Abstract
Aim: The goal of this study is to synthesize new metal complexes containing N-methyl-1-(pyridin-2-yl)methanimine and azide ligands as α-glucosidase inhibitors for Type 2 diabetes. Materials & methods: The target complexes (12–16) were synthesized by reacting N-methyl-1-(pyridin-2-yl)methanimine (L1) with sodium azide in the presence of corresponding metal salts. The investigation of target protein interactions, vibrational, electronic and nonlinear optical properties for these complexes was performed by molecular docking and density functional theory studies. Results: Among these complexes, complex 13 (IC50 = 0.2802 ± 0.62 μM) containing Hg ion showed the highest α-glucosidase inhibitory property. On the other hand, significant results were detected for complexes containing Cu and Ag ions. Conclusion: Complex 13 may be an alternate anti-diabetic inhibitor according to in vitro/docking results.
Keywords: : α-glucosidase/docking, azide, Cu/Hg/Cd/Ag/Zn ions, Schiff base, spectral elucidation, TD/DFT methods
GRAPHICAL ABSTRACT
Plain language summary
Summary points.
In this study, new effective α-glucosidase inhibitor series of Schiff base-azide metal complexes {[Cu2(L1)2(N3)3], (12), [Hg(L1)2(N3)·H2O] (13), [Cd(L1)2(N3)], (14), [Ag(L1)(N3)2·2H2O], (15), [Zn2(L1)2(N3)2Cl2·4H2O], (16); L1: N-methyl-1-(pyridin-2-yl)methanimine} were synthesized. The electronic spectral features for these complexes characterized by 1H and 13C NMR (except for complex 12 including Cu ion), mass (LC-HRMS) and FT–IR spectra were examined by UV–Vis spectra.
In complexes 12–16, being α-glucosidase inhibitory potential, the IC50 values of 0.2802 ± 0.62 and 420.88 ± 1.42 μM range were obtained.
Considering the CAM-B3LYP and ωB97XD/6–311+G(d,p)//LanL2DZ levels of density functional theory (DFT), the optimal ground state geometries and vibrational wavenumbers of the complexes 12–16 were computed, and the obtained results were compared with the corresponding experimental wavenumbers.
The corresponding theoretical wavelengths, oscillator strengths as well as major contributions to the electronic transitions were calculated at TD-DFT/ωB97XD and TD-DFT/CAM-B3LYP levels. Additionally, the molecular orbitals and atoms/functional groups that make important contributions to electronic transitions were determined with the SWizard and Chemissian programs.
Moreover, to survey nonlinear optical parameters of the complexes 12–16, the CAM-B3LYP and ωB97XD/6–311+G(d,p)//LanL2DZ levels of density functional theory were used.
Demonstration of target protein interactions with complexes 12–16 has been fulfilled by molecular docking studies.
It is deduced that while IC50 results of the 15 and 12 against α-glucosidase are obtained remarkably, the β and γ results that complexes 15 and 12 are highly efficient in terms of microscopic second- and third-nonlinear optics properties, respectively.
The experimental vibrational wavenumbers and absorption wavelength results were compared with corresponding theoretical ones, it was observed that the results were quite compatible.
1. Background
Many studies of azide metal complexes, in which azide ions (N3-) coordinate with various metal ions, such as Cu(II) [1–4], Zn(II) [5], Co(II) [5], Mn(II) [5], etc. to form stable complexes, are notable in coordination chemistry. The structural, spectral, optical, magnetic [6–8] and pharmaceutical [9,10] properties of these metal complexes have attracted great attention due to their different reactivity and potential applications in various fields. The azide ion serves as a versatile ligand with its linear trigonal planar geometry that can bind to metal centers via nitrogen atoms. Furthermore, this ligand has a high degree of coordination flexibility, which can also participate in the formation of monodentate, bidentate or polydentate complexes. Depending on the metal ion and coordination environment, azide complexes exist a wide variety of geometries, including distorted trigonal bipyramidal [3], distorted square pyramidal [5] and distorted octahedral [5,7] and bridge structures [1,2,4,11]. It was reported that synthesis of metal complexes including azide and/or heterocyclic ligands with various transition metals, such as cobalt [4,6–8], copper [1–4], zinc [5], iron [8], erbium [12] and platinum [13]. The utilization of these complexes underscores their role in advancing our understanding and utilization of these intriguing materials. Additionally, azide/azide derivative metal complexes, as well as compounds with heterocyclic rings including nitrogen, have been investigated for many biological activities. These include antimicrobial and antifungal properties [3], as well as anticancer potential [14,15]. Their capacity to interact with biological targets and demonstrate selective cytotoxicity underscores their significance in biomedical research [16–19]. In this context, studying azide metal complexes having various application potentials in the fields of physics, chemistry, material science and biology is valuable in terms of the synthesis of new azide metal complexes, revealing new results by investigating their different properties, as well as contributing to the development of new technologies.
α-Glucosidase enzyme (EC 3.2.1.20), having a significant impact on the digestion and absorption of carbohydrates, is found in the small intestines of the digestive system and is known to provide the digestive process by breaking down complex carbohydrates into simple sugars [20–24]. With the help of this enzyme, the 1,4-glycosidic bond in polysaccharides is hydrolyzed, resulting in the creation of α-D-glucose. As a result, during absorption, α-D-glucose is formed and enters the circulation, which raises postprandial hyperglycemia [23,25]. The activity of the α-glucosidase enzyme is a process regulated by using α-glucosidase inhibitors, which bind to the α-glucosidase enzyme and reduce or stop its activity, as well as slowing down the digestive process and allowing carbohydrates to be absorbed more slowly. It is noteworthy that α-glucosidase inhibitors are widely used in the treatment of Type 2 diabetes and that Type 2 diabetes is a common metabolic disease worldwide. Acarbose, miglitol and voglibose, which function as α-glucosidase inhibitors, are currently used as clinically prescribed drugs in the treatment of Type 2 diabetes [22]. Although some side effects of these inhibitors cause digestive system problems such as abdominal pain, diarrhea and bloating, there are already many studies using α-glucosidase inhibitor [26–40]. Therefore, scientists are studying new enzyme inhibitors [41–45] associated with various diseases. They aim to develop options and clinical treatments [46] that are both safer and more effective against specific targets. The synthesis of more effective new inhibitors taking these effects into consideration continues to be an important research topic today.
In the development of science and technology, it is extremely important to correlate quantum chemical results with experimental results and calculate quantities that can be obtained directly from experimental measurements. Many studies conducted in recent years show that theoretical and experimental research are handled together and density functional theory (DFT) methods are used extensively. The electronic structure and characteristics of molecular systems (organic, inorganic, etc.) are extensively studied. This is achieved using the DFT levels [47–50], providing valuable insights into their properties and behavior. Additionally, further advancements in understanding are facilitated through the application of more sophisticated computational methods [51–53]. The ligands used in these systems allow examining the electronic structure, geometry, spectral and optical properties of coordination compounds consisting of metal ions/atoms and obtaining detailed information about the structure. In addition to comparing the values calculated using DFT methods with experimental data, it is important to shed light on the stability, electronic transitions, optical properties and magnetic behavior of metal complexes and to enable the design of new materials for various applications. The development of improved exchange-correlation functionals and the inclusion of solvent effects have further enhanced the predictive power of DFT in studying metal complex behavior. It seems that the selection of functionals suitable for the complex structure from the developed exchange-correlation functionals and the inclusion of solvent effects further increase the predictive power of DFT in examining metal complex behavior. CAM-B3LYP (a long-range corrected version of B3LYP using the Coulomb attenuation method from Handy and co-workers) [54] and ωB97XD (the last functional including empirical dispersion and long-range corrections suggested by Head-Gordon and co-workers) [55,56] functionals are known to reliably predict the geometries of metal complexes, energies and spectroscopic properties, making them a popular choice for a wide range of applications.
Azomethine and azide groups are regarded as highly potent ligands in medicinal applications of coordination chemistry [5,17,20,22]. Because these ligands have a high degree of coordination flexibility, which can also participate in the formation of monodentate, bidentate or multidentate complexes. Moreover, it is well-known that they play important roles in potential biological activities. Therefore, these ligands were chosen in the synthesis of target complexes. So as to design and create more powerful α-glucosidase inhibitors, novel five metal complexes (12–16) including N-methyl-1-(pyridin-2-yl)methanimine (L1) and azide (N3-) ligands were synthesized as a result of our ongoing studies. The structures of these complexes (12–16) were determined by 1H and 13C NMR (except for complex 12 including Cu ion), mass (LC-HRMS) and FTIR spectra. The IC50 values against α-glucosidase were obtained. In vitro α-glucosidase study on synthesized complexes and examination of their electronic properties by experimental and DFT methods play an important role in determining the structure–activity relationship. In addition, molecular docking study was conducted to determine the interactions of the target protein with the ligands and to interpret the activity results. Furthermore, the unique properties and capabilities of the DFT/ωB97XD and DFT/CAM-B3LYP functionalities allow these methods to obtain deeper insights into the behavior of metal complexes in structure–property relationships.
2. Experimental & computational details
2.1. Spectroscopic implements
The NMR spectrometer (Varian Infinity Plus, Oxford, UK) was utilized to define the structure of the synthesized the N-methyl-1-(pyridin-2-yl)methanimine (L1) ligand and complexes 13–16. The synthesized compounds were dissolved in 500 μl of DMSO-d6 containing tetramethylsilane (0.03 vol. %) as an internal standard. Temperature of the samples during the data acquisition was at 29°C. The 1H and 13C NMR measurements were fulfilled at 300 and 75 MHz, respectively. The 1H and 13C NMR spectra for each sample were obtained with 32 and 5000 accumulations, and pulse angles were 45 and 20 degrees, respectively. The structures of novel five metal complexes (12–16) were determined by using the SHIMADZU LCMS-9030 (OR, USA) model triple quadrupole liquid chromatograph mass spectrometer. The concentrations of samples were prepared at 50 ppm in methanol. LC-MS spectra were recorded at the following conditions: Mobile phase A: acetonitrile; Mobile phase B: water; pump system A:B concentration: 50:50%; flow: 0.5000 ml/min; pump A and B pressures: 30 bar; oven temperature: 28.7°C; interface type: electrospray ionisation; interface voltage (+)/(-): +4.50/-3.50 kV. The FT-IR spectra in the range of 4000–400 cm-1 for complexes 12–16 in powder form were recorded at the Perkin Elmer UATR-TWO (Perkin Elmer Spectrum-two equipped with ATR) spectrophotometer (CT, USA). The methanolic solutions of samples were prepared at 10-5 M and electronic absorption wavelengths (UV-Vis spectra) in the 900-200 nm range were obtained with the SHIMADZU UV-2600 UV-Vis spectrophotometer. Using the CHNS-932 (Leco, MI, USA) elemental analyzer, elemental studies of the synthesized complexes 12–16 in powder form were carried out.
2.2. The synthesis of the N-methyl-1-(pyridin-2-yl)methanimine (L1) & complexes 12–16 {[Cu2(L1)2(N3)3], (12), [Hg(L1)2(N3)·H2O], (13), [Cd(L1)2(N3)], (14), [Ag(L1)(N3)2·2H2O], (15), [Zn2(L1)2(N3)2Cl2·4H2O], (16)}
2-Pyridinecarboxaldehyde (99%), methylamine solution (33 wt.% in absolute ethanol), sodium azide (≥99.5%), copper(II) acetate monohydrate (≥99.0%, Cu(OAc)2·H2O), mercury(II) acetate (≥98.0%, Hg(OAc)2), cadmium(II) acetate (99.995%, Cd(OAc)2), silver nitrate (≥99.0%, AgNO3), zinc chloride (≥98%, ZnCl2) reagents, magnesium sulfate (anhydrous, ≥99.5%), triethylamine (for synthesis), ethanol (absolute for analysis), dimethyl sulfoxide-D6 (DMSO-d6, ≥99.8%) purchased from Sigma-Aldrich (Sternheim, Germany) and Merck (Darmstadt, Germany) were utilized without any special reprocessing. Figure 1 depicts the synthesis methods of complexes 12–16.
Figure 1.

Synthesis of complexes 12–16.
rt: Room temperature.
The synthesis and NMR spectra of the N-methyl-1-(pyridin-2-yl)methanimine is presented in Supplementary Material (Supplementary Figure S1).
Synthesis of complexes 12–16 (Figure 1): 1 mmol of corresponding metal salts ((Cu(OAc)2·H2O/Hg(OAc)2/Cd(OAc)2/AgNO3/ZnCl2) were added to N-methyl-1-(pyridin-2-yl)methanimine (L1) (2 mmol) dissolved in 15 ml of absolute ethanol. After 20 min, 2 mmol NaN3 dissolved in 5 ml water was added to this solution. The mixture was stirred at ambient temperature for overnight (complexes 12–14) or for 3 h (complexes 15,16), then allowed to evaporate at ambient temperature. It took around 7 to 10 days to acquire complex compounds in powder form. The structures of these products (12–16) were confirmed by 1H and 13C NMR (except for complex 12 including Cu ion), and mass spectroscopy. Complex 12: Anal. Calc. for C14H16Cu2N13: C, 34.08; H, 3.27; N, 36.90; found: C, 34.42; H, 3.45; N, 36.63. Calculated exact mass (m/z): 492.0244 (C14H16Cu2N13); found: LC-HRMS (m/z): 492.9582 ([M]+). Complex 13: 1H NMR (DMSO-d6, 300 MHz) δ/ppm: 8.66 (1H, d, J = 4.6 Hz), 8.62 (1H, s), 7.99–8.05 (1H, m), 7.91–7.94 (1H, m), 7.60–7.66 (1H, m), 3.51 (3H, s). 13C NMR (DMSO-d6, 75 MHz) δ/ppm: 162.65, 162.60, 150.70, 150.65, 147.55, 140.98, 140.90, 129.24, 129.22, 129.20, 129.16, 128.30, 46.93, 46.84. Anal. Calc. for C14H18HgN7O: C, 33.57; H, 3.62; N, 19.57; found: C, 33.86; H, 3.84; N, 19.02. Calculated exact mass (m/z): 502.1279 (C14H18HgN7O); found: LC-HRMS (m/z): 502.7797 ([M]+). Complex 14: 1H NMR (DMSO-d6, 300 MHz) δ/ppm: 8.72 (1H, d, J = 4.7 Hz), 8.51 (1H, d, J = 1.5 Hz), 8.05–8.12 (1H, m), 7.90–7.93 (1H, m), 7.63–7.68 (1H, m), 3.41 (3H, s). 13C NMR (DMSO-d6, 75 MHz) δ/ppm: 161.89, 150.06, 150.02, 140.70, 128.49, 127.45, 46.36. Anal. Calc. for C14H16CdN7: C, 42.60; H, 4.09; N, 24.84; found: C, 42.12; H, 3.86; N, 25.27. Calculated exact mass (m/z): 396.0501 (C14H16CdN7); found: LC-HRMS (m/z): 399.0392 ([M+2H]+). Complex 15: 1H NMR (DMSO-d6, 300 MHz) δ/ppm: 8.64 (1H, d, J = 4.2 Hz), 8.52 (1H, d, J = 1.5 Hz), 7.99–8.05 (1H, m), 7.86–7.88 (1H, m), 7.58–7.62 (1H, m), 3.56 (3H, s). 13C NMR (DMSO-d6, 75 MHz) δ/ppm: 163.83, 151.61, 150.98, 139.06, 127.43, 125.50, 48.66. Anal. Calc. for C7H12AgN8O2: C, 24.15; H, 3.47; N, 32.19; found: C, 24.84; H, 3.75; N, 31.82. Calculated exact mass (m/z): 347.0134 (C7H12AgN8O2); found: LC-HRMS (m/z): 348.7087 ([M]+). Complex 16: 1H NMR (DMSO-d6, 300 MHz) δ/ppm: 8.65 (1H, s, br), 8.58 (1H, s), 8.16 (1H, s, br), 8.00 (1H, s, br), 7.73 (1H, s, br), 3.37 (3H, s). 13C NMR (DMSO-d6, 75 MHz) δ/ppm: 162.43, 162.39, 149.76, 149.66, 140.64, 140.61, 128.22, 128.14, 128.08, 126.09, 126.02, 45.62, 45.54. Anal. Calc. for C14H24Cl2N10O4Zn2: C, 28.12; H, 4.04; N, 23.42; found: C, 28.65; H, 4.48; N, 22.94. Calculated exact mass (m/z): 593.9942 (C14H24Cl2N10O4Zn2); found: LC-HRMS (m/z): 597.1465 ([M+2H]+). The 1H and 13C NMR (except for complex 12 including Cu ion), and mass (LC-HRMS) spectra were presented in Supplementary Material (Supplementary Figures S2–S10).
2.3. α-Glucosidase inhibitor activity
The IC50 values against α-glucosidase of complexes 12–16 were defined taking into considering reported previously [30–32,36]. The detailed technique of α-glucosidase assay was provided in Supplementary Material. Taking into account Ac (the absorbance of the control) and As (the absorbance of the samples) in the equation below, the IC50 values against α-glucosidase for complexes 12–16 were obtained.
The calculations of IC50 for complexes 12–16 were used the Graphpad Software.
2.4. Computational details
To investigate the features of structural, vibrational, electronic and nonlinear optics (NLO), the Gaussian 16, Revision C.01 [57] and GaussView 6 [58] programs were utilized. The CAM-B3LYP [54] and ωB97XD methods [55,56] with 6–311+G(d,p)//LanL2DZ [59,60] basis set in the ground state were applied to obtain the optimum geometrical structures, vibrational frequencies and natural bond orbital (NBO) results for complexes 12 and 16. For complexes 13–15, the same properties were examined by considering DFT/CAM-B3LYP and ωB97XD/LanL2DZ levels. Time-dependent DFT (TD–DFT) levels [61] with the conductor-like polarizable continuum model (CPCM) [62] were used to investigate electronic absorption wavelengths (λ) and oscillator strength of complexes 12–16 in MeOH. The SWizard [63] and Chemissian [64] programs were applied to define major contributions in electronic transitions. By using the CAM-B3LYP and ωB97XD levels in gas phase, the NLO properties (the parameters of microscopic polarizabilities (α and Δα), first- and second-order hyperpolarizabilities (β and γ) via equations presented in Supplementary Material were examined. Finally, the DFT/CAM-B3LYP level was utilized to survey the molecular electrostatic potential (MEP) surfaces.
2.5. Docking procedure
AutoDock4 program [65] and AutoDockTools (ADT1.5.7) [66] was utilized to examine the interactions between the Saccharomyces cerevisiae isomaltase (Protein Data Bank [PDB] ID: 3A4A) template structure provided from the Protein Data Bank and synthesized complexes 12–16. By considering the optimized structures of complexes 12–16 and genistein obtained at the DFT/CAM-B3LYP method, the inhibition constants (Ki) and binding energies of these complexes were calculated from docking findings. The Discovery Studio 4.0 program [67] provides 2D and 3D views of interactions between amino acid residues in enzyme structure and complexes 12, 13 and 15. The docking details are also given in Supplementary Material.
3. Results & discussion
3.1. The chemical descriptions of the produced complexes
The N-methyl-1-(pyridin-2-yl)methanimine (L1) was characterized by 1H and 13C NMR spectra. The structures of synthesized complexes 12–16 were verified by using 1H and 13C NMR (except for complex 12 including Cu ion), mass spectra (Supplementary Figures S2–S10), as well as FTIR spectra. The aromatic proton signals for L1 appeared at the range of 7.44 and 7.93 ppm, aromatic CH neighbor pyridine-N and CH protons belonging to imine were observed at 8.61 and 8.33 ppm, respectively (in 1H NMR spectra given in Supplementary Figure S1). In 1H NMR spectra of complexes 13–16, the aromatic protons, aromatic CH neighbor pyridine-N and CH protons belonging to imine shifting the peak downfield were obtained at the range from 7.58 to 8.16 ppm, from 8.64 to 8.72 ppm, and from 8.51 to 8.62 ppm, respectively (Supplementary Figures S3, S5, S7 & S9). The aromatic carbon signals for the L1 ligand emerged at the range of 120.89 and 163.84 ppm, while these carbon signals for complexes 13–16 were detected at the range from 125.50 to 163.83 ppm (in 13C NMR spectra given in Supplementary Figures S1, S3, S5, S7 & S9). The carbon signal belonging to imine in L1 ligand was observed at 154.81 ppm, whereas this signal shifting the peak upfield was obtained at the range of 149.76 and 151.61 ppm in complexes 13–16 (Supplementary Figures S3, S5, S7 & S9). Considering these chemical shifts in upfield or downfield in NMR spectra of L1 ligand and complexes 13–16, the forming of complexes was supported. Besides, the structures of complexes 12–16 verified by mass spectra (Supplementary Figures S2, S4, S6, S8 & S10).
Supplementary Figure S11 displays the optimized structures (obtained at DFT/CAM-B3LYP) of complexes 12–16 confirmed by 1H and 13C NMR (except for complex 12 including Cu ion), and mass spectra (Supplementary Figures S2–S10). DFT/ωB97XD and DFT/CAM-B3LYP levels were applied to obtain the geometrical parameters (bond lengths and angles and dihedral angles) given in Supplementary Table S1. The azide ions (N3-), can both enable the assembly of binuclear complexes and bind to transition metal atoms with different coordination modes [68].
Complex 12 [(Cu)2(L1)2(N3)3] consists of the coordination of two L1 ligands and one azide ligand with Cu(II) ion by forming a bridge with two Cu(II) ions over the N atom of the azide ligand. It was understood that a complex structure with distorted tetrahedral geometry was formed by the coordination of the L1 ligand and azide ligands to Cu(II) ions via N atoms.
Complex 13 [Hg(L1)2(N3)·H2O] and complex 14 [Cd(L1)2(N3)] consist of the coordination of two L1 ligands and an azide ligand with the Hg(II) and Cd(II) ions, respectively. It was understood that both complex structures with distorted trigonal bipyrimidal geometry was formed when two L1 ligands and an azide ligand coordinated to the Hg(II) and Cd(II) ions via N atoms. Besides, complex 13 includes the uncoordinated water molecule.
Complex 15 [Ag(L1)(N3)2·2H2O] is comprised of an L1 ligand and the coordination of two azide ligands with an Ag(I) ion. It was understood that a complex structure with distorted tetrahedral geometry was formed when an L1 ligands and two azide ligands coordinated to the Ag(I) ion via N atoms. Complex 15 contains also the uncoordinated two water molecules.
Complex 16 [(Zn)2(L1)2(N3)2Cl2·4H2O] consists of the coordination of two L1 and azide ligands and a Cl atom with Zn(II) ion by forming a tetrahedral bridge with two Zn(II) ions over the N atoms of two azide ligands. It was understood that a complex structure with distorted trigonal bipyramidal geometry was formed by the coordination of an L1 and two azide ligands and a Cl atom to Zn(II) ion via N atoms.
These results are consistent with the coordination geometries of Cu(II), Zn(II) and Ag(I) coordinated with different ligands [1,2,5,6,69,70].
Supplementary Table S1 contains theoretical some geometrical parameters for complexes 12–16. These complexes are comprised of five-membered chelate rings occurred by L1 ligand with metal ions. For complexes 12–16, the M(metal)–N2 bond lengths between the metal ion and pyridine N atom of the complexes were obtained at 2.143, 2.350, 2.418, 2.253, 2.280 Å by using DFT/CAM-B3LYP level. The similar bond lengths between the metal ion and imine N atom of the complexes were calculated at 2.086, 2.509, 2.263, 2.355, 2.135 Å by using the same DFT level, respectively. In addition, in all complexes, the theoretical bond lengths obtained by the DFT/ωB97XD level are similar trends. For complexes 12–16, the N2–M–N5 bond angles among imine N, metal ion and azide N in the coordination environment were found to be 78.09, 71.93, 68.82, 73.12, 74.87° with DFT/CAM-B3LYP level, respectively. On the other hand, the N–M–N bond angles among imine/pyridine N, metal ion and azide N in the coordination environment were observed the values greater or less than 90°, regardless of coordination geometry. These results indicate the presence of a distorted coordination geometry in the synthesized complexes 12–16. Despite the coordination environment/geometry differences, the theoretical results of the complexes are compatible with the previously reported ones [1,2,4,5,69–71].
Studying conjugative interactions, charge transfer in molecular systems and intra- and inter-molecular bonding are accomplished successfully with NBO analysis [72–74]. By considering the second-order perturbation approach [75,76], the hyperconjugative interactions were obtained. In NBO analysis, the large E(2) energy value is an indication that the interactions between electron acceptors and electron donors are more intense and intramolecular and intermolecular charge transitions are high. The natural bond orbital calculation outcomes of the synthesized complexes 12–16 are given in Supplementary Table S2. The interactions between LP(n) nitrogen orbital pairs and LP*(n) metal(II)/(I) (Cu(II), Hg(II), Cd(II), Ag(I), Zn(II)) displayed the n→n* interactions indicate the coordination environment around the metal ions in the complex structures. The E(2) values belonging to these interactions were obtained at the range of 6.01 and 47.25 kcal/mol at the DFT/CAM-B3LYP level (Supplementary Table S2). Additional significant interactions in the complexes indicate stabilization within the L1 ligand. It is concluded that the bonding and antibonding orbitals exhibit charge transfer (CT) interactions, which cause the system to stabilize.
The FTIR spectra for the synthesized complexes 12–16 demonstrate the wavenumbers in the range of 4000–400 cm-1. The corresponding theoretical vibrational frequencies computed at the DFT/ωB97XD and DFT/CAM-B3LYP levels to get closer to the experimental IR spectra were multiplied by 0.96. The experimental and theoretical characteristic vibrational wavenumbers were presented in Supplementary Figure S12, Supplementary Tables S1 & S4. The νCH(ring), νCH(imine) and νCH(methyl) stretching modes appeared at the range from 3095 to 2926 cm-1 for complex 12, from 3137 to 2931 cm-1 for complex 13, from 3089 to 2919 cm-1 for complex 14, from 3070 to 2917 cm-1 for complex 15, from 3070 to 2910 cm-1 for complex 16. These results are coherent with previously reported for the Schiff base metal complexes [3,18,21,28,29,31,68]. The corresponding ones for complexes 12–16 obtained at DFT/ωB97XD were found at 3093–2904 cm-1 range, 3150–2895 cm-1 range, 3089–2897 cm-1 range, 3117–2927 cm-1 range and 3073–2910 cm-1 range, respectively (Supplementary Tables S3 & S4). The vibrational bands emerging at the range of 2032 and 2060 cm-1 were assigned as νNN asymmetric stretching modes belonging to the azide ligand. These findings are compatible with the modes reported in different complexes [2,4,5]. Theoretical these bands were calculated at 1885 and 2178 cm-1, and 1844 and 2185 cm-1 ranges, in order of the DFT/ωB97XD and DFT/CAM-B3LYP levels. Moreover, the νC=N(imine) stretching modes of the C=N bond (imine) of the L1 for complexes 12–16 observed at 1654, 1653, 1658, 1648 and 1655 cm-1 were obtained theoretically at 1662, 1671, 1687, 1692 and 1701 cm-1 (with DFT/ωB97XD level), respectively (Supplementary Tables S3 & S4). Besides, the νNC(imine) stretching modes of the N–CH3 bond of the L1 for complexes 12–16 emerged at the range of 1015 and 1028 cm-1 range, and the corresponding values computed at DFT/ωB97XD and DFT/CAM-B3LYP levels were obtained at the range of 1011 and 1025 cm-1, 1006 and 1023 cm-1, respectively. The νCC(ring) stretching modes attributed to ring C=C bond emerged within the ranges 1588–1623 cm-1 in FTIR spectra, 1589–1614 cm-1 in DFT/ωB97XD level, and 1588–1615 cm-1 in DFT/CAM-B3LYP level, as can be seen in Supplementary Tables S3 & S4. These results are consistent with similar group frequency values in different structures in previous reports [5,20,22,26–28,71]. The vibration bands appearing at 1341, 1304, 1330, 1333 and 1340 cm-1, which are assigned as the νNN(azide) symmetric stretching vibration, are attributed to the coordination of the metal ion with the N atom of the azide. The corresponding vibrational modes obtained at DFT/ωB97XD and DFT/CAM-B3LYP levels were found at the range from 1291 to 1362, and from 1283 to 1359 cm-1, respectively. The βHCC(ring) in-plane bending modes appeared at the range of 1039 and 1573 cm-1 in FTIR spectra were found at the range 1039 and 1591 cm-1 in DFT/ωB97XD level, and range 1036 and 1590 cm-1 DFT/CAM-B3LYP level. Furthermore, the detailed stretching, in-plane and out-of-plane bending, torsion vibrational modes belonging to ring, azide, imine and methyl groups for complexes 12–16 are tabulated in Supplementary Table S4.
3.2. Analysis of the electronic absorption spectra & electronic transitions
The electronic absorption spectra and the plot of experimental Eg (optical band gap energy) in methanol for synthesized complexes 12–16 are given in Figure 2. The corresponding theoretical wavelengths, oscillator strengths as well as major contributions to the electronic transitions were computed at TD-DFT/ωB97XD and TD-DFT/CAM-B3LYP levels. Additionally, the molecular orbitals and atoms/functional groups that make important contributions to electronic transitions were determined with the SWizard [63] and Chemissian [64] programs. According to Figure 2A, three absorption peaks appeared in bridge-structured complexes 12 and 16, while two maximum absorption peaks emerged in other complexes 13–15. In complex 12, the absorption peaks appeared at 371.1, 264.5 and 203.5 nm, while in complex 16, the absorption peaks emerged at 280.6, 231.5 and 213.8 nm. The absorption peaks of 251.1 and 207.3 nm for complex 13, 264.5 and 244.5 nm for complex 14, and 258.4 and 204.6 nm for complex 15 were detected (Supplementary Table S5). These peaks were attributed to intra-ligand π→π*, ligand-to-metal/metal-to-ligand and n→π* transitions. Figure 3 displays ligand-ligand charge transfer in the high energy region as charge transitions within the ligand.
Figure 2.

Optical spectra. (A) The UV-Vis spectra and (B) the plot of experimental Eg (optical band gap energy) in methanol for complexes 12–16.
Figure 3.
Frontier molecular orbitals and their energies that contribute the most to the electronic transition obtained by the DFT/CAM-B3LYP method in methanol for complexes 12–16.
DFT/CAM-B3LYP: Density functional theory/Coulomb attenuation method- Becke, 3–parameter, Lee–Yang–Parr; FMO: Frontier molecular orbital: HOMO: Highest occupied molecular orbital; LUMO: Lowest occupied molecular orbital.
It has been reported that large contributions from metal orbitals to the lowest occupied molecular orbital (LUMO) in complex structures can increase the π-acceptor property of the ligand [77]. The metal orbital contributions to the highest occupied molecular orbital (HOMO)/HOMOs and LUMO/LUMOs of complexes 12–16 were obtained at the range from 2 to 37% (HOMO/HOMOs) and from 3 to 37% (LUMO/LUMOs), by using TD-CAM-B3LYP level. The LUMO/LUMOs contributions of the L1 including 2-pyridyl and imine groups, and azide (N3) ligands for complexes 12–16 vary between 20 and 100% (in L1), 0 and 45% (in N3), respectively (Supplementary Table S5). Due to the obtained larger LUMO/LUMOs contributions for L1, the L1 ligand has the more electron acceptor property. It is clear from Figure 3 that the electron density for complex 12 is localized in the Cu ion, L1, and azide on one side for the HOMO α-level, localized in the L1 and azide on the other side for LUMO α-level. In complexes 13 and 14, HOMO and LUMO α-levels are localized on the L1 ligand including pyridyl and imine group in opposite parts. For the HOMO α-levels of complexes 15 and 16, the electron density is mostly localized on the metal ions and azide ligand, while in the LUMO α-levels, it is predominantly located on the L1 ligand. (Figure 3). Supplementary Table S5 shows the remarkable metal-to-ligand charge transfer (MLCT), ligand-to-ligand charge transfer (LLCT) and ligand-to-metal charge transfer (LMCT) transitions. The 54% contribution for HOMO→LUMO α spin for complex 12 depicts the MLCT transition obtained at the contributions of Cu (20%), L1 (30%) containing pyridyl (22%) and imine (8%) groups and N3 (50%) ligand in HOMO, and L1 (61%) and N3 (36%) in LUMO. In complex 13, the absorption wavelength obtained at 536.2 nm (f = 0.0007) displays the contribution of 99% HOMO→LUMO changing from pyridyl (45%), imine (51%), H2O (2%) to pyridyl (60%), imine (38%), N3 (2%) and attributed to the LLCT transition. The LMCT transition for complex 15 was calculated at the contribution of 43% H-1→L β spin coming from pyridyl (7%), imine (2%), N3 (90%) to Ag (37%), pyridyl (11%), imine (9%), N3 (43%). The detailed percentage contributions in HOMOs→LUMOs transitions and assignments for HOMOs and LUMOs were presented in Supplementary Table S5, considering the metal ion and its environments. The frontier molecular orbital contributions in electronic transitions obtained DFT/ωB97XD level are presented in Supplementary Table S6.
MEP surfaces are used to understand the structure–activity relationship (SAR) for many molecular systems. These surfaces are related to electron density. The red and yellow areas on the MEP surface illustrate the negatively charged region with high electron density while the blue areas are spaces with positive charges. Green areas depict spaces where there is no potential and it is neutral [78]. MEP surfaces of complexes 12–16 were described by DFT/CAM-B3LYP level. The regions of complexes 12–16 with the highest electron density are shown in Supplementary Figure S13, where lone pairs of electronegative atoms are linked to the strong red electron clouds around the azide ligand. Areas showing nucleophilic reactivity were observed around the -CH groups in the L1 ligand, which has the highest positive potential. In this way, the possible hydrogen bond interactions and reactive parts can be revealed in these complexes.
It is concluded from Figure 2B that the experimental optical band gap energy (Eg) values of complexes 12–16 were obtained at 4.01, 4.37, 4.09, 4.48 and 4.16 eV, in order. The HOMO-LUMO band gap results obtained from DFT/CAM-B3LYP calculations in α-spin level were obtained at 5.172, 3.797, 2.952, 5.672 and 6.422 eV, respectively. It seems that there are consistent results between experimental and theoretical approach errors. Molecular parameters χ (electronegativity), η (chemical hardness), S (chemical softness), ω (electrophilicity), φ (nucleophilicity) were calculated using HOMO and LUMO energy values [79] and their results are given in Supplementary Table S7. Chemical hardness (η) is known as the resistance of chemical systems to changing their electronic distribution. The η, χ, S parameters are indicators of chemical reactivity and stability of systems. The χ, η, S, ω and φ parameters of complexes 12–16 in α spin obtained at the DFT/CAM-B3LYP level were found at the range from 2.957 to 4.645 eV, from 1.476 to 3.211 eV, from 0.311 to 0.678 1/eV, from 2.302 to 3.360 eV, from 0.298 to 0.434 1/eV, respectively, seen in Supplementary Table S7. The low η parameters obtained in these synthesized complexes indicate that intra-ligand charge transfer (ILCT) occurs. These results are coherent with previously reported ones in different structures [77,80].
3.3. The analysis of nonlinear optics parameters
The synthesis and related computational research of NLO materials have been intensively studied due to their potential applications in optoelectronics and telecommunications technology, such as data storage and processing [81–87]. Theoretical NLO studies are taken into consideration in order to calculate the response of molecules to these properties if doing an experimental investigation is not feasible on the NLO properties of newly synthesized molecules. The findings obtained from these studies have the potential to shed light on electronic devices that are considered to be developed or designed. In this context, the and Δα indicating the linear optical (LO), β/γ demonstrating the second-/third-order NLO parameters of complexes 12–16 were obtained at DFT/ωB97XD and DFT/CAM-B3LYP levels in gas phase by using equations (7)-(10) [81,83] given in Supplementary Material, as can be seen in Supplementary Table S8.
To compare/evaluate the obtained theoretical results, p-nitroaniline (pNA) [88] and urea [89,90] were referenced as prototype compounds. According to the DFT/CAM-B3LYP level, the mean polarizability () values of complexes 12–16 in the gas phase obtained at 29.78 × 10-24 and 49.75 × 10-24 esu ranges exhibited higher than that of pNA (17 × 10-24 esu) [88]. It has been observed that the polarization tendency increases in the large molecules. On the other hand, the β parameters of complexes 12–16 in the gas phase were obtained at the range of 2.28 × 10-30 and 18.21 × 10-30 esu, 2.33 × 10-30 and 17.05 × 10-30 esu in order of the DFT/ωB97XD and /CAM-B3LYP levels (Supplementary Table S8). While these results were observed to be greater than the β results of urea (0.32 × 10-30 esu [89] and 0.130 × 10-30 esu [90]), lower values than those of pNA (9.2 × 10-30 esu [88]) were obtained, except for complexes 13,16. Based on the γ values for complexes 12–16 in the gas phase in the same levels, the highest values of 85.71 × 10-36 and 96.93 × 10-36 esu for complex 12 were calculated while the smallest values of 1.66 × 10-36 and 15.32 × 10-36 esu for complex 15. It is observed that the γ values of complex 12 are about 5.7/12.2 and 6.5/13.8-times higher than those of pNA (15 × 10-36 esu) [88]/urea (7 × 10-36 esu) [89]. It is concluded from the β and γ results that complexes 15 and 12 are high efficiency regarding the microscopic second- and third-NLO features, respectively. Depending on the electronic structure of the complexes, the effects of the metal ion and its coordination environment in the structure were also observed in the differences in the calculated NLO parameters.
3.4. α-Glucosidase inhibitory activity
The IC50 values for complexes 12–16 were obtained at the range from 0.2802 ± 0.62 and 420.88 ± 1.42 μM, seen in Table 1. Among these complexes against α-glucosidase, the complex 13 (IC50 = 0.2802 ± 0.62 μM) exhibits the strongest inhibitory activity, whereas the complex 14 (IC50 = 420.88 ± 1.42 μM) displays the lowest inhibition effect. Complex 13 (IC50 = 0.2802 ± 0.62 μM) was observed at 48.39-times more effective compared with genistein (IC50 = 13.56 ± 1.21 μM), utilized as a standard in this study.
Table 1. In vitro inhibition IC50 values (μM) for α-glucosidase of complexes 12–16 and ligands.
| Ligand/complex | IC50 (μM)† |
|---|---|
| N-methyl-1-(pyridin-2-yl)methanimine (L1) | Not active |
| Sodium azide (NaN3) | Not active |
| Complex 12 [Cu2(L1)2(N3)3] | 1.562 ± 0.83 |
| Complex 13 [Hg(L1)2(N3)·H2O] | 0.2802 ± 0.62 |
| Complex 14 [Cd(L1)2(N3)] | 420.88 ± 1.42 |
| Complex 15 [Ag(L1)(N3)2·2H2O] | 1.274 ± 0.32 |
| Complex 16 [Zn2(L1)2(N3)2Cl2·4H2O] | 171.5 ± 1.72 |
| Genistein | 13.56 ± 1.21 |
IC50 values represent the means ± S.E.M. of three parallel measurements (p < 0.05).
For the SAR evaluation based on Table 1, the following results are deduced:
The coordination of complex 13 [Hg(L1)2(N3)·H2O] and complex 14 [Cd(L1)2(N3)] were defined as a distorted trigonal bipyrimidal geometry. Complex 13 (IC50 = 0.2802 ± 0.62 μM), having the larger atomic diameter of Hg, has better inhibitory activity than complex 14 (IC50 = 420.88 ± 1.42 μM).
The coordination of complex 12 [(Cu)2(L1)2(N3)3] and 16 [(Zn)2(L1)2(N3)2Cl2·4H2O] have a bridge structure with azide ligand. However, complex 12 was defined as a distorted tetrahedral geometry formed by the coordination of the L1 ligand and azide ligands to Cu(II) ions via N atoms while complex 16 was described as a distorted trigonal bipyramidal geometry formed by the coordination of an L1 and two azide ligands, and a Cl atom to Zn(II) ion via N atoms. Besides the differences in coordination geometry and environment between complexes 12 and 16, complex 12 (IC50 = 1.562 ± 0.83 μM) with a larger atomic diameter showed a higher inhibition effect than complex 16 (IC50 = 171.5 ± 1.72 μM).
Moreover, the α-glucosidase inhibition values of complexes 12 and 15 obtained at 1.562 ± 0.83 μM and 1.274 ± 0.32 μM, respectively. It can be stated that the small inhibition difference between complexes 15 and 12 is due to the difference in coordination in the structures as well as the fact that the atomic diameter of Cu metal is smaller than that of Ag.
As a result, it is clear that there are differences in IC50 values depending on the role of metal ions and the coordination environment.
The different organic compounds were reported as α-glucosidase inhibitors with the IC50 values ranging from 4.51 ± 0.09 μM to 45.7 ± 0.23 μM [37–39]. The IC50 value of the Zn complex including N2O2 was observed at 0.23 mM [20]. In addition, the IC50 values of Cr, Mn, Co and Ni complexes including pyridine derivatives could not be observed but were obtained between 0.22 and 0.92 mM in other complexes [30,91,92]. Moreover, IC50 values for transition metal complexes between 6-methylpicolinic acid/Schiff-base and isothiocyanate were obtained at the range of 0.2376 ± 0.82 and >600 μM [30,71].
Additionally, it can be stated that complex 13 (IC50 = 0.2802 ± 0.62) may be a good candidate for α-glucosidase inhibitor when the inhibition result is compared with previous studies involving similar metal ions [31,32,40,71,93,94]. The IC50 values of complexes 12 (1.562 ± 0.83) and 15 (1.274 ± 0.32) were also found to be significant. It is concluded that some metal complexes containing Schiff base and azide ligands tend to increase α-glucosidase inhibition.
Since the structures in the synthesized complexes are not of similar geometry and coordination, a direct correlation is not expected between density functional theory calculations and the reactivity and α-glucosidase inhibitor activity of the compounds. In addition, the reactivities of the compounds depend on many different parameters (steric effect, reaction medium, structures of other components and nucleophilic and electrophilic behaviors, etc.). For these reasons, a direct relationship between the chemical reactivity parameter results calculated using DFT methods for these complexes and their inhibitor activities could not be obtained. However, an inverse correlation was observed between electric dipole moment and inhibitor activity of synthesized complexes. Complex 13, the strongest inhibitor (IC50 = 0.280 μM) in this study, has the lowest dipole moment (μ = 5.87 Debye), while complex 14 and 16, the weakest inhibitors (IC50 = 420.88 and 171.50 μM, respectively), have the highest dipole moment (μ = 12.25 and 11.77 Debye, respectively). On the other hand, a linear increase was seen between LUMO energies with rising inhibitory activity. Complex 13 has the highest ELUMO energy, -1.058 eV, whereas ELUMO energy of complex 14 is -1.538 eV.
3.5. Molecular docking & inhibition mechanism
Using the AutoDockTools (ADT1.5.7) [66] graphical user interface with the AutoDock4 software, the ligand–protein interactions of the complexes 12–16 were determined, and S. cerevisiae isomaltase (PDB ID: 3A4A) as target protein was considered as rigid. The binding energies and estimated inhibition constants (Ki) were defined by docking results. Discovery Studio 4.0 program [67] was utilized to display the 2D and 3D structures of the interactions between ligands and amino acid residues (Figures 4–6 & Supplementary Figures S14 & S15).
Figure 4.

3D- and 2D-structures displaying interactions between amino acid residues and ligands of the most active complex 12.
Figure 5.

3D- and 2D-structures displaying interactions between amino acid residues and ligands of the most active complex 13.
Figure 6.

3D- and 2D-structures displaying interactions between amino acid residues and ligands of the most active complex 15.
The binding energy values of the genistein taken as a standard inhibitor [71] and complexes 12–16 were found to be -6.35, -4.77, -4.97, -4.43, -4.93 and -4.70 kcal/mol, respectively. The Ki values related to these energies were calculated at 22.06 μM, 316.95 μM, 226.71 μM, 567.97 μM, 241.40 μM and 358.10 μM, respectively, seen in Table 2. The best inhibition values against α-glucosidase among complexes were obtained for complex 13 (IC50 = 0.2802 ± 0.62 μM), at same time the lowest Ki value (26.71 μM) and binding energy (-4.97 kcal/mol) were calculated by considering docking results. The in vitro results were supported by docking study in order of complex from lowest to highest inhibitory activity.
Table 2. Protein–ligand interactions and distance values for complexes 12–16 and genistein.
| Substrate | Receptor | Interaction | Distance (Å) | Ki (μM) | Binding energy (kcal/mol) | Ref. |
|---|---|---|---|---|---|---|
| Complex 12 | 316.95 | -4.77 | ||||
| Azide (N) | Glu322 (OE1) | Charge–Charge | 5.14 | |||
| Azide (N) | Phe321(O) | Conventional H-Bond | 3.03 | |||
| Azide (N) | Phe321 | Pi–Cation | 4.92 | |||
| L1 | Leu323 (CD1) | Pi–Sigma | 3.34 | |||
| Complex 13 | 226.71 | -4.97 | ||||
| Azide (N) | Lys523 (O) | Conventional H-Bond | 2.53 | |||
| Imine-CH3 (C) | Gly361 (O) | Carbon H-Bond | 3.36 | |||
| Imine (C) | Phe321 | Pi–Cation | 4.35 | |||
| L1 | Leu323 (C) | Pi–Sigma | 3.43 | |||
| L1 | Lys524 | Pi–Alkyl | 5.37 | |||
| Complex 14 | 567.97 | -4.43 | ||||
| L1 (N) | Phe321 | Pi–Cation | 4.60 | |||
| L1 | Leu323 | Pi–Alkyl | 4.98 | |||
| L1 | Lys524 | Pi–Alkyl | 5.38 | |||
| Complex 15 | 241.40 | -4.93 | ||||
| Azide (N) | Glu411 (OE2) | Charge–Charge | 4.91 | |||
| Azide (N) | Asp242 (OD2) | Charge–Charge | 4.03 | |||
| Azide (N) | Asp307 (OD1) | Charge–Charge | 5.09 | |||
| Azide (N) | Asp307 (OD1) | Charge–Charge | 4.08 | |||
| Azide (N) | Pro312 (O) | Conventional H-Bond | 2.82 | |||
| L1 (C) | His280 | Pi–Cation | 4.88 | |||
| L1 | Tyr158 | Pi–Pi Stacked | 4.65 | |||
| L1 | Lys156 | Pi–Alkyl | 5.34 | |||
| Complex 16 | 358.10 | -4.70 | ||||
| Azide (N) | Val404 (O) | Conventional H-Bond | 3.13 | |||
| Azide (N) | Tyr407 (O) | Conventional H-Bond | 3.03 | |||
| Imine (N) | Tyr416 | Pi–Cation | 3.75 | |||
| Imine (C) | Tyr416 | Pi–Cation | 3.33 | |||
| Imine-CH3 (C) | Tyr416 | Pi–Cation | 3.50 | |||
| L1 | Glu405 (OE2) | Pi–Anion | 4.06 | |||
| L1 | Val404 (C)/Glu405 (N) | Amide–Pi Stacked | 5.57 | |||
| L1 | Val404 | Pi–Alkyl | 5.50 | |||
| Genistein | 22.06 | -6.35 | [71] | |||
| C=O (benzopyran-4-one) | Ile272 | Conventional H-Bond | 2.73 | |||
| CO (5-hydroxy-benzopyran-4-one) | Ser298 | Conventional H-Bond | 2.45 | |||
| CO (7-hydroxy-benzopyran-4-one) | Leu297 | Conventional H-Bond | 3.29 | |||
| CO (7-hydroxy-benzopyran-4-one) | Ser291 | Conventional H-Bond | 2.88 | |||
| CO (7-hydroxy-benzopyran-4-one) | Ala292 | Conventional H-Bond | 3.14 | |||
| Phenyl | Arg263 | Pi–Alkyl | 4.19 | |||
| Phenyl | Ile272 | Pi–Alkyl | 5.25 |
Furthermore, the direct/indirect interactions with the enzyme active site change attaching to the coordination environment of complexes 12–16. The interactions of some parts of complexes with amino acid residues are given in Table 2, Figures 4–6 & Supplementary Figures S14 & S15. These interactions can be listed as the conventional hydrogen bond, pi–cation, pi–sigma, pi–alkyl and charge–charge. By virtue of the docking results of complexes 12, 13 and 15 exhibiting the best inhibition values against α-glucosidase among complexes, the interaction distance ranges from 2.53 to 4.92 Å emerged at the conventional hydrogen bond/pi-cation interactions between azide (N)/imine(C)/L1(C) and amino acid residues Phe321(O)/Phe321, Lys523(O)/Phe321 and Pro312(O)/His280 for complexes 12, 13 and 15, respectively (Table 2). Besides, the other remarkable interactions demonstrate the pi–sigma for complexes 12, 13, the pi–alkyl for complexes 13, 15, as well as the pi–pi stacked for complex 15. These interactions, having distance ranging from 3.34 to 5.37 Å, were observed between L1 with amino acid residues Leu323(CD1)/Leu323(C) for pi–sigma, Lys524/Lys156 for pi–alkyl, and Tyr158 for pi–pi stacked. In comparison these results of the genistein, the conventional hydrogen bond and pi–alkyl interactions obtained at 2.73/2.45/3.29/2.88/3.14 and 4.19/5.25 Å distances are remarkable for Ile272/Ser298/Leu297/Ser291/Ala292 and Arg263/Ile272 residues with the C=O (benzopyran-4-one)/CO (5-hydroxy-benzopyran-4-one)/CO (7-hydroxy-benzopyran-4-one) and phenyl ring, seen in Table 2. Other than these interactions, having amide pi–pi stacked and pi–anion for complex 16 appeared at the distances of 5.57 and 4.06 Å, were depicted between Val404(C)/Glu405(N) and Glu405(OE2) residues and the N-methyl-1-(pyridin-2-yl)methanimine (L1).
The 1,4-glycosidic bond in polysaccharides is hydrolyzed via catalyzed by α-glucosidase, resulting in the creation of α-D-glucose. As a result, during absorption, α-D-glucose is formed and enters the circulation, which raises postprandial hyperglycemia [23,25]. α-glucosidase contains Asp, Glu, Gly, His, Phe, Thr, Tyr, Leu, Lys, etc. residues in the structure. Additionally, especially Arg, Asp, Glu, Trp and Tyr residues are located at the entrance of the active site of α-glucosidase and they have been played critical for the enzyme's catalytic activity [95–97]. As can be seen in Table 2, Figures 4–6 & Supplementary Figures S14 & S15, the synthesized complexes interacted with some of these residues by conventional hydrogen bond, pi–cation, pi–sigma, pi–alkyl and charge–charge interactions. It is considered that these interactions with the active pocket of glucosidase can hinder the substrate from entering, decrease the catalytic activity, and finally cause the inhibition of glucosidase activity. Moreover, hydrogen bond interactions can decrease the hydrophilicity of α-glucosidase and rise its hydrophobicity that supports to increase the stability of the inhibitor–enzyme complex.
A similar trend was obtained between docking Ki values and in vitro inhibition results. It is concluded that the donor and acceptor properties in molecular systems may give rise to similarities and differences in localization, or delocalization of electron density in the molecule, as well as interactions with amino acid residues in the enzyme structure. Furthermore, by considering the enzyme kinetic studies [26,28,98], the majority of the metal complexes are non-competitive α-glucosidase inhibitors. Considering the literature, it is predicted that 12–16 complexes may be non-competitive inhibitors.
3.6. Molecular dynamic simulation
Molecular dynamic (MD) simulation calculations were performed for complexes 12–16 in ligands with the receptor using the GROMACS simulation package [99] via WebGro server [100]. Root mean square deviation (RMSD) profiles for protein–ligand complexes were assessed. The protein–ligand topology was generated by using Discovery Studio 4.0 program (Systèmes D. Dassault Systèmes Biovia; CA, USA) [67]. The MD simulation details are also given in Supplementary Material.
Considering that the stability of the system is inversely proportional to the amplitude of the fluctuations, the decrease in the stability of the systems is expressed by a higher RMSD value and higher fluctuations. It was determined that the α-glucosidase enzyme has an average RMSD of 0.24 nm. It appears to remain stable throughout the MD simulation, except for some large fluctuations between 20 and 30 ns (Supplementary Figure S16). The average RMSD of the α-glucosidase–compound-13 system of 0.25 nm was obtained. While some fluctuations were observed during 15–25 ns, it was observed that the MD simulation remained stable until 50 ns.
The mean RMSD values of α-glucosidase–compound-12/14/15 were calculated as 0.23, 0.26 and 0.31 nm, respectively. It was observed that α-glucosidase-12/14/15 compounds were stable during the MD simulation, except for a few minor/major deviations at 10–15, 15–20 and 10–20 ns, respectively. (Supplementary Figure S16). On the other hand, the average RMSD value of α-glucosidase–compound-16 was obtained as 0.31 nm. It showed high stability throughout the entire 50 ns MD simulation, while some minor fluctuations were observed during 10–15 and 35–45 ns. Supplementary Figure S16 shows the RMSD plots for all complexes. It is thought that the amino acids in α-glucosidase cause fluctuations during the MD simulation because they have different structural regions (such as helices, β sheets and coil) [25].
4. Conclusion
In vitro α-glucosidase inhibitions and docking studies were performed on the N-methyl-1-(pyridin-2-yl)methanimine (L1) and complexes 12–16 {[Cu2(L1)2(N3)3], (12), [Hg(L1)2(N3)·H2O] (13), [Cd(L1)2(N3)2], (14), [Ag(L1)2(N3)2·2H2O], (15), [Zn2(L1)2(N3)2Cl2·4H2O], (16)} synthesized and characterized by 1H and 13C NMR (except for complex 12 including Cu ion), mass (LC-HRMS) and FT–IR spectra. The electronic spectral behaviors were investigated by UV–Vis spectra. The IC50 values of complexes 12–16 were obtained at the range from 0.2802 ± 0.62 to 420.88 ± 1.42 μM. The complex 13 demonstrates the best α-glucosidase inhibitory activity, while the complex 14 has the lowest inhibitory activity. Besides, the IC50 values of complexes 12 (1.562 ± 0.83) and 15 (1.274 ± 0.32) were also found to be significant. These results were backed up by the docking ones presenting receptor–ligand interactions. Additionally, by considering a total of 50 ns MD simulations, the molecular dynamics study revealed that the compounds were stable during MD simulation. Furthermore, the detailed theoretical characteristic analyses were fulfilled using ωB97XD and CAM-B3LYP levels of density functional theory. The coordination geometries of complexes 12–16 were confirmed by NBO results appearing among lone pair electrons (n) of nitrogen atoms and metal ions. The high contribution to LUMO/LUMOs results shows that the L1 ligand has the strongest electron acceptor properties than the N3 ligand. It is deduced that while IC50 values of the 15 and 12 against α-glucosidase are obtained remarkably, the β and γ results that complexes 15 and 12 are highly efficient in terms of microscopic second- and third-NLO properties, respectively. Among Schiff base–azide metal complexes, for Type-2 diabetes, complex 13 may be an alternate anti-diabetic inhibitor according to in vitro/docking results, as well as for NLO material, complexes 12 and 15 may be recommended as a candidate.
Supplementary Material
Funding Statement
The work was supported by the Scientific Research Projects Unit of Sakarya University (2021-7-25-103).
Supplemental material
Supplemental data for this article can be accessed at https://doi.org/10.1080/17568919.2024.2342650
Author contributions
D Avcı: methodology, software, formal analysis, writing-original draft, writing review and editing. Ö Özgea: data curation, formal analysis, visualization. F Sönmezc: formal analysis, investigation, writing, review and editing. Ö Tamera: investigation, software. A Başoğlua: methodology, software. Y Atalaya: methodology, software. BZ Kurt: formal analysis.
Financial disclosure
The work was supported by the Scientific Research Projects Unit of Sakarya University (2021-7-25-103). The authors have no other relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript apart from those disclosed.
Competing interests disclosure
The authors have no competing interests or relevant affiliations with any organization or entity with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert testimony, grants or patents received or pending, or royalties.
Writing disclosure
No writing assistance was utilized in the production of this manuscript.
Data availability statement
The data that support the findings of this study are available from the corresponding author upon reasonable request.
References
Papers of special note have been highlighted as: • of interest; •• of considerable interest
- 1.Song Y, Massera C, Roubeau O, et al. An unusual open cubane structure in a μ1,1-azido- and alkoxo-bridged tetranuclear copper(II) complex, [Cu4L2(μ1,1-N3)2]·5H2O (H3L = N,N′-(2-hydroxylpropane-1,3-diyl)bis-salicylideneimine). Inorg. Chem. 2004;43:6842–6847. doi: 10.1021/ic049317g [DOI] [PubMed] [Google Scholar]
- 2.Liu X, Yang Q, Su Z, et al. 3D high-energy-density and low sensitivity materials: synthesis, structure and physicochemical properties of an azide–Cu(II) complex with 3,5-dinitrobenzoic acid. RSC Adv. 2014;4:16087–16093. doi: 10.1039/C4RA00635F [DOI] [Google Scholar]
- 3.Charlot M-F, Kahn O, Chaillet M, et al. Interaction between copper(II) ions through the Azido bridge: concept of spin polarization and ab initio calculations on model systems. J. Am. Chem. Soc. 1986;108:2574–2581. doi: 10.1021/ja00270a014 [DOI] [Google Scholar]
- 4.Lehleh A, Beghidja A, Beghidja C, et al. Synthesis, crystal structure and thermal decomposition of Cu(II), Co(II), Mn(II) complexes with hetero-ligands containing cysteic acid, 4,4′-dimethyl-2,2′-bipyridine and azide. C R Chimie. 2011;14:462–470. doi: 10.1016/j.crci.2010.12.002 [DOI] [Google Scholar]
- 5.Montazerozohori M, Mojahedi Jahromi S, Masoudiasl A, et al. Nano structure zinc (II) Schiff base complexes of a N3-tridentate ligand as new biological active agents: spectral, thermal behaviors and crystal structure of zinc azide complex. Spectrochim Acta A. 2015;138:517–528. doi: 10.1016/j.saa.2014.11.055 [DOI] [PubMed] [Google Scholar]
- 6.Wang Li-Y, Zhao B, Zhang Chen-X, et al. The first azide(μ1,1)-bridged binuclear cobalt(II)-imino nitroxide complex with ferromagnetic behavior. Inorg. Chem. 2003;42:5804–5806. doi: 10.1021/ic034710s [DOI] [PubMed] [Google Scholar]
- 7.Pastor-Ramírez C, Zamorano-Ulloa R, Vazquez-Lima H, et al. Synthesis and magneto-structural study of distorted octahedral CoII complexes with picolinamide and azide (μ1,1)-bridge. ChemistrySelect. 2021;6:1394–1403. doi: 10.1002/slct.202004323 [DOI] [Google Scholar]
- 8.Wang Y-Q, Yue Q, Qi Y, et al. Manganese (II), iron (II), and mixed-metal metal–organic frameworks based on chains with mixed carboxylate and azide bridges: magnetic coupling and slow relaxation. Inorg. Chem. 2013;52:4259–4268. doi: 10.1021/ic302162c [DOI] [PubMed] [Google Scholar]
- 9.Hou J, Liu X, Shen J, et al. The impact of click chemistry in medicinal chemistry. Expert Opin. Drug Discov. 2012;7(6):489–501. doi: 10.1517/17460441.2012.682725 [DOI] [PubMed] [Google Scholar]
- 10.Gao P, Sun L, Zhou J, et al. Discovery of novel anti-HIV agents via Cu(I)-catalyzed azide-alkyne cycloaddition (CuAAC) click chemistry-based approach. Expert Opin. Drug Discov. 2016;11(9):857–871. doi: 10.1080/17460441.2016.1210125 [DOI] [PubMed] [Google Scholar]
- 11.Villanueva M, Mesa JL, Urtiaga MK, et al. A new alternating ferro- and antiferromagnetic, one-dimensional manganese(II) azide complex, [Mn(dpa)(N3)2]2 crystal structure and spectroscopic and magnetic properties. Eur. J. Inorg. Chem. 2001;1581–1586. doi: [DOI] [Google Scholar]
- 12.Cui G, Li Y, Shi T, et al. Synthesis and characterization of Eu(III) complexes of modified cellulose and poly(N-isopropylacrylamide). Carbohydr. Polym. 2013;94:77–81. doi: 10.1016/j.carbpol.2013.01.045 [DOI] [PubMed] [Google Scholar]
- 13.Zhang J, Wang L, Zhong A, et al. Deep red PhOLED from dimeric salophen Platinum(II) complexes. Dyes Pigm. 2019;162:590–598. doi: 10.1016/j.dyepig.2018.10.053 [DOI] [Google Scholar]
- 14.Zhao Y, Chen S, Shen F, et al. In vitro neutralization of autocrine IL-10 affects Op18/stathmin signaling in non-small cell lung cancer cells. Oncol. Lett. 2019;41:501–511. doi: 10.3892/or.2018.6795 [DOI] [PubMed] [Google Scholar]
- 15.Wei S, Sun T, Du J, et al. Xanthohumol, a prenylated flavonoid from Hops, exerts anticancer effects against gastric cancer in vitro. Oncol. Lett. 2018;40:3213–3222. doi: 10.3892/or.2018.6723 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Yim CB, Dijkgraaf I, Merkx R, et al. Synthesis of DOTA-conjugated multimeric [Tyr3]octreotide peptides via a combination of Cu(I)-catalyzed “click” cycloaddition and thio acid/sulfonyl azide “sulfo-click” amidation and their in vivo evaluation. J. Med. Chem. 2010;53:3944–3953. doi: 10.1021/jm100246m [DOI] [PubMed] [Google Scholar]
- 17.Thamilarasan V, Jayamani A, Sengottuvelan N. Synthesis, molecular structure, biological properties and molecular docking studies on MnII, CoII and ZnII complexes containing bipyridine–azide ligands. Eur. J. Med. Chem. 2015;89(7):266–278. doi: 10.1016/j.ejmech.2014.09.073 [DOI] [PubMed] [Google Scholar]
- 18.Jiang X, Yan M. Comparing the impact on the prognosis of acute myocardial infarction critical patients of using midazolam, propofol, and dexmedetomidine for sedation. BMC Cardiovasc. Disord. 2021;21:584. doi: 10.1186/s12872-021-02385-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Cao J, Chen C, Wang Y, et al. Influence of autologous dendritic cells on cytokine-induced killer cell proliferation, cell phenotype and antitumor activity in vitro. Oncol. Lett. 2016;12:2033–2037. doi: 10.3892/ol.2016.4839 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Balan K, Ratha P, Prakash G, et al. Evaluation of in vitro α-amylase and α-glucosidase inhibitory potential of N2O2 schiff base Zn complex. Arab J. Chem. 2017;10:732–738. doi: 10.1016/j.arabjc.2014.07.002 [DOI] [Google Scholar]; • In the study titled evaluation of the in vitro α-glucosidase inhibitory potential of Schiff base Zn complex containing N2O2, the IC50 value for α-glucosidase inhibition was found to be 0.23 mM.
- 21.Tripath IP, Kumar MM, Ruchita T, et al. Synthesis, spectral, electrochemical analysis and screening for α-glucosidase inhibition of some complexes of cobalt (II) and ethylenediamine. Res. J. Chem. Sci. 2014;4(6):13–17. [Google Scholar]; • By performing synthesis, spectral, electrochemical analysis, and α-glucosidase inhibition study of some Co(II) complexes containing ethylenediamine, α-glucosidase inhibition values for [Co(en)3]·SO4 and [Co(en)3]·2Cl complexes were reported as 927.89 μM and 784.12 μM in low level.
- 22.Miyazaki R, Yasui H, Yoshikawa Y. α-glucosidase inhibition by new Schiff base complexes of Zn(II). Open. J. Inorg. Chem. 2016;6(2):114–124. doi: 10.4236/ojic.2016.62007 [DOI] [Google Scholar]
- 23.Raju BC, Ashok AK, Kumar JA, et al. α-Glucosidase inhibitory antihyperglycemic activity of substituted chromenone derivatives. Bioorg. Med. Chem. 2010;18(1):358–365. doi: 10.1016/j.bmc.2009.10.047 [DOI] [PubMed] [Google Scholar]
- 24.Yao Y, Sang W, Zhou M, et al. Antioxidant and α-glucosidase inhibitory activity of colored grains in China. J. Agric. Food Chem. 2010;58(2):770–774. doi: 10.1021/jf903234c [DOI] [PubMed] [Google Scholar]
- 25.Khan M, Ahad G, Alam A, et al. Synthesis of new bis(dimethylamino)benzophenone hydrazone for diabetic management: in-vitro and in-silico approach. Heliyon. 2024;10:e23323. doi: 10.1016/j.heliyon.2023.e23323 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Zheng J, Ma L. Silver(I) complexes of 2,4-dihydroxybenzaldehyde–amino acid Schiff bases-Novel noncompetitive a-glucosidase inhibitors. Bioorg. Med. Chem. Lett. 2015;25:2156–2161. doi: 10.1016/j.bmcl.2015.03.078 [DOI] [PubMed] [Google Scholar]; •• A series of silver(I) complexes containing Schiff bases were designed and tested for a-glucosidase inhibition, and all the silver complexes showed strong inhibitory activity at the micromolar level.
- 27.Deswal Y, Asija S, Dubey A, et al. Cobalt(II), nickel(II), copper(II) and zinc(II) complexes of thiadiazole based Schiff base ligands: synthesis, structural characterization, DFT, antidiabetic and molecular docking studies. J. Mol. Struct. 2022;1253:132266. doi: 10.1016/j.molstruc.2021.132266 [DOI] [Google Scholar]; •• The findings of this study served as the foundation for our investigation, and it also aided in the target compounds' chemical synthesis.
- 28.Zheng JW, Ma L. Assessment of silver(I) complexes of salicylaldehyde derivatives-histidine Schiff base as novel α-glucosidase inhibitors. Chin. Chem. Lett. 2016;27(2):283–286. doi: 10.1016/j.cclet.2015.11.015 [DOI] [Google Scholar]; •• The findings of this study served as the foundation for our investigation, and it also aided in the target compounds' chemical synthesis.
- 29.Yoshikawa Y, Hirata R, Yasui H, et al. Alpha-glucosidase inhibitory effect of anti-diabetic metal ions and their complexes. Biochimie 2009;91:1339–1341. doi: 10.1016/j.biochi.2009.06.005 [DOI] [PubMed] [Google Scholar]
- 30.Avcı D, Altürk S, Sönmez F, et al. Three novel Cu(II), Cd(II) and Cr(III) complexes of 6-Methylpyridine-2-carboxylic acid with thiocyanate: synthesis, crystal structures, DFT calculations, molecular docking and α-Glucosidase inhibition studies. Tetrahedron. 2018;74:7198–7208. doi: 10.1016/j.tet.2018.10.054 [DOI] [Google Scholar]
- 31.Avcı D, Altürk S, Sönmez F, et al. A novel series of mixed-ligand M(II) complexes containing 2,2′-bipyridyl as potent α-glucosidase inhibitor: synthesis, crystal structure, DFT calculations, and molecular docking. JBIC J. Biol. Inorg. Chem. 2019;24:747–764. doi: 10.1007/s00775-019-01688-9 [DOI] [PubMed] [Google Scholar]; •• The rationale for our study was based on comparison with the results of this study; additionally, this study found that the target compounds were more effective against α-glucosidase.
- 32.Avcı D, Altürk S, Sönmez F, et al. Synthesis, spectral properties, in vitro α-glucosidase inhibitory activity and quantum chemical calculations of novel mixed-ligand M(II) complexes containing 1,10-phenanthroline. Appl. Organomet. Chem. 2020;34:e5412. doi: 10.1002/aoc.5412 [DOI] [Google Scholar]
- 33.Gollapalli M, Taha M, Javid MT, et al. Synthesis of benzothiazole derivatives as a potent α-glucosidase inhibitor. Bioorg. Chem. 2019;85:33–48. doi: 10.1016/j.bioorg.2018.12.021 [DOI] [PubMed] [Google Scholar]
- 34.Ghani U. Re-exploring promising a-glucosidase inhibitors for potential development into oral anti-diabetic drugs: finding needle in the haystack. Eur. J. Med. Chem. 2015;103:133–162. doi: 10.1016/j.ejmech.2015.08.043 [DOI] [PubMed] [Google Scholar]
- 35.Santos CM, Freitas M, Fernandes E. A comprehensive review on xanthone derivatives as α-glucosidase inhibitors. Eur. J. Med. Chem. 2018;157:1460–1479. doi: 10.1016/j.ejmech.2018.07.073 [DOI] [PubMed] [Google Scholar]
- 36.Sun H, Ding W, Song X, et al. Synthesis of 6-hydroxyaurone analogues and evaluation of their α-glucosidase inhibitory and glucose consumption-promoting activity: development of highly active 5,6-disubstituted derivatives. Bioorg. Med. Chem. Lett. 2017;27(15):3226–3230. doi: 10.1016/j.bmcl.2017.06.040 [DOI] [PubMed] [Google Scholar]
- 37.Wang G, Peng Z, Wang J, et al. Synthesis, in vitro evaluation and molecular docking studies of novel triazine-triazole derivatives as potential α-glucosidase inhibitors. Eur. J. Med. Chem. 2017;125:423–429. doi: 10.1016/j.ejmech.2016.09.067 [DOI] [PubMed] [Google Scholar]
- 38.Fan M, Yang W, Peng Z, et al. Chromone-based benzohydrazide derivatives as potential α-glucosidase inhibitor: synthesis, biological evaluation and molecular docking study. Bioorg. Chem. 2023;131:106276. doi: 10.1016/j.bioorg.2022.106276 [DOI] [PubMed] [Google Scholar]
- 39.Wang G, Chen M, Wang J, et al. Synthesis, biological evaluation and molecular docking studies of chromone hydrazone derivatives as α-glucosidase inhibitors. Bioorg. Med. Chem. Lett. 2017;27:2957–2961. doi: 10.1016/j.bmcl.2017.05.007 [DOI] [PubMed] [Google Scholar]
- 40.Avcı D, Altürk S, Sönmez F, et al. A new dinuclear copper (II) complex of 2,5–Furandicarboxyclic acid with 4(5)–Methylimidazole as a high potential α-glucosidase inhibitor: Synthesis, Crystal structure, Cytotoxicity study, and TD/DFT calculations. Appl. Organomet. Chem. 2019;33(3):e4725. doi: 10.1002/aoc.4725 [DOI] [Google Scholar]
- 41.Cui G, Zhao K, You K, et al. Synthesis and characterization of phenylboronic acid-containing polymer for glucose-triggered drug delivery+. Sci. Technol. Adv. Mater. 2020;21:1–10. doi: 10.1080/14686996.2019.1700394 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Cui G, Bai Y, Li W, et al. Synthesis and characterization of Eu(III) complexes of modified D-glucosamine and poly(N-isopropylacrylamide). Mater. Sci. Eng. C. 2017;78:603–608. doi: 10.1016/j.msec.2017.03.059 [DOI] [PubMed] [Google Scholar]
- 43.Zhao J, Liu Y, Zhu L, et al. Tumor cell membrane-coated continuous electrochemical sensor for GLUT1 inhibitor screening. J. Pharm. Anal. 2023;13:673–682. doi: 10.1016/j.jpha.2023.04.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Jing Y, Li L, Zhu-jun Y, et al. Polypeptide from Moschus suppresses lipopolysaccharide-induced inflammation by inhibiting NF-κ B-ROS/NLRP3 pathway*. Chin. J. Integr. Med. 2023;29(10):895–904. doi: 10.1007/s11655-023-3598-z [DOI] [PubMed] [Google Scholar]
- 45.Sohrabi M, Binaeizadeh MR, Iraji A, et al. A review on a-glucosidase inhibitory activity of first row transition metal complexes: a futuristic strategy for treatment of Type 2 diabetes. RSC Adv. 2022;12:12011–12052. doi: 10.1039/D2RA00067A [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Wei Zeng MD, Lu Yu MD, Jiarui Wu MB, et al. Clinical characteristics and long-term follow-up outcomes of myelin oligodendrocyte glycoprotein antibody-associated disease in Han Chinese participants. Medicine 2023;102:40. doi: 10.1097/MD.0000000000035391 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Gul S, Jan F, Alam A, et al. Synthesis, molecular docking and DFT analysis of novel bis-Schiff base derivatives with thiobarbituric acid for α-glucosidase inhibition assessment. Sci Rep. 2024;14:3419. doi: 10.1038/s41598-024-54021-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Ahmad R, Alam A, Khan M, et al. Antioxidant activity, molecular docking and quantum studies of new Bis-Schiff bases based on benzyl phenyl ketone moiety. ChemistrySelect. 2023;8:e202302338. doi: 10.1002/slct.202302338 [DOI] [Google Scholar]
- 49.Ahmad R, Khan M, Alam A, et al. Synthesis, molecular structure and urease inhibitory activity of novel bis-Schiff bases of benzyl phenyl ketone: a combined theoretical and experimental approach. Saudi. Pharm. J. 2023;31:101688. doi: 10.1016/j.jsps.2023.06.021 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Waheed M, Idris S, Jan F, et al. Exploring the synthesis, structure, spectroscopy and biological activities of novel 4-benzylidene-1-(2-(2,4-dichloro phenyl)acetyl) thiosemicarbazide derivatives: an integrated experimental and theoretical investigation. Saudi Pharm. J. 2023;31:101874. doi: 10.1016/j.jsps.2023.101874 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Li B, Wang W, Zhao L, et al. Aggregation-Induced Emission-Based Macrophage-Like Nanoparticles for Targeted Photothermal Therapy and Virus Transmission Blockage in Monkeypox. Adv. Mater. 2024;36:2305378. doi: 10.1002/adma.202305378 [DOI] [PubMed] [Google Scholar]
- 52.Zhang J, Zhong A, Huang G, et al. Enhanced efficiency with CDCA co-adsorption for dye-sensitized solar cells based on metallosalophen complexes. Solar Energy. 2020;209:316–324. doi: 10.1016/j.solener.2020.08.096 [DOI] [Google Scholar]
- 53.Idris S, Jan F, Waheed M, et al. Multifaceted bioactivity of thiosemicarbazide derivatives: synthesis, characterization, and DFT investigations on inhibition of α-amylase, hydroxyl radical scavenging, and iron chelating activities with molecular docking insights. J Mol Struct. 2024;1304:137669. doi: 10.1016/j.molstruc.2024.137669 [DOI] [Google Scholar]
- 54.Yanai T, Tew DP, Handy NC. A new hybrid Exchange-correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem Phys Lett. 2004;393:51–57. doi: 10.1016/j.cplett.2004.06.011 [DOI] [Google Scholar]
- 55.Chai J-D, Head-Gordon M. Systematic optimization of long-range corrected hybrid density functionals. J Chem Phys. 2008;128:084106. doi: 10.1063/1.2834918 [DOI] [PubMed] [Google Scholar]
- 56.Chai J-D, Head-Gordon M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys Chem Chem Phys. 2008;10:6615–6620. doi: 10.1039/B810189B [DOI] [PubMed] [Google Scholar]
- 57.Frisch MJ, Trucks GW, Schlegel HB, et al. Gaussian 16. Revision C.01. Wallingford (CT): Gaussian, Inc.; 2016. [Google Scholar]
- 58.Dennington R, Keith TA, Millam JM. GaussView. Version 6.1. Shawnee Mission (KS); Semichem Inc.; 2016. [Google Scholar]
- 59.Frisch MJ, Pople JA, Binkley JS. Self-consistent molecular orbital methods 25. Supplementary functions for Gaussian basis set. J Chem Phys. 1984;80:3265–3269. doi: 10.1063/1.447079 [DOI] [Google Scholar]
- 60.Hay PY, Wadt WR. Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J Chem Phys. 1985;82:270–283. doi: 10.1063/1.448799 [DOI] [Google Scholar]
- 61.Runge E, Gross EKU. Density-functional theory for time-dependent systems. Phys Rev Lett. 1984;52:997–1000. doi: 10.1103/PhysRevLett.52.997 [DOI] [Google Scholar]
- 62.Miertus S, Scrocco E, Tomasi J. Electrostatic interaction of a solute with a continuum. A direct utilizaion of Ab initiomolecular potentials for the prevision of solvent effects. J Chem Phys. 1981;55:117–129. doi: 10.1016/0301-0104(81)85090-2 [DOI] [Google Scholar]
- 63.Gorelsky SI. SWizard Program. Revision 4.5. Canada; University of Ottawa; 2010. Available from: http://www.sg.chem.net/ [Google Scholar]
- 64.Skripnikov L. Chemissian: software to analyze spectra, build density maps and molecular orbitals. Version 4.43. 2016. Available from: https://www.chemissian.com/
- 65.Morris GM, Huey R, Lindstrom W, et al. Autodock4 and AutoDockTools4: automated docking with selective receptor flexibility. J Comput Chem. 2009;30:2785–2791. doi: 10.1002/jcc.21256 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Sanner MF. Python: a programming language for software integration and development. J Mol Graphics Mod. 1999;17:57–61. [PubMed] [Google Scholar]
- 67.Systèmes D. Dassault Systèmes Biovia. San Diego (CA) USA; 2016. [Google Scholar]
- 68.Manca G, Cano J, Ruiz E. Exchange interactions in azido-bridged ligand NiII complexes: a theoretical analysis. Inorg Chem. 2009;48(7):3139–3144. doi: 10.1021/ic802397k [DOI] [PubMed] [Google Scholar]
- 69.Amirnasr M, Khalaji AD, Falvello LR, et al. Synthesis, characterization, and crystal structures of [Ag(ca2en)(PPh3)(X)] (X = N3, SCN) complexes. Polyhedron. 2006;25:1967–1970. doi: 10.1016/j.poly.2005.12.021 [DOI] [Google Scholar]
- 70.Wu H, Wang F, Shi F, et al. Three- and four-coordinate Ag(I) complexes of crotonate and bis(benzimidazole)-2-oxapropane ligands: syntheses, crystal structures, DNA-binding studies and antioxidant activities. Transit Met Chem. 2015;40:555–564. doi: 10.1007/s11243-015-9948-8 [DOI] [Google Scholar]
- 71.Özge Ö, Avcı D, Sönmez F, et al. Synthesis, DFT calculations, α-glucosidase inhibitor activity, and docking studies on Schiff base metal complexes containing isothiocyanate. Appl Organomet Chem. 2023;37:e7084. doi: 10.1002/aoc.7084 [DOI] [Google Scholar]; •• The rationale for our study was based on comparison with the results of this study; additionally, this study found that the target compounds were more effective against α-glucosidase.
- 72.Weinhold F, Landis C. Valency and bonding: A natural bond orbital donor– acceptor perspective. Cambridge, UK: Cambridge University Press; 2005. [Google Scholar]
- 73.Altürk S, Avcı D, Tamer Ö, et al. A cobalt(II) complex with 6-methylpicolinate: synthesis, characterization, second- and third-order nonlinear optical properties, and DFT calculations. J Phys Chem Solids. 2016;98:71–80. doi: 10.1016/j.jpcs.2016.06.008 [DOI] [Google Scholar]
- 74.Reed AE, Weinhold F. Natural bond orbital analysis of near-Hartree–Fock water dimer. J Chem Phys. 1983;78:4066–4073. doi: 10.1063/1.445134 [DOI] [Google Scholar]
- 75.Glendening ED, Reed AE, Carpenter JE, et al. NBO Version 3.1. TCI. Madison: University of Wisconsin; 1998. [Google Scholar]
- 76.Foster JP, Weinhold F. Natural hybrid orbitals. J Am Chem Soc. 1980;102:7211–7218. doi: 10.1021/ja00544a007 [DOI] [Google Scholar]
- 77.Gorelsky SI, Lever ABP. The electronic structure and spectra of [Ru(NH3)4(LL)]2+ (LL = bpy, bpz, bqdi) studied by density functional theory and INDO/S. Charge transfer character of electronic transitions and their solvatochromism. Can J Anal Sci Spectr. 2003;48(1):93–105. [Google Scholar]
- 78.Luque FJ, Orozco M, Bhadane PK, et al. SCRF calculation of the effect of water on the topology of the molecular electrostatic potential. J Phys Chem. 1993;97:9380–9384. doi: 10.1021/j100139a021 [DOI] [Google Scholar]
- 79.Pearson RG. Absolute electronegativity and hardness correlated with molecular orbital theory. Proc Nati Acad Sci USA. 1986;83:8440–8441. doi: 10.1073/pnas.83.22.8440 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Avcı D, Özge Ö, Başoğlu A, et al. Synthesis, crystal structures, and DFT calculations: novel Mn(II), Co(II) and Ni(II) complexes of N-(pyridin-2-ylmethylene)methanamine with isothiocyanate as promising optical materials. Opt Quantum Electron. 2023;55:408. doi: 10.1007/s11082-023-04668-1 [DOI] [Google Scholar]
- 81.Oudar JL, Zyss J. Relations between microscopic and macroscopic lowest-order optical nonlinearities of molecular crystals with one-or two-dimensional units. Phys Rev A. 1982;26:2028–2048. doi: 10.1103/PhysRevA.26.2028 [DOI] [PubMed] [Google Scholar]
- 82.Pierce BM. A theoretical analysis of third-order nonlinear optical properties of linear polyenes and benzene. J Chem Phys. 1989;91(2):791–811. doi: 10.1063/1.457132 [DOI] [Google Scholar]
- 83.Avcı D. Second and third-order nonlinear optical properties and molecular parameters of azo chromophores: semiempirical analysis. Spectrochim Acta A. 2011;82:37–43. doi: 10.1016/j.saa.2011.06.037 [DOI] [PubMed] [Google Scholar]
- 84.Chemla DS, Zysss J. Nonlinear Optical Properties of Organic Molecules and Crystals. London: Academic Press; 1987. [Google Scholar]
- 85.Faizan M, Mehkoom M, Afroz Z, et al. Experimental and computational investigation of novel dihydrated organic single crystal of 2,4,6-triaminopyrimidine and 3,5-dintrobenzoic acid: linear and nonlinear optical response with limiting performance. J Solid State Chem. 2021;300:122255. doi: 10.1016/j.jssc.2021.122255 [DOI] [Google Scholar]
- 86.Kamaal S, Mehkoom M, Ali A, et al. Potential third-order nonlinear optical response facilitated by intramolecular charge transfer in a simple Schiff base molecule: experimental and theoretical exploration. ACS Omega. 2021;6:6185–6194. doi: 10.1021/acsomega.0c05557 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Mehkoom M, Afzal SM, Ahmad S, et al. The new pyrazoline derivative 5-(3,4-Dimethoxyphenyl)-3-(2,5-dimethyl-thiophene-3-yl),-4,5-dihydro-pyrazole-1-carbothioic acid amide (DDPA) as an advisable candidate for optical linearity, nonlinearity, and limiting performance. J Mol Liq. 2022;345:117018. doi: 10.1016/j.molliq.2021.117018 [DOI] [Google Scholar]
- 88.Cheng LT, Tam W, Stevenson SH, et al. Experimental investigations of organic molecular nonlinear optical polarizabilities. 1. Methods and results on benzene and stilbene derivatives. J Phys Chem. 1991;95:10631–10643. doi: 10.1021/j100179a026 [DOI] [Google Scholar]
- 89.Ledoux I, Zyss J. Influence of the molecular environment in solution measurements of the second-order optical susceptibility for urea and derivatives. Chem Phys. 1982;73:203–213. doi: 10.1016/0301-0104(82)85161-6 [DOI] [Google Scholar]
- 90.Adant C, Dupuis M, Bredas J. Ab initio study of the nonlinear optical properties of urea: electron correlation and dispersion effects. Int J Quant Chem. 1995;56:497–507. doi: 10.1002/qua.560560853 [DOI] [Google Scholar]
- 91.Yoshikawa Y, Ueda E, Kawabe K, et al. Development of new insulinomimetic zinc(II) picolinate complexes with a Zn(N2O2) coordination mode: structure characterization, in vitro, and in vivo studies. JBIC J Biol Inorg Chem. 2002;7:68–73. doi: 10.1007/s007750100266 [DOI] [PubMed] [Google Scholar]
- 92.Yasumatsu N, Yoshikawa Y, Adachi Y, et al. Antidiabetic copper (II)-picolinate: impact of the first transition metal in the metallopicolinate complexes. Bioorg Med Chem. 2007;15:4917–4922. doi: 10.1016/j.bmc.2007.04.062 [DOI] [PubMed] [Google Scholar]
- 93.Avcı D, Altürk S, Sönmez F, et al. Synthesis, DFT calculations and molecular docking study of mixed ligand metal complexes containing 4,4′-dimethyl-2,2′-bipyridyl as α-glucosidase inhibitors. J Mol Struct. 2020;1205:127665. doi: 10.1016/j.molstruc.2019.127655 [DOI] [Google Scholar]
- 94.Avcı D, Altürk S, Sönmez F, et al. Novel metal complexes containing 6-methylpyridine-2-carboxylic acid as potent α-glucosidase inhibitor: synthesis, crystal structures, DFT calculations, and molecular docking. Mol Divers. 2021;25:171–189. doi: 10.1007/s11030-020-10037-x [DOI] [PubMed] [Google Scholar]
- 95.Hermans MM, Kroos MA, van Beeumen J, et al. Human lysosomal a-glucosidase characterization of the catalytic site. J Biol Chem. 1991;266(21):13507–13512. doi: 10.1016/S0021-9258(18)92727-4 [DOI] [PubMed] [Google Scholar]
- 96.Rye CS, Withers SG. Glycosidase mechanisms. Curr Opin Chem Biol. 2000;4:573–580. doi: 10.1016/S1367-5931(00)00135-6 [DOI] [PubMed] [Google Scholar]
- 97.Dong Q, Hu N, Yue H, et al. Inhibitory activity and mechanism investigation of hypericin as a novel α-glucosidase inhibitor. Molecules. 2021;26:4566. doi: 10.3390/molecules26154566 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Wang Y, Ma L, Li Z, et al. Synergetic inhibition of metal ions and genistein on α-glucosidase. FEBS Lett. 2004;576:46–50. doi: 10.1016/j.febslet.2004.08.059 [DOI] [PubMed] [Google Scholar]
- 99.Abraham MJ, Murtola T, Schulz R, et al. GROMACS: high performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX. 2015;1:19–25. doi: 10.1016/j.softx.2015.06.001 [DOI] [Google Scholar]
- 100.WebGRO for Macromolecular Simulations. Available from: https://simlab.uams.edu/
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data that support the findings of this study are available from the corresponding author upon reasonable request.

