DOI:
10.1039/D4NJ03197K
(Paper)
New J. Chem., 2024,
48, 15747-15759
Synthesis, crystal structure, thermal analysis, spectroscopic, optical polarizability, and DFT studies, and molecular docking approaches of novel 2-methyl-benzylammonium derivatives for potential anti-inflammatory control†
Received
17th July 2024
, Accepted 21st August 2024
First published on 3rd September 2024
Abstract
This work aims to investigate novel structures based on 2-methylbenzylamine cations [(C8H12N)2Co(SCN)4] (1) and [(C8H12N)SCN] (2). The novel complexes were characterized and investigated by various techniques such as differential thermogravimetry analysis, FT-IR, UV-visible spectroscopy, impedance complex analysis, molecular modeling based on DFT calculations, and molecular docking as potent anti-inflammatory agents. Based on the reported results of these characterization tools, the desired complex phases were confirmed. These novel compounds were characterized by FTIR analysis, which supported the presence of surface ligand groups of thiocyanates, and UV-visible spectroscopy showed the optical transparencies of the titled compounds in addition to the confirmation of the electronic transition. Complex packing occurred through the N–H⋯S and N–H⋯N H-bond interactions, forming a ring in addition to CH-interactions, resulting in a 3-D network. Finally, molecular docking occurred for both complexes, which suggests that the complexes have anti-inflammatory potential.
1. Introduction
Cobalt complexes are of great interest due to their diverse architectures, which include a wide variety of assemblies due to their extended configurations such as turbinated networks,1 molecular drawstrings, chains,2 square grids,3 and stairway frameworks.4–8 Research has been conducted on the design and synthesis of novel compounds based on tricyanic cobaltate, which has a general chemical formula of C2Co(SCN)4. The aim of this contribution is twofold: firstly, to investigate how the organic cation (C) can add more interesting properties to the structure, such as protonic conductivity,9–12 ferroelectricity,13 nonlinear optics, and luminescence,14,15 and secondly, how it can control the inorganic chain [Co(SCN)4]n− packing.16–18 Many antibiotic compounds are either derivatives of known drugs or their metal complexes, underscoring the importance of coordinating ligands. The biological efficacy of these compounds depends not only on the choice of metal but also on the bioactive moiety. Thus, a variety of bioactive ligands have been employed to assess the biological potential of their metal complexes.19,20 Transition metal complexes from the 3d and 4d series have shown significant interactions with large biological molecules like DNA, RNA, and proteins, displaying strong binding abilities. These complexes have demonstrated promising antimicrobial and anticancer properties.21,22 Medicinal chemists are increasingly interested in the creation of novel metal-based chemotherapy compounds. Numerous research studies have explored the biological effects of these metal complexes, such as their abilities to combat cancer, bacteria, fungi, and oxidative stress. Some investigations have focused on the synergy between antibiotics and transition metals to improve antibacterial efficacy.
In this present work, single crystals of [(C8H12N)2Co(SCN)4] and [(C8H12N)CNS] have been synthesized by room-temperature evaporation. The crystal structures have been determined by X-ray diffraction. Both complexes have been investigated by spectroscopic and optical techniques and characterized using DFT setups. The thermal behaviors were studied and showed phase transition and decomposition in both complexes; additionally, molecular docking suggests that the complexes may act as anti-inflammatory agents.
2. Experimental details
2.1. Synthesis
2.1.1. Preparation of [(C8H12N)2Co(SCN)4].
All of the experimental reagents were used without any further purification. The synthesis of [(C8H12N)2Co(SCN)4] requires a prior preparation of HSCN. To (2 g, 10 mmol) of cobalt chloride, 20 mL of HSCN acid (2 M) was added. Then, 5 mL of a dilute solution of 2-methyl benzylamine (C8H11N) (2 M) was added dropwise with constant stirring. The final blue solution was left at room temperature. After a few days, blue block-shaped crystals of [(C8H12N)2Co(SCN)4] appeared in the solution.
2.1.2. Preparation of [(C8H12N)CNS].
The title compound [(C8H12N)CNS] has been synthesized by reacting an aqueous solution of 2 methylbenzylamine with thiocyanate (HSCN) at 75–85 °C, followed by extraction of [(C8H12N)CNS] with ethanol, and then recrystallization using ethanol as a recrystallization solvent.
2.2. Structure analysis using X-rays
Single-crystal DRX data were collected on a Bruker-Nonius Kappa CCD diffractometer using Mo-Kα radiation (170 K, λ = 0.71073 A). The crystal structure was solved by direct methods and refined by the full-matrix methods based on F2 using the SHELXL program23 included in the WINGX software package.24 The molecular structures were drawn using Diamond 3.1 supplied by Crystal Impact.25 Crystal data and details of the structure refinement are given in Table 1.
Table 1 Synthesis conditions and crystallographic data for [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN]
Formula |
[(C8H12N)2Co(SCN)4] |
[(C8H12N)SCN] |
Where w = 1/[σ2(Fo2) + (0.0215P)2 + 1.7725P] and P = (Fo2 + 2Fc2)/3. |
System |
Monoclinic |
Triclinic |
Space group |
C2/c |
P |
a, b, c (Å) |
18.6374(5), 5.0271(1), 28.1614(8) |
6.5707(3), 10.0248(5), 14.7305(8) |
α, β, γ (°) |
108.374(2) |
100.787(2), 91.137(3), 90.534(3) |
V (Å3) |
2503.99(11) |
952.88(8) |
D
x
(Mg m−3) |
1.519 |
1.257 |
Z
|
4 |
4 |
M
r
(g mol−1) |
535.62 |
180.27 |
T (K) |
170 |
170 |
θ
max, θmin (°) |
28.4, 2.3 |
28.3,0.4 |
μ (mm−1) |
1.04 |
0.29 |
Color |
Needle, blue |
Block, colourless |
Crystal size (mm3) |
0.48 × 0.06 × 0.06 |
0.40 × 0.36 × 0.22 |
T
max, Tmin |
0.746, 0.643 |
0.746, 0.651 |
Diffractometer |
Bruker-Nonius Kappa CCD |
Bruker-Nonius Kappa CCD |
Measured reflections |
5794 |
8562 |
Independent reflections |
3102 |
8562 |
F(000) |
1108 |
384 |
h
|
−24 to 23 |
−8 to 8 |
k
|
−6 to 6 |
−12 to 13 |
l
|
−37 to 37 |
0 to 19 |
Parameters refined |
|
|
R[F2 > 2σ(F2)] |
0.055 |
0.058 |
wR(F2) |
0.112 |
0.131 |
S
|
1.06 |
1.03 |
δρmax, δρmin (e Å−3) |
0.48, −0.43 |
0.50, −0.37 |
CCDC number |
2061208
|
2087028
|
2.3. Physical measurements
2.3.1. Infrared spectroscopy.
IR measurements were performed using a NICOLET IR 200 FT-IR spectrometer, with scans run over the range of 400–4000 cm−1.
2.3.2. UV solid state spectroscopy.
UV measurement was performed using a PerkinElmer Lambda spectrophotometer. Scans were run over the range of 200–800 cm−1.
2.3.3. The thermal analyses.
The thermal analysis spectra were obtained via simultaneous thermogravimetry-differential thermal analysis (TG-DTA) using the PYRIS 1 TGA instrument using 10.1 mg of [(C8H12N)2Co(SCN)4] and 9.80 mg of [(C8H12N)CNS], at a heating rate of 5 °C min−1 for the titled compound in the temperature range of 300–880 K, with the samples placed in aluminum pans under an inert atmosphere (nitrogen gas).
2.4. Computational details
The geometry optimization of the cobalt complex and thiocyanic salt was carried out using first-principles density functional theory (DFT) computations with the wB97XD functional and the def2SVP basis set. The choice of the ωB97XD/def2SVP method for the theoretical investigation of the electronic and structural studies of 1 and 2 is grounded on its proven accuracy in studying complexes analogous to 1 and 2 as shown by other independent studies.26–281 and 2 were optimized to the lowest minima to enhance the accuracy of the results obtained and avoid the impact of imaginary frequency.29 Structural optimization was carried out at the ωB97XD/def2SVP level of theory using Gaussian 1630 and visualized using GaussView 6.0.31 The visualization of the Frontier molecular orbital (FMO) was carried out using ChemCraft,32 while the visualization of the molecular electrostatic potential (MESP) was carried out using visual molecular dynamics (VMD) software.33 The analysis of the natural bond orbital (NBO) was carried out using the NBO 7.0 module34 embedded in Gaussian 16.
2.5. Molecular docking protocol
The molecular docking was carried out to investigate the anti-inflammatory potency of 1 and 2 using an in silico molecular docking approach with 1 and 2 as ligands using two anti-inflammatory proteins downloaded from the protein data bank (PDB) viahttps://www.rcsb.org with RCSB codes 1D8A35 and 2QIO,36 respectively. The pre-docking protein preparation of ID8A and 2QIO involved the deletion of water, defining the binding site, deletion of native ligands and extraneous polypeptide chains distal from the SBD sphere (ligand binding site), addition of polar hydrogen, and extraction of the attributes of the SBD sphere along the X, Y, and Z coordinates. All these were accomplished using Biovia Discovery Studio,37 while the in silico receptor–ligand docking was accomplished using AutoDock Vina,38 while the receptor–ligand interaction was visualized using PyMol39 for the 3D interaction and Biovia for the 2D interaction and the H-bond as well.
3. Results and discussion
3.1. Crystallographic study
Crystallographic data and structure refinements of [(C8H12N)2Co(SCN)4] and [(C8H12N)CNS] are listed in Table 1.
3.1.1. Crystal structure of [(C8H12N)2Co(SCN)4] (1).
The asymmetric unit of (1) comprises one tetra(isothiocyanate) cobalt [Co(NCS)4]2− anion and two 2-methylbenzylammonium cations (Fig. 1(a)). Selected bond distances and angles are given in Table 2.
|
| Fig. 1 ORTEP views of [(C8H12N)2Co(SCN)4] (a) and [(C8H12N)SCN] (b). | |
Table 2 Selected bond lengths and bond angles of the two cobalt complexes in the title compound
[(C8H12N)2Co(SCN)4] |
[(C8H12N)SCN] |
Co1–N2i |
1.956(3) |
N1–C1 |
1.491(6) |
Co1–N2 |
1.956(3) |
N1–H1C |
0.89(3) |
Co1–N3i |
1.959(3) |
N1–H1D |
0.90(3) |
Co1–N3 |
1.959(3) |
N1–H1E |
0.90(3) |
S1–C10 |
1.627(4) |
N2–H2A |
0.90(3) |
S2–C11 |
1.634(4) |
N2–H2B |
0.90(3) |
N2–C10 |
1.154(4) |
N2–H2C |
0.91(3) |
N3–C11 |
1.157(4) |
N4–C25 |
1.495(7) |
N1–C2 |
1.499(4) |
|
|
N2i–Co1–N2 |
115.87(18) |
C1–N1–H1C |
110(4) |
N2i–Co1–N3i |
106.41(12) |
C1–N1–H1D |
108(4) |
N2–Co1–N3i |
111.76(12) |
H1C–N1–H1D |
108(5) |
N2i–Co1–N3 |
111.76(12) |
C1–N1–H1E |
103(3) |
N2–Co1–N3 |
106.41(12) |
H1C–N1–H1E |
121(5) |
N3i–Co1–N3 |
104.05(18) |
H1D–N1–H1E |
106(5) |
C10–N2–Co1 |
171.0(3) |
N1–C1–C2 |
111.9(4) |
C11–N3–Co1 |
166.3(3) |
N1–C1–H1A |
109.2 |
N2–C10–S1 |
178.2(4) |
C2–C1–H1A |
109.2 |
N3–C11–S2 |
177.7(3) |
N1–C1–H1B |
109.2 |
|
|
C2–C1–H1B |
109.2 |
|
|
H1A–C1–H1B |
107.9 |
Symmetry code: (i) −x + 1, y, −z + 1/2. |
|
|
The coordination geometry of the central Co(II) ions can be described as a slightly distorted tetrahedron surrounded by four nitrogen atoms (Fig. S1(a), ESI†) and arranged along the b-axis direction. These anions lie to form anionic layers parallel to the (a,c) plane at y = ¼ and y = ¾. The main geometrical characteristics of the [Co(SCN)4]2− anion (Fig. S1(a), ESI†) are given in Table 2, where the average Co–N bond distance is 1.956 (3) A° and the N–Co–N bond angles vary in the range of 180.0–106°. These values are in agreement with those found in complexes containing the [M(NCS)n]m− anion such as [Ni(SCN)4]2−40 and [In(SCN)6]3−.41 The SCN− ligands connect to Co(II) atoms using the end-to-end bridging mode, producing a three-dimensional linear chain (3D) along the [100] direction via μ1,4-SCN–thiocyanate bridges (Fig. 2(a)). These cations are located at layers parallel to [001] edges (Fig S2(a), ESI†). The values of the bond distances C(2)–N(1): 1.499 (4) Å and C(2)–C(3): 1.515 (3) Å are in good agreement with those found in similar complexes [C6H5CH2P(Cl)(C6H5)3]2[Co(NCS)4],42 [2ClBzTPP]2[Co(NCS)4],43 (4NO2BzTPP)2[Zn(NCS)4]44 and [4NO2BzPy]2[Co(NCS)4].45
|
| Fig. 2 (a) Projection along the b-axis of the structure of [(C8H12N)2Co(SCN)4] (a) and [(C8H12N)SCN], (b) dotted lines indicate hydrogen bonds. | |
Each terminal nitrogen atom of the [C8H12N]+ group shares its protons in hydrogen bonding interactions of the type N–H⋯S(SCN), forming a three-dimensional network (Table 3). The N–H(C8H12N) ⋯S(SCN) distance values vary between 3.360(3) and 3.505(3) Å. The strength of the hydrogen bond can be interpreted according to the criteria concerning the distances N⋯S.46 These hydrogen bridging bonds (types: donors and acceptors) and the π–π and CH–π interactions ensure the cohesion and stability of the crystalline edifice (Fig. 3(a)). In addition, the stability of the compound is improved by the CH⋯π interactions between the CH frame and the aromatic rings with a distance of 3.458 Å to 3.993 Å.
Table 3 Hydrogen-bond geometry (Å, °)
D–H⋯A |
D–H |
H⋯A |
D⋯A |
D–H⋯A |
[(C8H12N)2Co(SCN)4] |
N1–H1A⋯S2ii |
0.90(2) |
2.59(3) |
3.360(3) |
144(3) |
N1–H1B⋯S1iii |
0.92(2) |
2.51(2) |
3.403(3) |
164(4) |
N1–H1C⋯S2 |
0.91(2) |
2.61(2) |
3.455(3) |
154(3) |
N1–H1C⋯S2iv |
0.91(2) |
2.90(3) |
3.505(3) |
125(3) |
|
[(C8H12N)SCN] |
N1–H1C⋯N5 |
0.89(3) |
2.02(3) |
2.886(7) |
163(5) |
N1–H1D⋯S1i |
0.90(3) |
2.52(3) |
3.409(5) |
170(5) |
N1–H1E⋯S2 |
0.90(3) |
2.53(3) |
3.367(5) |
155(5) |
N2–H2A⋯S2 |
0.90(3) |
2.68(4) |
3.409(5) |
139(5) |
N2–H2B⋯S3 |
0.90(3) |
2.38(3) |
3.280(6) |
173(5) |
N2–H2C⋯S4ii |
0.91(3) |
2.50(4) |
3.327(5) |
151(5) |
N3–H3A⋯N6 |
0.90(3) |
1.99(3) |
2.876(6) |
169(5) |
N3–H3B⋯S2iii |
0.92(3) |
2.51(3) |
3.388(5) |
159(5) |
N3–H3C⋯S1ii |
0.91(3) |
2.64(4) |
3.433(4) |
147(4) |
N4–H4A⋯S1 |
0.91(3) |
2.59(4) |
3.386(5) |
146(5) |
N4–H4B⋯N8 |
0.91(3) |
1.99(3) |
2.890(8) |
171(6) |
N4–H4C⋯N7 |
0.89(3) |
2.21(4) |
3.010(6) |
149(5) |
|
| Fig. 3 CH–π and π–π stacking in [(C8H12N)2Co(SCN)4] (a) and [(C8H12N)SCN] (b). | |
3.1.2. Crystal structure of [(C8H12N)SCN] (2).
The asymmetric unit of (2) (Fig. 1(b)) contains one 2-methylbenzyl ammonium cation (C1, C2, C3, C5, C6, C7, or C8/N1) and one thiocyanate anion (S1, C33, or N5). The 2-methylbenzyl ammonium cation (Fig. S2(b), ESI†) ring adopts a slightly distorted chair conformation, with puckering parameters: θ = 120.8 (5)°, while for an ideal chair configuration, θ has a value of 109 to 180°. The bond lengths and bond angles (Table 2) are in normal ranges and are comparable with those reported earlier for similar compounds.47,48 The different types of H-bonds for [(C8H12N)CNS] are shown in Fig. 2(b) as dashed light red lines and their relative geometrical parameters are listed in Table 3.
The compound exhibits N–H⋯N and N–H⋯S H-bonds which give rise to 3D chains. Regarding the donor–acceptor bond lengths, all the hydrogen bonds in the studied system are found to be weak (D–A > 3 Å) and primarily involve electrostatic interactions.49 The intermolecular hydrogen bonding interactions link neighboring thiocyanate anions through N–H⋯N hydrogen bonds with lengths from 1.99(3) Å to 2.21(4) Å and through N–H⋯S hydrogen bonds with lengths from 2.38(3) Å to 2.52(3) Å (Table 3), contributing to the R26(16) rings as shown in Fig. S3 (ESI†), in addition to the CH–π stacking interactions (Fig. 3(b)) between aromatic rings which provides more stability to the three-dimensional framework with a stacking distance of 3.872 Å.
[(C8H12N)2Co(SCN)4]: symmetry codes: (ii) x, y + 1, z; (iii) x, y − 1, z; (iv) −x + 1/2, y + 1/2, −z + 1/2.
[(C8H12N)SCN]: symmetry code: (i) x − 1, y, z; (ii) x, y + 1, z; (iii) x + 1, y, z.
3.2. Structural analysis
The structural properties (bond length and bind angle) of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] were investigated theoretically using GaussView and compared with the experimental data to compare the degree of consonance or dissonance complemented with the calculation of the root mean square deviation (RMSD). Table 4 shows the structural properties of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] with special emphasis on the experimental and theoretical bond length and bond angle of the selected bond labels, while Fig. 4 shows the optimized structures of [(C8H12N)2Co(SCN)4] (Fig. 4(a)) and [(C8H12N)SCN] (Fig. 4(b)), respectively.
Table 4 The bond length and bond angles of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] investigated at the ωB97XD/def2SVP level of theory
Bond length (Å) |
[(C8H12N)2Co(SCN)4] |
[(C8H12N)SCN] |
Atom label |
Experimental |
Theoretical |
Atom label |
Experimental |
Theoretical |
Co1–N2 |
1.956 |
1.874 |
S2–C34 |
1.656 |
1.640 |
Co1–N3 |
1.959 |
1.820 |
N1–C1 |
1.491 |
1.176 |
S1–C10 |
1.627 |
1.611 |
C3–C4 |
1.396 |
1.393 |
N2–C10 |
1.154 |
1.182 |
C6–C7 |
1.386 |
1.340 |
N3–C11 |
1.157 |
1.180 |
N2–C9 |
1.467 |
1.490 |
RMSD |
|
0.068 |
|
|
0.130 |
Bond angle (°) |
[(C8H12N)2Co(SCN)4] |
[(C8H12N)SCN] |
Atom label |
Experimental |
Theoretical |
Atom label |
Experimental |
Theoretical |
C10–N2–Co1 |
171.0 |
159.021 |
C1–N1–H1 |
110.000 |
103.511 |
C11–N3–Co1 |
166.3 |
153.064 |
C17–N3–H3 |
108.000 |
111.375 |
N2–C10–S1 |
178.2 |
178.111 |
C7–C6–C5 |
118.800 |
119.992 |
N3–C11–S2 |
177.7 |
177.261 |
C4–C5–C6 |
120.300 |
119.207 |
N1–C2–C3 |
110.4 |
111.384 |
C6–C7–H7 |
119.500 |
119.396 |
RMSD |
|
7.301 |
|
|
3.058 |
|
| Fig. 4 The optimized structures of [(C8H12N)2Co(SCN)4] (a) and [(C8H12N)SCN] (b) optimized at the ωB97XD/def2SVP level and visualized using ChemCraft. | |
As presented in Table 4, the five intramolecular bonds considered in 1 were Co1–N2, Co1–N3, S1–C10, N2–C10, and N3–C11. The first three were longer in the experimental data, while in the theoretical data, they were slightly shorter with differences in magnitude of 0.082 Å, 0.139 Å, and 0.016 Å, respectively. This implies that the theoretical values were 4.2%, 7.1%, and 1.0% different from the corresponding experimental values. The last two bonds were slightly longer theoretically than experimentally with differences in magnitude being 0.028 Å and 0.023 Å, respectively, which further implies that the experimental values were 2.4% and 1.9% lower than the theoretical magnitudes, respectively. Conversely, the first four of the five bond labels considered in [(C8H12N)SCN] were higher experimentally than theoretically. Experimentally, the magnitudes of the S2–C34, N1–C1, C3–C4, and C6–C7 bond labels were 1.565 Å, 1.491 Å, 1.396 Å and 1.386 Å, respectively, while the theoretical values of 1.640 Å, 1.176 Å, 1.393 Å and 1.340 Å were 0.016 Å, 0,315 Å, 0.003 Å and 0.046 Å, respectively, which correspond to the magnitude differences of 1.0%, 21. 1%, 0.2%, and 3.3%, respectively. Furthermore, the N2–C9 bond had a theoretical value of 1.490 Å and a slightly lower experimental magnitude of 1.467 Å. Overall, the RMSD of [(C8H12N)2Co(SCN)4] was found to be lower with a value of 0.068 compared to that of 2 with a higher value of 0.130.
The bond angles of the selected intramolecular bond labels were also studied at the same level of theory and a great degree of harmony between the experimental and theoretical values was observed as shown by the not-so-high RMSD of 7.301 and 3.058 for [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN], respectively. In [(C8H12N)2Co(SCN)4], the C10–N2–Co1, C11–N3–Co1, N2–C10–S1, and N3–C11–S2 bond angles were experimentally higher than the theoretical values with the differences in the values being 11.979, 13.236, 0.089 and 0.439, respectively, thereby indicating that the last two showed a higher degree of consonance compared to the first two. On the other hand, the N1–C2–C3 bond angle was theoretically higher with a theoretical value of 11.384, while the experimental value was 0.88343 degrees lower with a magnitude of 110.4. Conversely, the [(C8H12N)SCN] bond angles showed a higher degree of closeness between the experimental and theoretical values. The C1–N1–H1, C4–C5–C6, and C6–C7–H7 bond angles were higher experimentally than theoretically with the differences of 6.5, 1.1, and 0.1 degrees, respectively, while the C17–N3–H3 and C7–C6–C5 bond angles were higher theoretically than experimentally with differences in value being 3.4 and 1.2 degrees, respectively.
3.3. Infrared spectroscopy investigation
Fig. 5 displays the IR spectrum of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN]. The characteristic vibrations can prove the presence of the thiocyanate ligand and its binding mode to the Co(II) ion center for the formation of the anionic complex [Co (SCN)4]. The strong band at 2079 cm−1 can be assigned to the stretching vibration of the CN bond of the thiocyanic ligand. The band at 840 cm−1 is attributed to the C–S bond stretching vibration. The band at 490 cm−1 is ascribed to the bending vibration of N–C–S. This vibrational assignment of the thiocyanate indicates the binding of the thiocyanate ligand to the metal (II) center via its N-terminal atom. The assignment of these bands to thiocyanate vibrations and the determination of its coordination mode are based on previously reported results such as those for (C2N6H12) [Co (NCS)4]. H2O50 and [2NAPMeTPP]2[Co(NCS)4].51
|
| Fig. 5 Infrared absorption spectra of [(C8H12N)2Co(SCN)4] (a) and [(C8H12N)SCN] (b). | |
The spectrum also shows characteristic vibrations for 2-methylbenzylamine. The broad bands in the range of 3600–2300 cm−1 correspond to the stretching vibrations of the organic and hydroxyl groups ν(N–H) and ν(C–H). The band at 1504 cm−1 corresponds to ν(CC) stretching vibrations. The band at 1450 cm−1 can be assigned to the CH2 deformation.52
The bands at 1244 and 1180 cm−1 can be attributed to the ring deformation. The weak bands at 1166 and 1021 cm−1 can be assigned to the CH2 twisting. The weak band at 870 cm−1 can be attributed to the ring deformation.53
3.4. Solid-state spectroscopy
The luminescence properties have been tested for the solid states of [(C8H12N)2Co(SCN)4] and [(C8H12N)CNS] at room temperature, in the region [200–800 nm]. As depicted in Fig. 6, the compounds show different luminescence behaviors; the two characteristic bonds for [(C8H12N)2Co(SCN)4] at 598, 456, and 402 nm are assigned to d → d*, charge transfer and n → π* transitions, respectively, and for [(C8H12N)SCN], one characteristic bond at 290 nm is assigned to charge transfer transition.
|
| Fig. 6 Solid state UV-vis spectra of [(C8H12N)2Co(SCN)4] (a) and [(C8H12N)SCN] (b). | |
The calculation of Eg revealed that [(C8H12N)2Co(SCN)4] exhibits semiconductor behavior with a gap energy value of 2.09 eV and [(C8H12N)SCN] exhibits isolant behavior with a gap energy of 3.8 eV (Fig. S4, ESI†). These behaviors are probably due to the interactions in the molecular solid, the charge transfer between the central metal and its coordinated ligands, and especially due to the presence of the thiocyanate anions that may affect the emission.54
3.5. Electronic properties
3.5.1. Quantum reactivity descriptors.
Following the advances in the modern-day understanding of the Frontier molecular orbital (FMO) theory55 of chemical reactivity, there is a global consensus on the indispensable pertinence of the two major frontier molecular orbitals, namely, the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) to the reactivity and invariably the stability of a compound.56,57Table 5 presents the electronic properties of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] as well as the magnitudes of their global quantum reactivity descriptors, while Fig. 7 is a pictorial illustration of the HOMOs and LUMOs of [(C8H12N)2Co(SCN)4] (Fig. 7(a)) and [(C8H12N)SCN] (Fig. 7(b)), respectively.
Table 5 The electronic properties and the quantum descriptors of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] investigated at the ωB97XD/def2SVP level of theory
Complex |
HOMO (eV) |
LUMO (eV) |
E
gp (eV) |
IP (eV) |
EA (eV) |
S (eV) |
η (eV) |
μ (eV) |
χ (eV) |
ω (eV) |
A1 |
−7.645 |
−4.640 |
3.005 |
7.645 |
4.640 |
0.333 |
1.502 |
6.142 |
−6.142 |
12.557 |
A2 |
−5.665 |
−5.319 |
0.346 |
5.665 |
5.319 |
2.894 |
0.173 |
5.492 |
−5.492 |
87.281 |
|
| Fig. 7 the HOMO–LUMO graphics of [(C8H12N)2Co(SCN)4] (a) and [(C8H12N)SCN] (b) visualized using ChemCraft. | |
Fundamentally, the difference between the HOMO and the LUMO is generally known as the energy gap, Egp, and can be used to theoretically compare the chemical reactivity of compounds because the higher the energy gap the less reactive the compound and the greater the stability, while a lower energy gap indicates higher reactivity and lower stability. [(C8H12N)2Co(SCN)4] has an Egp of 3.005 eV and that of [(C8H12N)SCN] has a calculated value of 0.346 eV. This deductively implies that [(C8H12N)2Co(SCN)4] is predicted to have a lower reactivity compared to [(C8H12N)SCN] and may be accounted for by the presence of Co as a complexation metal in [(C8H12N)2Co(SCN)4], while in [(C8H12N)SCN] without a complexing metal, the reactivity is higher. This also implies that [(C8H12N)2Co(SCN)4] is more stable than [(C8H12N)SCN]. This variation in the comparative reactivity and stability strength of the Co complex and [(C8H12N)SCN] is further elucidated by the chemical softness and chemical hardness which give more information about the reactivity and stability of a compound, respectively. This implies that the higher the chemical softness, the higher the reactivity and vice versa, while the higher chemical hardness corresponds to the greater stability for both complexes; [(C8H12N)2Co(SCN)4] with a chemical softness of 0.333 eV is equally predicted to be less reactive than [(C8H12N)SCN] with a chemical softness of 2.894 eV, while the lower chemical hardness of [(C8H12N)SCN] corroborates this. Furthermore, application of Koopman's approximations can be used to obtain certain other quantum descriptors namely: chemical potential (μ), chemical hardness (η), chemical softness (S), electrophilicity index (ω), and electronegativity (χ) using eqn (1)–(7).58
| | (3) |
| | (4) |
| | (5) |
| | (6) |
3.5.2. Natural bond orbitals (NBOs).
The study of natural bond orbitals is aimed at investigating the inter-molecular charge transfer from donor orbitals to acceptor orbitals and in this specific case, this is done using the Second Order Perturbation Theory of the Fock Matrix in the NBO basis.34Table 6 presents the results of the NBO analysis of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN], including the transitions, the donor orbitals, acceptor orbitals, occupancies as well as stabilization energies which were computed based on eqn (8). | | (8) |
where fij is the off-diagonal element on the Fock matrix, E(j) − E(i) are the diagonal elements and E(2) represent the second order stabilization energy of the system.59 From Table 6, the transitions with the greatest contribution to the stability of 1 are LP(3) → LP(1)*, σ → LP(1)*, σ → LP(4)* and σ* → σ*. The LP(3) → LP(1)* transition from S13 to C6 had the highest stabilization energy of 177.17 kcal mol−1, thereby implying that this transition contributed the most to the orbital stabilization of [(C8H12N)2Co(SCN)4]. Furthermore, the σ → LP(1)* transition with an E(2) of 138.90 kcal mol−1 was due to the electron transfer from the complexing metal, Co1 to C6. The σ → LP(4)* transition occurred twice with notable stabilization energies of 135.80 and 135.53 kcal mol−1, respectively, while the σ* → σ* transition from Co1–N3 to Co1–N5 had an E(2) of 113.42 kcal mol−1. Conversely, in [(C8H12N)SCN], the E(2) values were comparatively lower than those of [(C8H12N)2Co(SCN)4] which equally corroborates the FMO observation that this complex is predicted to be comparatively more stable but less reactive than [(C8H12N)SCN]. The values of the stabilization energies of [(C8H12N)SCN] ranged from 10.45 kcal mol−1 to 69.59 kcal mol−1 and were due to five different transitions namely: LP(3) → π*, π* → σ*, LP(2) → π*, LP(1) → σ* and π* → σ*. The orbital transition with the highest contribution to the stability of [(C8H12N)SCN] is the LP(3) → π* with an E(2) of 69.69 kcal mol−1, while the π* → σ* transition had the least contribution with an E(2) of 10.45 kcal mol−1.
Table 6 The natural bond orbital (NBO) of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] investigated at the ωB97XD/def2SVP level of theory
Transition |
Donor |
Occupancy |
Acceptor |
Occupancy |
E
(2) (kcal mol−1) |
E
(j)−(i) (a.u.) |
F(i,j) (a.u.) |
[(C8H12N)2Co(SCN)4] |
LP(3) → LP(1)* |
S13 |
1.63739 |
C6 |
0.88420 |
177.17 |
0.15 |
0.167 |
σ → LP(1)* |
Co1 |
1.61138 |
C6 |
0.88420 |
138.90 |
0.28 |
0.201 |
σ → LP(4)* |
Co1–N5 |
0.21653 |
Co1 |
0.39697 |
135.80 |
0.11 |
0.196 |
σ → LP(4)* |
Co1–N3 |
0.30559 |
Co1 |
0.39697 |
135.53 |
0.12 |
0.195 |
σ* → σ* |
Co1–N3 |
0.30559 |
Co1–N5 |
0.21653 |
113.42 |
0.01 |
0.069 |
[(C8H12N)SCN] |
LP(3) → π* |
S3 |
1.76244 |
N1–C2 |
0.22839 |
69.59 |
0.41 |
0.150 |
π* → σ* |
N1–C2 |
0.18127 |
N1–C3 |
0.06469 |
37.65 |
0.46 |
0.335 |
LP(2) → π* |
S3 |
1.76463 |
N1–C2 |
0.18127 |
37.08 |
0.58 |
0.133 |
LP(1) → σ* |
S3 |
1.98291 |
N1–C2 |
0.06469 |
15.56 |
1.54 |
0.140 |
π* → σ* |
N1–C2 |
0.18127 |
C2–S3 |
0.01068 |
10.45 |
0.10 |
0.091 |
3.5.3. Molecular electrostatic potential (MESP).
The intra-molecular differential charge distribution is a phenomenon that can be accounted for by the fact that the nucleophilicity or electrophilicity of the atoms that constitute a molecule cannot be the same, hence in a molecule, the regions with a greater electron-pulling tendency are usually characterized by their possession of lone pairs, high electronegativity as well as high electron affinity.60,61 Structurally, [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] have some structural resemblance, but due to the inherent structural differences, they have different positions of the nucleophilic and electrophilic regions as illustrated by the red and blue regions, respectively, in Fig. 8.
|
| Fig. 8 The graphical illustration of the molecular electrostatic potentials of [(C8H12N)2Co(SCN)4] (a) and [(C8H12N)SCN] (b) visualized using VMD. | |
3.6. Molecular docking analysis
The bioactivity of a compound depends on its structural complementarity with the target receptor62–64 and can be investigated using the in silico molecular docking simulation as this approach can provide useful insights into the binding strength and affinity of bonding between a receptor and a ligand.65 The results as provided in Table 7 summarize the binding affinity (kcal mol−1) and amino acid residues for the receptor–ligand interaction between [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] with 1D8A and 2QIO, respectively, while Fig. 9 shows the 2D, 3D and the H-bond of the molecular docking as carried out in this study. As presented in Table 7, the anti-inflammatory potentials of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] were specific for 1D8A and 2QIO, respectively, as the docking scores of 1D8A-(a) for[(C8H12N)2Co(SCN)4] and the 2QIO-(b) for [(C8H12N)SCN] interactions had 0.0 binding affinity in all the 9 poses. Table 7 shows that in the 1D8A-(a) interaction, the mean binding affinity was −5.0 kcal mol−1, while the first and ninth pose binding affinities were −5.4 and −4.6 kcal mol−1 with the amino acid residues arising from the interaction of 1 with the 1D8A receptor being a two-fold interaction of 1 with the alanine 189 (ALA:189) of the polypeptide chain B of 1D8A at distances of 2.47 Å and 2.91 Å, respectively. On the other hand, the 2QIO-(b) interaction had a slightly greater mean binding affinity of −5.1 kcal mol−1 with the first and ninth poses having binding affinities of −5.5 kcal mol−1 and −5.3 kcal mol−1, while the amino acid residues were three and were at the alanine 190 (ALA:190) of the polypeptide chain C with interaction bond lengths of 2.58 Å, 2.63 Å and 2.97 Å, respectively. Based on the foregoing, the anti-inflammatory potential of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] has been established using in silico bioactivity prediction.
Table 7 The binding affinity and amino acid residues of the receptor–ligand interactions of [(C8H12N)2Co(SCN)4] and [(C8H12N)SCN] with 1D8A and 2QIO
MODE |
Anti-inflammatory potential |
Binding affinity (kcal mol−1) |
Amino acid residue |
Binding affinity (kcal mol−1) |
Binding affinity (kcal mol−1) |
Binding affinity (kcal mol−1) |
Amino acid residue |
1D8A-A1 |
1D8A-A1 |
2QIO-A1 |
1D8A-A2 |
2QIO-A2 |
2QIO-A2 |
1 |
−5.5 |
B:ALA:189:O (2.91 Å and 2.47 Å) |
0.0 |
0.0 |
−5.5 |
C:ALA:190:O (2.97 Å, 2.58 Å and 2.63 Å) |
2 |
−5.4 |
|
0.0 |
0.0 |
−5.5 |
|
3 |
−5.1 |
|
0.0 |
0.0 |
−5.3 |
|
4 |
−4.9 |
|
0.0 |
0.0 |
−5.2 |
|
5 |
−4.9 |
|
0.0 |
0.0 |
−5.0 |
|
6 |
−4.8 |
|
0.0 |
0.0 |
−4.9 |
|
7 |
−4.7 |
|
0.0 |
0.0 |
−4.9 |
|
8 |
−4.7 |
|
0.0 |
0.0 |
−4.9 |
|
9 |
−4.6 |
|
0.0 |
0.0 |
−4.8 |
|
Mean |
−5.0 |
|
0.0 |
0.0 |
−5.1 |
|
|
| Fig. 9 The graphical illustration of the molecular docking interaction between 1D8A and 2QIO with [(C8H12N)2Co(SCN)4] (a) and [(C8H12N)SCN] (b); molecular docking analysis was performed using Biovia Discovery Studio, while the in silico receptor–ligand docking was accomplished using AutoDock Vina, while the receptor–ligand interaction was visualized using PyMol for the 3D interaction and Biovia for the 2D interaction and the H-bond as well. | |
3.7. Thermal properties
DTA and TGA were carried out using the PYRIS 1 TGA instrument. The thermal curves of the two coordination compounds are given in Fig. 10(a) and (b).
|
| Fig. 10 DTA/TG curves of [(C8H12N)2Co(SCN)4] (a) and [(C8H12N)SCN] (b). | |
[(C8H12N)2Co(SCN)4] shows two endothermic peaks at 390 K and 470 K. The first one corresponds to the phase transition without the loss of weight detected by differential thermogravimetry and the second one corresponds to the degradation of the organic part and some of the thiocyanate ligands from the metal compound, as confirmed by TGA analysis. The decomposition in the [420–480 K] range has the same variation for the compound (1), and the decomposition of the resulting Co(NCS)4 is carried out at a higher temperature.66 During the process, the SO group, free from hydrogen bonding, is oxidized to SO2 at [520–560 K].66–68
For [(C8H12N)SCN], the TGA studies shows a phase transition at 323 K without any weight loss. The decomposition process begins at 453 K for [(C8H12N)SCN], during this endothermic process, the mass loss in this step can be attributed to the successive decomposition of the titled complex. This endothermic step takes place when the intermolecular N–H⋯S hydrogen bonding breaks apart during the deamination process and one of the three N–H groups per complex molecule is no longer engaged in hydrogen bonding. During the latest process, the SO group is oxidized to SO2 at 520 K.
4. Conclusion
Novel crystalline compounds [(C8H12N)2Co(SCN)4] (1) and [(C8H12N)SCN] (2) were synthesized and analyzed using XRD single crystallography, FTIR spectroscopy, differential thermal gravimetric analysis, complex impedance measurements, and magnetic studies. The crystal structure (1) shows that Co(II) atoms are bridged by four SCN− anions generating 3-D chains assured by hydrogen bonds, van der Waals, and CH–π interactions. Crystal (2) shows that intermolecular cohesion is ensured by N–H⋯S, N–H⋯N hydrogen bonds and CH–π stacking interactions. Furthermore, frontier molecular orbital analysis and NBO analyses have been calculated using the DFT method.
TGA/DTA coupled analysis shows that these compounds undergo a phase transition at 390 K for (1) and at about 323 K for (2). Optical properties were investigated through absorption measurements. It is found that the gap energy value of the formed compounds is 2.09 eV for (1) and 3.80 eV for (2); according to these results, we can predict that our finding is a semiconductor. The molecular docking occurred for both complexes, suggesting that the complexes have an anti-inflammatory potential.
Data availability
CCDC 2061208 and 2087028 contain the supplementary crystallographic data for the obtained compounds. The DSC diagram and B3LYP-converged geometry, and the antibacterial activities of the title complex are also provided in the ESI.†
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
The authors extend their appreciation to the Researchers Supporting Project, King Saud University, Riyadh, Saudi Arabia, for funding this work through grant No. RSP2024R353.
References
- A. P. Singh, A. Ali and R. Gupta, Dalton Trans., 2010, 39, 8135–8138 RSC.
- E. V. Andrusenko, A. S. Novikov, G. L. Starova and N. A. Bokach, Inorg. Chim. Acta, 2016, 447, 142–149 CrossRef CAS.
- S. G. Telfer and J. D. Wuest, Cryst. Growth Des., 2009, 9, 1923–1931 CrossRef CAS.
- J. Seo, S. H. Vegi, C. K. Ranaweera, N. K. Baradanahalli, J.-H. Han, D. Koli and S. V. Babu, ECS J. Solid State Sci. Technol., 2018, 8, P3009 CrossRef.
- L. Li, S. Chen, R.-M. Zhou, Y. Bai and D.-B. Dang, Spectrochim. Acta, Part A, 2014, 120, 401–404 CrossRef CAS PubMed.
- A. Z. Spentzos, S. A. Ford, B. Muldowney, T. Liu, P. Cui, M. R. Gau, P. J. Carroll and N. C. Tomson, Organometallics, 2024, 43(13), 1425–1437 CrossRef CAS.
- X. Feng, Y. Song, J. S. Chen, Z. Li, E. Y. Chen, M. Kaufmann, C. Wang and W. Lin, Chem. Sci., 2019, 10, 2193–2198 RSC.
- V. Bernales, A. B. League, Z. Li, N. M. Schweitzer, A. W. Peters, R. K. Carlson, J. T. Hupp, C. J. Cramer, O. K. Farha and L. Gagliardi, J. Phys. Chem. C, 2016, 120, 23576–23583 CrossRef CAS.
- R. S. Joseyphus, E. Viswanathan, C. J. Dhanaraj and J. Joseph, J. King Saud Univ., 2012, 24, 233–236 CrossRef.
- A. H. Ahmed and M. G. Moustafa, J. Saudi Chem. Soc., 2020, 24, 381–392 CrossRef CAS.
- J. Makhlouf, H. Louis, I. Benjamin, B. B. Isang, C. F. Fidelis, A. Valkonen and W. Smirani, Polyhedron, 2023, 235, 116369 CrossRef CAS.
- J. Makhlouf, A. Ahsin, Y. El Bakri, A. Valkonen, R. Al-Salahi and W. Smirani, Inorg. Chem. Commun., 2023, 158, 111537 CrossRef CAS.
- S. Liu, Y.-F. Deng, Z.-Y. Chen, L. Meng, X. Chang, Z. Zheng and Y.-Z. Zhang, CCS Chem., 2021, 3, 2530–2538 CrossRef CAS.
- J. Makhlouf, Y. El Bakri, A. Valkonen, K. Saravanan, S. Ahmad, R. Al-Salahi and W. Smirani, Inorg. Chem. Commun., 2024, 159, 111723 CrossRef CAS.
- A. Hannachi, A. Valkonen, C. J. G. García, M. Rzaigui and W. Smirani, Polyhedron, 2019, 173, 114122 CrossRef CAS.
- O. A. Abbas, O. M. El-Roudi and S. A. Abdel-Latif, Appl. Organomet. Chem., 2023, 37, e7236 CrossRef CAS.
- N. S. Abdel-Kader, S. A. Abdel-Latif, A. L. El-Ansary and M. A. Hemeda, Polycyclic Aromat. Compd., 2023, 1–32 Search PubMed.
- W. Omara, S. A. Abdel-Latif, H. M. E. Azzazy and N. S. Abdel-Kader, Appl. Organomet. Chem., 2024, 38, e7439 CrossRef CAS.
- J. Makhlouf, Y. El Bakri, K. Saravanan, A. Valkonen, C. J. Gomez Garcia and W. Smirani Sta, Inorg. Chem. Commun., 2023, 154, 110914 CrossRef CAS.
- A. M. Abu-Dief, R. M. El-Khatib, T. El-Dabea, S. A. Abdel-Latif, I. O. Barnawi, F. S. Aljohani, K. Al-Ghamdi and M. A. A. A. El-Remaily, Appl. Organomet. Chem., 2024, e7463 CrossRef CAS.
- J. Makhlouf, Y. El Bakri, K. Saravanan, A. Valkonen and W. Smirani, J. Mol. Struct., 2023, 1281, 135049 CrossRef CAS.
- A.-Q. S. Soliman, S. A. Abdel-Latif, S. Abdel-Khalik, S. M. Abbas and O. M. Ahmed, Design, synthesis, structural characterization, molecular docking, antibacterial, anticancer activities, and density functional theory calculations of novel MnII, CoII, NiII, and CuII complexes based on pyrazolone-sulfadiazine azo-dye ligand, J. Mol. Struct., 2024, 1318, 139402 CrossRef CAS.
- G. M. Sheldrick, Acta Crystallogr., Sect. C: Struct, Chem., 2015, 71, 3–8 Search PubMed.
- L. J. Farrugia, J. Appl. Crystallogr., 2012, 45, 849–854 CrossRef CAS.
- J. H. Kaneko, T. Tanaka, T. Imai, Y. Tanimura, M. Katagiri, T. Nishitani, H. Takeuchi, T. Sawamura and T. Iida, Nucl. Instrum. Methods Phys. Res., Sect. A, 2003, 505, 187–190 CrossRef CAS.
- N. Cabrera, J. R. Mora, E. Márquez, V. Flores-Morales, L. Calle and E. Cortés, SAR QSAR Environ. Res., 2021, 32, 29–50 CrossRef CAS PubMed.
- O. Noureddine, N. Issaoui, S. Gatfaoui, O. Al-Dossary and H. Marouani, J. King Saud Univ., 2021, 33, 101283 CrossRef.
- M. I. Sorour, A. H. Marcus and S. Matsika, Molecules, 2022, 27, 4062 CrossRef CAS PubMed.
- S. Spicher and S. Grimme, J. Chem. Theory Comput., 2021, 17, 1701–1714 CrossRef CAS PubMed.
-
M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson and H. Nakatsuji, GaussView 5.0., Wallingford, E.U.A., 2016.
-
R. Dennington, T. A. Keith and J. M. Millam, GaussView, Version 5.0.8., Semichem Inc., Shawnee Mission, KS, USA., 2009.
- S. V. Zaitseva, S. A. Zdanovich and O. I. Koifmana, Macroheterocycles, 2012, 5(1), 81–86 CrossRef.
- W. Humphrey, A. Dalke and K. Schulten, J. Mol. Graphics, 1996, 14, 33–38 CrossRef CAS PubMed.
- E. D. Glendening, C. R. Landis and F. Weinhold, J. Comput. Chem., 2013, 34, 1429–1437 CrossRef CAS PubMed.
- C. W. Levy, A. Roujeinikova, S. Sedelnikova, P. J. Baker, A. R. Stuitje, A. R. Slabas, D. W. Rice and J. B. Rafferty, Nature, 1999, 398, 383–384 CrossRef CAS PubMed.
- S. K. Tipparaju, D. C. Mulhearn, G. M. Klein, Y. Chen, S. Tapadar, M. H. Bishop, S. Yang, J. Chen, M. Ghassemi and B. D. Santarsiero, ChemMedChem, 2008, 3, 1250–1268 CrossRef CAS PubMed.
-
D. S. Biovia and B. D. S. Biovia, San Diego, CA, USA, 2019. Discovery Studio Visualizer, 2019, vol. 12.
- O. Trott and A. J. Olson, J. Comput. Chem., 2010, 31, 455–461 CrossRef CAS PubMed.
- W. L. DeLano, Pymol: An open-source molecular graphics tool, CCP4 Newsl. Protein Crystallogr., 2002, 40(1), 82–92 Search PubMed.
- A. Ferchichi, J. Makhlouf, Y. El Bakri, K. Saravanan, A. Valkonen, H. E. Hashem, S. Ahmad and W. Smirani, Sci. Rep., 2022, 12, 1–18 CrossRef PubMed.
- F. W. B. Einstein, M. M. Gilbert, D. G. Tuck and P. L. Vogel, Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem., 1976, 32, 2234–2235 CrossRef.
- W.-Q. Chen, L.-J. Su, X.-Q. Cai, J.-J. Yang, Y.-L. Qian, X.-P. Liu, L.-M. Yang, J.-R. Zhou and C.-L. Ni, Synth. React. Inorg., Met.-Org., Nano-Met. Chem., 2014, 44, 980–985 CrossRef CAS.
- H.-Q. Ye, L.-J. Su, X.-X. Chen, X. Liao, Q.-T. Liu, X.-Y. Wu, J.-R. Zhou, L.-M. Yang and C.-L. Ni, Synth. Met., 2015, 199, 232–240 CrossRef CAS.
- H.-Q. Ye, J.-L. Xie, J.-Y. Yu, Q.-T. Liu, S.-L. Dai, W.-Q. Huang, J.-R. Zhou, L.-M. Yang and C.-L. Ni, Synth. Met., 2014, 197, 99–104 CrossRef CAS.
- H.-Q. Ye, Y.-Y. Li, R.-K. Huang, X.-P. Liu, W.-Q. Chen, J.-R. Zhou, L.-M. Yang and C.-L. Ni, J. Struct. Chem., 2014, 55, 691–696 CrossRef CAS.
- I. M. Walker and P. J. McCarthy, Inorg. Chem., 1984, 23, 1842–1845 CrossRef CAS.
- A. Shimada, Yos Okaya and M. Nakaura, Acta Crystallogr., 1955, 8, 819–822 CrossRef CAS.
- D. N. Smith and M. E. Devey, Genome, 1994, 37, 977–983 CrossRef CAS PubMed.
- T. Steiner, Angew. Chem., Int. Ed., 2002, 41, 48–76 CrossRef CAS.
- E. M. Rakhman’ko, Y. V. Matveichuk and V. V. Yasinetskii, J. Anal. Chem., 2015, 70, 178–185 CrossRef.
- H.-T. Cai, Q.-T. Liu, H.-Q. Ye, L.-J. Su, X.-X. Zheng, J.-N. Li, S.-H. Ou, J.-R. Zhou, L.-M. Yang and C.-L. Ni, Spectrochim. Acta, Part A, 2015, 142, 239–245 CrossRef CAS PubMed.
- G. Ionita, G. Marinescu, C. Ilie, D. F. Anghel, D. K. Smith and V. Chechik, Langmuir, 2013, 29, 9173–9178 CrossRef CAS PubMed.
- S. J. Osborne, S. Wellens, C. Ward, S. Felton, R. M. Bowman, K. Binnemans, M. Swadźba-Kwaśny, H. Q. N. Gunaratne and P. Nockemann, Dalton Trans., 2015, 44, 11286–11289 RSC.
- C. Näther, I. Jess and S. Mangelsen, Z. Anorg. Allg. Chem., 2023, 649, e202300120 CrossRef.
- J. I. Seeman, Chem. Rec., 2022, 22, e202100302 CrossRef CAS PubMed.
- A. I. Ikeuba, A. U. Agobi, L. Hitler, B. J. Omang, F. C. Asogwa, I. Benjamin and M. C. Udoinyang,
et al.
, Chem. Afr., 2023, 6(2), 983–997 CrossRef CAS.
- P. Nkoe, A.-L. E. Manicum, H. Louis, F. P. Malan, W. J. Nzondomyo, K. Chukwuemeka, S. A. Sithole, A. Imojara, C. M. Chima and E. C. Agwamba, J. Mol. Liq., 2023, 370, 121045 CrossRef CAS.
- A. D. Udoikono, H. Louis, E. A. Eno, E. C. Agwamba, T. O. Unimuke, A. T. Igbalagh, H. O. Edet, J. O. Odey and A. S. Adeyinka, J. Photochem. Photobiol., 2022, 10, 100116 CrossRef CAS.
- F. Weinhold, C. R. Landis and E. D. Glendening, Int. Rev. Phys. Chem., 2016, 35(3), 399–440 Search PubMed.
- A. Varadwaj, H. M. Marques and P. R. Varadwaj, Molecules, 2019, 24, 379 Search PubMed.
-
Murray, J. S. and Sen, K., ed., Molecular electrostatic potentials: concepts and applications, 1996 Search PubMed.
- I. G. Osigbemhe, H. Louis, E. M. Khan, E. E. Etim, E. E. Oyo-Ita, A. P. Oviawe, H. O. Edet and F. Obuye, Appl. Biochem. Biotechnol., 2022, 194, 5680–5701 CrossRef CAS PubMed.
- L. Pinzi and G. Rastelli, Int. J. Mol. Sci., 2019, 20, 4331 CrossRef CAS PubMed.
- J. Fan, A. Fu and L. Zhang, Quant. Biol., 2019, 7, 83–89 CrossRef CAS.
- P. S. Idante, G. C. Apebende, H. Louis, I. Benjamin, U. J. Undiandeye and I. J. Ikot, J. Indian Chem. Soc., 2023, 100, 100806 CrossRef CAS.
- J. Werner, M. Rams, Z. Tomkowicz, T. Runcevski, R. E. Dinnebier, S. Suckert and C. Näther, Inorg. Chem., 2015, 54, 2893–2901 CrossRef CAS PubMed.
- J. Werner, T. Runčevski, R. Dinnebier, S. G. Ebbinghaus, S. Suckert and C. Näther, Eur. J. Inorg. Chem., 2015, 3236–3245 CrossRef CAS.
- S. Wöhlert, T. Runčevski, R. E. Dinnebier, S. G. Ebbinghaus and C. Näther, Cryst. Growth Des., 2014, 14, 1902–1913 CrossRef.
Footnote |
† Electronic supplementary information (ESI) available: CCDC 2061208 and 2087028 contain the supplementary crystallographic data for the obtained compounds. The DSC diagram and B3LYP-converged geometry and the antibacterial activities of the title complex. CCDC 2061208 and 2087028. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4nj03197k |
|
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.