
    [Google Scholar]
  50. , , , , , . Metal based bioactive nitrogen and oxygen donor mono and bis Schiff bases: Design, synthesis, spectral characterization, computational analysis and antibacterial screening. Acta Chimica Slovenica. 2022;69:200-216. https://doi.org/10.17344/acsi.2021.7182
    [Google Scholar]
  51. , , , , , , , , . Green synthesis, SC-XRD, non-covalent interactive potential and electronic communication via DFT exploration of pyridine-based hydrazone. Crystals. 2020;10:778. https://doi.org/10.3390/cryst10090778
    [Google Scholar]
  52. , , , , , . Photovoltaic response promoted via intramolecular charge transfer in triphenylpyridine core with small acceptors: A DFT/TD-DFT study. Materials Science in Semiconductor Processing. 2025;186:109086. https://doi.org/10.1016/j.mssp.2024.109086
    [Google Scholar]
  53. , , , , , , , , , , , . Influence of acceptor tethering on the performance of nonlinear optical properties for pyrene-based materials with a-π-d-π-D architecture. Arabian Journal of Chemistry. 2022;15:103673. https://doi.org/10.1016/j.arabjc.2021.103673
    [Google Scholar]
  54. , , , , , . A novel thiazole based acceptor for fullerene-free organic solar cells. Dyes and Pigments. 2018;149:470-474. https://doi.org/10.1016/j.dyepig.2017.10.037
    [Google Scholar]
  55. , , , , , , , . A facile microwave assisted synthesis and structure elucidation of (3R)-3-alkyl-4,1-benzoxazepine-2,5-diones by crystallographic, spectroscopic and DFT studies. Spectrochimica acta. Part A, Molecular and Biomolecular Spectroscopy. 2020;230:117995. https://doi.org/10.1016/j.saa.2019.117995
    [Google Scholar]
  56. , , , . Theoretical investigation for exploring the antioxidant potential of chlorogenic acid: A density functional theory study. International Journal of Food Properties. 2016;19:745-751. https://doi.org/10.1080/10942912.2015.1042588
    [Google Scholar]
  57. . Understanding density functional theory (DFT) and completing it in practice. AIP Advances. 2014;4 https://doi.org/10.1063/1.4903408
    [Google Scholar]
  58. , , , , , . Enriching the compositional tailoring of NLO responsive dyes with diversity oriented electron acceptors as visible light harvesters: A DFT/TD-DFT approach. Molecular Physics. 2023;121 https://doi.org/10.1080/00268976.2022.2148585
    [Google Scholar]
  59. , , , , , , , , , . Synthesis, structure elucidation, SC-XRD/DFT, molecular modelling simulations and DNA binding studies of 3,5-diphenyl-4,5-dihydro-1 h-pyrazole chalcones. Journal of Biomolecular Structure & Dynamics. 2023;43:1831-1846. https://doi.org/10.1080/07391102.2023.2293260
    [Google Scholar]
  60. , , , , , , . Design, structural inspection of some new metal chelates based on Benzothiazol-pyrimidin-2-ylidene ligand: Biomedical studies and molecular docking approach. Inorganic Chemistry Communications. 2023;158:111587. https://doi.org/10.1016/j.inoche.2023.111587
    [Google Scholar]
  61. , , , , , , , , , . Design, synthesis, and in vitro cancer cell growth inhibition evaluation and antimalarial testing of trioxanes installed in cyclic 2-enoate substructures. European Journal of Medicinal Chemistry. 2013;69:294-309. https://doi.org/10.1016/j.ejmech.2013.08.008
    [Google Scholar]
  62. , , , . Three novel Ni(II), VO(II) and Cr(III) mononuclear complexes encompassing potentially tridentate imine ligand: Synthesis, structural characterization, DNA interaction, antimicrobial evaluation and anticancer activity. Applied Organometallic Chemistry. 2017;31 https://doi.org/10.1002/aoc.3750
    [Google Scholar]
  63. , . Kinetic parameters from thermogravimetric data. Nature. 1964;201:68-69. https://doi.org/10.1038/(2010)68a0
    [Google Scholar]
  64. , , , , , , , , , . Gaussian 09 (Revision A. 02) [Computer software]. Wallingford CT: Gaussian Inc.; .
  65. , , , , , , , . Synthesis, crystal structure, DNA/bovine serum albumin binding and antitumor activity of two transition metal complexes with 4‐acylpyrazolone derivative. Applied Organometallic Chemistry. 2019;33 https://doi.org/10.1002/aoc.4668
    [Google Scholar]
  66. , , . Revised basis sets for the LANL effective core potentials. Journal of Chemical Theory and Computation. 2008;4:1029-1031. https://doi.org/10.1021/ct8000409
    [Google Scholar]
  67. , , , , , . Synthesis, molecular structure, DFT studies, in silico docking and molecular dynamics simulations of 2,6 dimethoxychalcone derivatives as BRD4 inhibitors. Journal of Molecular Structure. 2021;1245:131032. https://doi.org/10.1016/j.molstruc.2021.131032
    [Google Scholar]
  68. , , , . DOCKTITE—A highly versatile step-by-step workflow for covalent docking and virtual screening in the molecular operating environment. Journal of Chemical Information and Modeling. 2015;55:398-406. https://doi.org/10.1021/ci500681r
    [Google Scholar]
  69. , . Activities of cinnamaldehyde from Boswellia serrata on MCF-7 breast cancer cell line. Journal For Innovative Development in Pharmaceutical and Technical Science (JIDPTS). 2020;3 https://doi.org/10.26438/ijsrbs/v7i4.3543
    [Google Scholar]
  70. , , , , , , , , , , , . Inhibition of the escherichia coli 6-oxopurine phosphoribosyltransferases by nucleoside phosphonates: Potential for new antibacterial agents. Journal of Medicinal Chemistry. 2013;56:6967-6984. https://doi.org/10.1021/jm400779n
    [Google Scholar]
  71. . Selective toxicity. The physico-chemical basis of therapy (7th ed). London, U.K.: Chapman and Hall; .
  72. , . Synthesis, spectral characterisation, electrochemical, and fluorescence studies of biologically active novel Schiff base complexes derived from e-4-(2-hydroxy-3-methoxybenzlideneamino)-n-(pyrimidin-2-yl)benzenesulfonamide. Turkish Journal of Chemistry. 2014;38:521-530. https://doi.org/10.3906/kim-1301-83
    [Google Scholar]
  73. , , , , , , , , . Synthesis, structural, spectral and biological evaluation of metals endowed 1,2,4-triazole. Bulletin of the Chemical Society of Ethiopia. 2020;34:335-351. https://doi.org/10.4314/bcse.v34i2.11
    [Google Scholar]
  74. , , , , , , , . Complexes of imino-1,2,4-triazole derivative with transition metals: Synthesis and antibacterial study. Russian Journal of General Chemistry. 2018;88:1707-1711. https://doi.org/10.1134/s1070363218080248
    [Google Scholar]
  75. , , , , , , . Innovation of Fe(III), Ni(II), and Pd(II) complexes derived from benzothiazole imidazolidin‐4‐ol ligand: Geometrical elucidation, theoretical calculation, and pharmaceutical studies. Applied Organometallic Chemistry. 2023;37 https://doi.org/10.1002/aoc.7162
    [Google Scholar]
  76. . Rapid colorimetric assay for cellular growth and survival: Application to proliferation and cytotoxicity assays. Journal of Immunological Methods. 1983;65:55-63. https://doi.org/10.1016/0022-1759(83)90303-4
    [Google Scholar]
  77. , , , , , , , . Fabrication of CuO/PdO nanocomposites for biomedical applications. Inorganic Chemistry Communications. 2024;170:113166. https://doi.org/10.1016/j.inoche.2024.113166
    [Google Scholar]
  78. . Antioxidant determinations by the use of a stable free radical. Nature. 1958;181:1199-1200. https://doi.org/10.1038/1811199a0
    [Google Scholar]
  79. , , , , , , . Microwave versus conventional synthesis, anticancer, DNA binding and docking studies of some 1,2,3-triazoles carrying benzothiazole. Arabian Journal of Chemistry. 2021;14:102997. https://doi.org/10.1016/j.arabjc.2021.102997
    [Google Scholar]
  80. , , , . IR spectra and structure of some azo dyes? p-azobenzene derivatives? in various aggregate states. Journal of Applied Spectroscopy. 1988;49:1020-1024. https://doi.org/10.1007/bf00657220
    [Google Scholar]
  81. , , . Ru(III) complexes of triazole based Schiff base and azo dye ligands: An insight into the molecular structure and catalytic role in oxidative dimerization of 2-aminophenol. Inorganic Chemistry Communications. 2021;129:108616. https://doi.org/10.1016/j.inoche.2021.108616
    [Google Scholar]
  82. , , . Synthesis and spectral studies of heterocyclic azo dye complexes with some transition metals. Journal of Physics: Conference Series. 2018;1003:012021. https://doi.org/10.1088/1742-6596/1003/1/012021
    [Google Scholar]
  83. , , , . Biochemical characterization of new gemifloxacin Schiff base (GMFX-o-phdn) metal complexes and evaluation of their antimicrobial activity against some phyto- or human pathogens. International Journal of Molecular Sciences. 2022;23:2110. https://doi.org/10.3390/ijms23042110
    [Google Scholar]
  84. , , . 2,4‐Dihydroxy‐5‐[(5‐mercapto‐1H‐1,2,4‐triazole‐3‐yl)diazenyl]benzaldehyde acetato, chloro and nitrato Cu(II) complexes: Synthesis, structural characterization, DNA binding and anticancer and antimicrobial activity. Applied Organometallic Chemistry. 2019;33 https://doi.org/10.1002/aoc.4707
    [Google Scholar]
  85. , , , , , , , , . Synthesis, theoretical investigations, biocidal screening, DNA binding, in vitro cytotoxicity and molecular docking of novel Cu (II), Pd (II) and Ag (I) complexes of chlorobenzylidene Schiff base: Promising antibiotic and anticancer agents. Applied Organometallic Chemistry. 2018;32 https://doi.org/10.1002/aoc.4527
    [Google Scholar]
  86. , , , , , , . Synthesis and intensive characterization for novel Zn(II), Pd(II), Cr(III) and VO(II)-Schiff base complexes; DNA-interaction, DFT, drug-likeness and molecular docking studies. Journal of Molecular Structure. 2021;1242:130693. https://doi.org/10.1016/j.molstruc.2021.130693
    [Google Scholar]
  87. , , , , , . Development of new 2-(Benzothiazol-2-ylimino)-2,3-dihydro-1H-imidazol-4-ol complexes as a robust catalysts for synthesis of thiazole 6-carbonitrile derivatives supported by DFT studies. Journal of Molecular Structure. 2023;1292:136188. https://doi.org/10.1016/j.molstruc.2023.136188
    [Google Scholar]
  88. , , , , . Development and structure elucidation of new VO2+, Mn2+, Zn2+, and Pd2+ complexes based on azomethine ferrocenyl ligand: DNA interaction, antimicrobial, antioxidant, anticancer activities, and molecular docking. Applied Organometallic Chemistry. 2021;35 https://doi.org/10.1002/aoc.6154
    [Google Scholar]
  89. , , , , . Development of novel guanidine iron (III) complexes as a powerful catalyst for the synthesis of tetrazolo[1,5-a]pyrimidine by green protocol. Sohag Journal of Sciences. 2024;9:7-15. https://doi.org/10.21608/sjsci.2023.211358.1080
    [Google Scholar]
  90. . Thermochemistry in gaussian. Gaussian Inc; . p. :1. https://doi.org/(Ochterski,2000)
  91. . An Introduction to Computational Chemistry Using G09W and Avogadro Software. Gaussian 09W Tutorial; .
  92. , , , , , . Structural, quantum chemical and spectroscopic investigations on photophysical properties of fluorescent saccharide sensor: Theoretical and experimental studies. ChemistrySelect. 2020;5:5227-5238. https://doi.org/10.1002/slct.202000966
    [Google Scholar]
  93. , , , , , . 1,2,3-Ttriazole substituted phthalocyanine metal complexes as potential inhibitors for anticholinesterase and antidiabetic enzymes with molecular docking studies. Journal of Biomolecular Structure & Dynamics. 2022;40:4429-4439. https://doi.org/10.1080/07391102.2020.1857842
    [Google Scholar]
  94. , , , , , , . Vibrational spectroscopic studies, Fukui functions, HOMO-LUMO, NLO, NBO analysis and molecular docking study of (E)-1-(1,3-benzodioxol-5-yl)-4,4-dimethylpent-1-en-3-one, a potential precursor to bioactive agents. Journal of Molecular Structure. 2016;1123:375-383. https://doi.org/10.1016/j.molstruc.2016.07.044
    [Google Scholar]
  95. , , . Synthesis, characterization, thermal studies, electrochemical behavior, antimicrobial, docking studies, and computational simulation of triazole‐thiol metal complexes. Applied Organometallic Chemistry. 2022;36 https://doi.org/10.1002/aoc.6647
    [Google Scholar]
  96. , , , . Molecular modelling, docking of triazole-Thiole ligand and some of its chelates: Synthesis, structural characterization, thermal behaviour, and antibacterial activity. Computational Biology and Chemistry. 2018;78:260-272. https://doi.org/10.1016/j.compbiolchem.2018.12.008
    [Google Scholar]
  97. , , , . 8-hydroxyquinolines: A review of their metal chelating properties and medicinal applications. Drug Design, Development and Therapy 2013:1157. https://doi.org/10.2147/dddt.s49763
    [Google Scholar]
  98. , . Microbial enzymes: Tools for biotechnological processes. Biomolecules. 2014;4:117-139. https://doi.org/10.3390/biom4010117
    [Google Scholar]
  99. , , . A review on the antimicrobial assessment of triazole-azomethine functionalized frameworks incorporating transition metals. Journal of Molecular Structure. 2023;1288:135744. https://doi.org/10.1016/j.molstruc.2023.135744
    [Google Scholar]
  100. , , . New antimicrobial strategies based on metal complexes. Chemistry. 2020;2:849-899. https://doi.org/10.3390/chemistry2040056
    [Google Scholar]
  101. , , , , , . Antibacterial and pharmacological evaluation of fluoroquinolones: A chemoinformatics approach. Genomics & Informatics. 2018;16:44-51. https://doi.org/10.5808/gi.2018.16.3.44
    [Google Scholar]
  102. , , , . Metal-based ethanolamine-derived compounds: A note on their synthesis, characterization and bioactivity. Journal of Enzyme Inhibition and Medicinal Chemistry. 2016;31:88-97. https://doi.org/10.1080/14756366.2016.1220375
    [Google Scholar]
  103. , . In vitro antibacterial, antifungal and cytotoxic activities of some triazole Schiff bases and their oxovanadium(IV) complexes. Journal of Enzyme Inhibition and Medicinal Chemistry. 2013;28:1291-1299. https://doi.org/10.3109/14756366.2012.735666
    [Google Scholar]
  104. , , , , . Novel series of nanosized mono- and homobi-nuclear metal complexes of sulfathiazole azo dye ligand: Synthesis, characterization, DNA-binding affinity, and anticancer activity. Inorganic Chemistry Communications. 2019;108:107496. https://doi.org/10.1016/j.inoche.2019.107496
    [Google Scholar]
  105. , , , , . Synthesis, characterization and antioxidant activity of some transition metal complexes with terpenoid derivatives. Journal of the Chilean Chemical Society. 2011;56:911-917. https://doi.org/10.4067/s0717-9707(2011)000400019
    [Google Scholar]
  106. , , . Antimicrobial testing of the extracts of Samanea saman (Jacq.) Merr. http://derpharmachemica.com/vol2-iss6/DPC-2010-2-6-73-83.pdf
  107. , , , , . Spectral, XRD, SEM and biological activities of transition metal complexes of polydentate ligands containing thiazole moiety. Spectrochimica Acta. Part A, Molecular and Biomolecular Spectroscopy. 2008;71:628-635. https://doi.org/10.1016/j.saa.2008.01.023
    [Google Scholar]
  108. , , , , , , , . Chemical and structural data of (1,2,3-triazol-4-yl)pyridine-containing coordination compounds. Data in brief. 2018;20:1397-1408. https://doi.org/10.1016/j.dib.2018.08.125
    [Google Scholar]
  109. , , , , , , , , , , , , , , , , , , , , . Physicochemical properties and mechanism of action of a new copper(ii) pyrazine-based complex with high anticancer activity and selectivity towards cancer cells. RSC Advances. 2024;14:36295-36307. https://doi.org/10.1039/d4ra06874b
    [Google Scholar]
  110. . Syntheses, crystal structures and antimicrobial activity of copper (II) complexes with the ligand N, N’-Bis (4-bromosalicylidene) propane-1, 2-diamine. Acta Chimica Slovenica. 2023;70 https://doi.org/10.17344/acsi.2024.8707
    [Google Scholar]
  111. , , , , , , , . Theoretical exploration of enhanced antioxidant activity in copper complexes of tetrahydroxystilbenes: Insights into mechanisms and molecular interactions. ACS Omega. 2024;9:9076-9089. https://doi.org/10.1021/acsomega.3c07885
    [Google Scholar]
Show Sections