ssrn-4937275
ssrn-4937275
ssrn-4937275
iew
investigation, molecular docking, antioxidant and antimicrobial screening
Abdulbasit Anoze Aliyu1, 2, Joshua Ayoola Obaleye1*, Abdullahi Ola Rajee1, Rawlings
A. Timothy3, Favour A. Nelson3, Monu Joy4
v
1Department of Chemistry, Faculty of Physical Sciences, University of Ilorin, Ilorin, Nigeria.
re
2Department of Chemistry, Faculty of Natural Sciences, Kogi State (Prince Abubakar Audu)
University, Anyigba, Kogi, Nigeria
Abstract
ot
A novel mixed ligand Cd(II) complex [Cd(nap)2(imd)3] bearing naproxen and imidazole ligands have
been synthesized. Characterization of the mono-nuclear complex was carried out via: 1H & 13C-NMR,
FTIR, UV-vis and MS and the structure validated with SC-XRD studies. The complex forms a
tn
monoclinic crystal structure with space group P21 and a = 13.581(8) Å, b = 8.290(5) Å, c = 16.693(10)
3
Å, a = γ = 90°, b = 109.352(18)°,V = 1773.1(18) Å . FTIR data confirmed the binding of the
naproxenato ligand via the deprotonated oxygen atom to the cadmium (II) ion. The coordination
rin
sphere around the Cd(II) center is a 7-coordinate system with a distorted pentagonal bipyramidal
geometry having a CdN3O4 chromophore with bidentate ligation of the carboxylate ligand naproxen.
Docking with the ADP-ribosyl transferase binding component protein (PDB: 6v1s) gave a binding
ep
affinity of -9.8 kcal/mol for the complex [Cd(nap)2(imd)3], establishing eight hydrogen bonds, with
the shortest bond length observed in serine. It exhibited greater binding affinity (-9.5 kcal/mol) and
established a higher number of hydrogen bonds compared to azithromycin. Tyrosine A:100 exhibited
Pr
the shortest bond length among the six hydrogen bond interactions detected during docking with
prostaglandin reductase 2 (PDB: 2zb7), resulting in a binding affinity of -8.6 kcal/mol. In comparison,
ibuprofen demonstrated lower binding affinity (-7.6 kcal/mol) and formed fewer hydrogen bonds.
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
Overall, molecular docking results strongly suggest that [Cd(nap)2(imd)3] exhibits high binding
affinities with both ADP-ribosyltransferase binding component and prostaglandin reductase 2,
ed
indicating its potential as a candidate for treating bacterial infections and inflammation. The complex
exhibited higher radical scavenging activity than the unbound ligands but was less than the ascorbic
acid standard in the in-vitro antioxidant experiment. The complex showed promising inhibitory action
iew
against some bacterial and fungal strains. The current results may be due to the combined impact of
the ligands and the metal ion during complex formation.
v
re
1.0 Introduction
Metal complexes have garnered a noticeable attention in the expansive field of coordination
chemistry, offering diverse applications ranging from ion exchange, hydrogen storage, and catalysis
er
to biological and medicinal functions [1-6]. The intricate synergy between ligands and metal ions
within these complexes performs a major function in ascertaining their properties and broadening
pe
their applications. Thus, the coordination behavior of a metal ion goes a long way in influencing the
biological activity of the resulting complex [7-10]. Carboxylic acid-derived ligands, a prominent
category in coordination chemistry, have garnered substantial interest owing to their diverse modes
ot
of coordination as well as robust association with heteroatomic ligands containing nitrogen atoms via
hydrogen bonding [11-13]. This versatility enables the design of intriguing structures with properties
tn
fine-tuned for various practical applications. Although there has been significant advancement in
particularly controlling experimental conditions viz: ratio of metal to ligand, reaction temperature,
pH, and choice solvent resulting to the obtained structures, have proven to be persistently tasking
[14,15]. Ligands such as imidazole and pyridine, frequently investigated as co-ligands with
ep
carboxylic acids, add another layer of complexity and versatility to these systems [16-18].
inhibiting cyclooxygenase enzymes, NSAIDs, including naproxen (Fig.1a), with IUPAC name (S)-2-
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
(6-methoxynaphthalen-2-yl) propanoic acid, exert their effects by blocking prostaglandin production
ed
responsible for pain and inflammation [19,20]. The therapeutic applications of NSAIDs span various
conditions, from osteoarthritis and rheumatoid arthritis to headaches, migraines, and dysmenorrhea
[21,22]. Notably, the interaction between NSAIDs or metal-NSAID complexes and DNA has been of
iew
profound significance due to their potential anticancer, antioxidant, and anti-inflammatory activities
[16,17, 23-25]. The primary clinical application of NSAIDs lies in their efficacy as anti-
v
The combination of NSAIDs with transition metal ions has been a notable area of interest,
re
exploring their potential therapeutic applications. Aspirin, Naproxen, and Ibuprofen among NSAIDs
eminently acts as coordination ligands in binding to metal ions to form stable complexes. Naproxen,
known for its painkilling and anti-inflammatory and properties, is widely employed as a
er
prominent ligand in several metal complexes [26-29]. Additionally, imidazole (Fig.1b), a versatile
pe
ligand known for its coordination ability and biological relevance [30-33], complements naproxen in
forming mixed ligand complexes. The exploration of mixed ligand complexes has become an
intriguing avenue for researchers, with the potential to offer enhanced stability and tailored
reactivity.
ot
Cadmium, a transition metal, presents an intriguing platform for the design of mixed ligand
tn
complexes, sharing some similarities with zinc in terms of coordination preferences due to their
investigations into their potential applications in various biological contexts [23,34-38]. While recent
studies have explored the cytotoxic effects of cadmium-containing compounds on cancer cells,
antimicrobial, antiviral, and antioxidant investigations have revealed their potential applications in
ep
combating microbial and viral infections, as well as mitigating oxidative stress [39-41]. However, a
comprehensive understanding of the crystal structure, quantum chemical properties, and molecular
Pr
interactions of cadmium complexes with mixed ligands is essential for elucidating their potential
biological activities.
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
This manuscript presents a comprehensive investigation into the cadmium complex formed by the
ed
mixed ligands naproxen and imidazole. Naproxen, a member of the phenylalkanoic acid class,
emerges as a pivotal ligand, contributing to the expanding knowledge of mixed ligand metal
complexes with potential applications in medicinal chemistry. The integrated use of crystallography,
iew
quantum chemistry, and molecular docking provides a multifaceted perspective on the cadmium
complex, paving the way for further exploration of its pharmacological properties. Our interest in
this study is driven by the scarcity of structural features for cadmium complexes formed with
v
naproxen and imidazole ligands, as this is the first of such complex to be reported, and the potential
re
benefits derived from exploring such mixed ligand systems. Through this investigation, we aim to
elucidate the structural and electronic features that underlie the bioactivity of the cadmium complex,
thereby contributing to the growing understanding of mixed ligand metal complexes and their
er
applications in medicinal chemistry.
pe
ot
tn
2.0 Experimentation
rin
All reagents are of reagent grade and were used as obtained without any further purification.
ep
DMSO are products of Sigma-Aldrich, USA. NaOH and Cd(OAc)2.2H2O are products of Merck
anthracis, Salmonella typhi, and Escherichia coli were used for the antimicrobial screening.
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
2.2 Synthesis of the complex
ed
The complex has been successfully synthesized via the direct reaction of naproxen (nap) with
imidazole (IMD) in a CH3OH/H2O solvent system (Scheme 1). A 10 mL methanolic solution of nap
(0.4 mmol, 92.1 mg) and NaOH (0.4 mmol, 16 mg) was subjected to continuous stirring for 1 hour at
iew
room temperature. Following this, (0.2 mmol, 53.3 mg) solution of Cd(OAc)2.2H2O dissolved using 5
mmol (27.2 mg) of imidazole was slowly incorporated into the reaction mixture. The entire mixture
v
was subjected to continuous stirring for an additional duration of 5 hours at ambient temperature.
re
After a gradual process of solvent evaporation for five days, colorless crystals were obtained, which
were deemed appropriate for study using SC-XRD. The crystals were harvested from the mother
liquor, subjected to air drying, and, after that, preserved for subsequent examination. Scheme 1
er
displays the synthesis pathway for the complex.
pe
[Cd(nap)2(imd)3]: Colourless crystals, Yield: 76 %, M.W = 775.13, m.p = 175 ᣞC, FTIR (cm-1):
3406, 3126, 3054, 2939, 2846, 2615, 2425, 2323, 2275, 2190, 2054, 2010, 1912, 1771, 1676, 1554,
1492, 1399, 1259, 1212, 1162, 1115, 1063, 1023, 925, 890, 853, 813, 751, 689, 650, 575, 528, 471.
Elemental Analysis (C37H38CdN6O6): Calc. (%): C, 57.33; H, 4.94; N, 10.84, found (%): C, 57.38; H,
4.88; N, 11.02. UV-Vis: λmax (nm): 265, 273, 321, 335, TOF-MS: 777.20 [M+2]+, 1H-NMR (400
ot
MHz, DMSO- d6) (δ/ppm): 12.03 (s, NH(1)-imd), 7.73 (s, H2-imd), 7.66 - 7.71 (m, 1H, Ar-H), 7.68 (dd,
1H, Ar-H), 7.47 (d, 1H, Ar-H), 7.24 (d, H4-imd), 7.10 (d, 1H, Ar-H), 7.08 (d, 1H, Ar-H), 7.01 (s, H5-
tn
imd), 3.86 (s, 3H, OCH3), 3.63 (q, 1H, -CHCH3), 1.77 (s, 3H, CH3), 1.40 (dd, J = 6.8, 5.6 Hz, 3H). 13C-
NMR (100-125 MHz, DMSO- d6) (δ/ppm): 176.64 (C(14) & C(15)) (-COO), 157.08 (C(2) & C(26))
(C-OMe), 140.28 (C(29), C(32) & C(35)) (Cimd(2)), 133.19 C(10) (2-napthalene), 129.38 (C(5)) (2-
rin
napthalene), 128.92 (C(9)) (2-napthalene), 127.80 (C(30), C(33) & C(36)) (Cimd(5)), 126.46 (C(8))
(2-napthalene), 125.60 (C(31), C(34) & C(37)) (Cimd(4)), 118.56 (C(3)) (2-napthalene), 106.14 (C(7))
(2-napthalene), 55.55 (C(1) & C(28)) (CH3-O), 47.28 (C(12) & C(16)) (CH-COO), 20.63 (C(13) &
C(17)) (CH3-C-COO).
ep
Pr
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
ed
iew
Scheme 1: The Synthetic pathway for obtaining the complex [Cd(nap)2(imd)3]
v
re
2.3 Instrumentation and physical determinations
FTIR spectra were obtained using a Bruker Alpha-II Platinum ATR spectrometer within a
range 4000 - 400 cm-1. The electronic absorption spectra were meaured in a Perkin Elmer lambda 25
er
UV/VIS spectrophotometer. NMR spectra were obtained using a Bruker Avance III spectrometer
using TMS as an internal standard with chemical shifts (δ) presented in ppm. The percentage of
pe
Carbon, Hydrogen and Nitrogen present was obtained using a Thermoscientific Flash 2000 organic
X-ray diffraction data from a single crystal of [Cd(nap)2(imd)3] was collected on a Bruker
tn
APEX-II X-ray diffractometer that has a Mo IμS microfocus X-ray source (λ = 0.71073 Å) and a CCD
detector. The measurement was conducted at 100 K utilizing an Oxford Cryosystems low-temperature
rin
device. The APEX-IV software suite was utilized for data collection, cell refinement, and data
reduction. Absorption adjustments were implemented with using SADABS [42]. The space group
assignments were determined by the investigation of systematic absences, E-statistics, and iterative
ep
refinement of the structures. The structure was identified by determined using a dual-space technique
using SHELXT [43] and then improved through full-matrix least squares with SHELXL [44].
Pr
Anisotropic displacement parameters were used to refine all atoms except hydrogen. The C-H atoms
were positioned based on idealized geometry and adjusted with fixed isotropic displacement
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
parameters using a riding model. CCDC 2323501 contains the supplementary crystallographic data
ed
for this paper. These data can be obtained free of charge from the Cambridge Crystallographic Data
iew
2.5 Computational Study
The structure under investigation in calculations utilizing density functional theory (DFT) was shown
v
using Gauss View 6.0.16. [45], while the Gaussian 16 suite of programs was utilized in conducting all
re
computations [46]. The ωB97XD/def2svp method was employed for structural optimization, as
illustrated in Fig.4. By using Chemcraft 3.0, the optimized structure as well as the molecular orbital
were visualized. er
Furthermore, the NBO 7.0 module was integrated into Gaussian 16 program to conduct Natural Bond
Orbitals (NBO) analysis [47]. This allowed for the exploration of orbital contributions, distribution of
pe
atomic charge as well as contributions to stabilization energy and charge transfer. The examination of
non-covalent interactions (NCI) was done using the Multiwfn 3.7 package [48]. To provide a visual
representation, 3-dimensional iso-surfaces were generated via VMD [49], along with 2-dimensional
ot
RDG maps presented using GNU plots. A delve into crystallographic analysis ensured the
investigation of both inter and intra-molecular interactions. Visualization of the crystal structure was
tn
done using Hirshfield surface analysis by the generation of fingerprint and Hirshfield surface plots
Molecular docking was used to simulate how a small molecule interacts with proteins at the atomic
ep
level. This allowed us to observe how these molecules behave inside the target protein’s binding site
while helping to deduce the essential biological processes [51]. The three primary objectives of
Pr
molecular docking include predicting the structure of the ligand-receptor, visualizing the
conformation as well as ranking such conformations using a scoring algorithm. The selected protein
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
from Chloridoids difficile (PDB:6v1s) The ADP-ribosyl transferase binding component (ARC) is a
ed
protein that is essential for the virulence of Chloridoids difficile and belonging to the ADP-ribosyl
transferase protein class, which are responsible for the ADP-ribosylation of other proteins [52]. ADP-
ribosylation is a post-translational modification that can alter the function of a protein. In the case of
iew
C. difficile, ARC is responsible for ADP-ribosylations in the Rho family of GTPases. This ADP-
ribosylation leads to the activation of Rho, which in turn causes the actin cytoskeleton to reorganize
[53]. This reorganization of the actin cytoskeleton is essential for the formation of the C. difficile
v
biofilm, which is a protective layer that allows the bacteria to survive in the environment. Also, a
re
protein from Homo sapiens, Prostaglandin reductase 2(PDB: 2zb7), is a pro-inflammatory mediator
that is involved in the initiation and perpetuation of inflammation. PGD2R activation leads to the
release of inflammatory cytokines and chemokines, which recruit inflammatory cells to the site of
er
inflammation. PGD2R also promotes the production of prostaglandin E2 (PGE2), which is another
pro-inflammatory mediator [54], was obtained from Protein Data Bank (PDB) (www.RSCPDB.org)
pe
and assembled on Biovia Discovery Studio 21 for analysis [55]. This included deleting native ligands
linked to the proteins from the PDB, adding missing hydrogen atoms, and removing water molecules.
The produced proteins, our chemical, and reference medications were then entered into Pyrx [56] for
ot
docking. Transformation of the proteins into macromolecular layout, while grid box argument helps in
centrally defining the crystal structures. The results of docking were estimated based on binding
tn
affinity, and conformations exhibiting some interactions via the active-site residues were carefully
selected through extensive research utilizing PyMOL and Discovery Studio. [57]. In increasing the
validity of the results, unique attention was given to elucidating the target protein’s interaction with
rin
the compounds. Crucially, Discovery Studio was used to do a thorough investigation of these
interactions, which included alkyl interactions, hydrogen and halogen bonds and Van der Waals
ep
forces. An important component of biological systems, hydrogen bonding, was specifically addressed
[58]. The tendency of the protein ligand interactions to recover from perturbations was inferred from
the bond length of H-atom; with shorter lengths suggesting stronger interaction between the ligand
Pr
and the protein. The thorough molecular docking method offered insightful information about the
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
2.6 Biological studies
ed
2.6.1 Antioxidant activity
iew
Each 2 mL containing 1000, 500, 250, 125, 62.5 µg/mL of the test sample and ascorbic acid
(standard) was added to a methanolic solution of DPPH (2.0 mL of 0.3 mM solution). The mixture
obtained was left to remain in the dark room for 30 mins after which absorbance was read at 517 nm
[59]. The experiments were done in triplicate, and average readings were taken. The percentage
v
scavenging activity for the radical (RSA) is calculated using formulae
re
er
Where Ac =absorbance of the control and As =absorbance of the sample
pe
To determine the IC50 (half-maximal inhibitory concentration) values [60], a plot of the Log of
concentration (μg/mL) against percentage inhibition of the samples was generated, and IC50 values
The phosphomolybdate assay was employed in evaluating the total antioxidant capacity of the
tn
compound [61]. Briefly, the method involves the addition of 3.0 mL of the phosphomolybdate reagent
(4.0 mM (NH4)2MoO4, 0.6 M H2SO4, and 28 mM Na3(PO4)) into a test tube containing 0.3 mL of test
rin
sample. The test tube and its content were after that enclosed with aluminium foil and incubated for
90 mins at 95 °C. Following its cooling to ambient temperature using ice, each solution’s absorbance
was recorded against a blank at 765 nm. All measurements were done in triplicate using ascorbic acid
ep
as a standard.
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
The antibacterial and antifungal activity of the compounds was evaluated using the
ed
agar well diffusion technique. The inhibitory action of the samples was evaluated against
(gram +) (Bacillus anthracis and Staphylococcus aureus), (gram –) (Salmonella typhi and
Escherichia coli) and fungal (Aspergillus niger and Candida albicans) strains. The samples
iew
were dissolved in DMSO. Accordingly, following the incubation at 37 °C, 24 and 48 hrs for
the bacteria and fungi, the inhibition zone diameters (mm) formed around each well
v
re
3.0 Results and Discussion
The asymmetric unit comprises three neutral imidazole units and two units of ligands contributing -2
pe
charges to a Cd(II) ion with the atoms situated in crystallographically separate locations. The Cd(II)
ion is a 7-coordinate system with a distorted pentagonal bipyramidal geometry. The distance between
carboxylate oxygen atoms and Cd(II) ranges from 2.42 to 2.498 Å. The naphthyl rings of both ligands
ot
are situated almost perpendicular to each other [∠C22–C23–C18–C9: 168.3(11)º]. Despite the non-
existence of significant π–π stacking interactions, [Cd(nap)2(imd)3] exhibits strong classical and non-
tn
classical intermolecular hydrogen bonding interactions. The N–H groups in all three imidazoles can
play the role of potential H-bond donors, whereas the oxygen atoms of the carboxylate group can act
rin
Å; N5–H5⋯O4(x, y-1, z): 2.839(15) Å]. The crystallographic information is presented in Table
ep
Pr
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
ed
Table 1. Crystal data and structure refinement for [Cd(nap)2(imd)3]
iew
Empirical formula C37H38CdN6O6
Temperature 100(2) K
v
Wavelength 0.71073 Å
re
Crystal system Monoclinic
Space group P 21 er
Unit cell dimensions a = 13.581(8) Å a= 90°.
pe
b = 8.290(5) Å b= 109.352(18)°.
c = 16.693(10) Å g = 90°.
Volume 1773.1(18) Å3
ot
Z 2
tn
F(000) 796
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
Independent reflections 7399 [R(int) = 0.2030]
ed
Completeness to theta = 25.242°100.0 %
iew
Data/restraints/parameters 7399 / 915 / 455
Goodness-of-fit on F2 1.031
v
re
R indices (all data) R1 = 0.1268, wR2 = 0.1766
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
Fig. 2: An ORTEP diagram for the asymmetric unit of [Cd(nap)2(imd)3] with atom numbering. H -
atoms are omitted for clarity.
ed
v iew
re
er
pe
ot
tn
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
It is established that shorter bond lengths indicate better reactivity between the atoms involved
ed
compared to longer bond lengths, which are more stable [64]. In this study, the various atoms in this
complex are studied based on their bond lengths and are compared to experimental values calculated
from the ChemDraw 3D PDB file. From our analysis, Cd43 is being used as the central metal atom,
iew
whereas experimentally, Cd1 is used. The closest atoms to the Cd central metal atom, which are
nitrogen and oxygen are being studied closely. The bond length of Cd43-O20 is 2.416 Å according to
Table 2, however the experimental bond Cd1-O3 has a bond length of 2.42 Å, which closely aligns
v
with our study results. Cd43-N58 has a bond length of 2.413 Ǻ while experiment Cd1-N10 has a value of
re
2.348 Ǻ, which is somewhat less compared to our obtained theoretical values results. The bond
lengths indicate that the link between Cd and N and O atoms is about equal in magnitude. The bond
between both atoms and the Cd bond length is at least 2.0 Ǻ, which signifies a double bond. Secondly,
er
other bonds that the analyzed system has in common with the experiment are also being studied. N51-
C53 has a magnitude of 1.318 Ǻ while the experimental N11-C17 is 1.316 Ǻ, which shows a very
pe
minimal difference of 0.002 Ǻ. The H82-O79 has a value of 2.085 Ǻ, and the experimental H82-O72 is
2.905 Ǻ, which is a very large difference from our result. This significant disparity is likewise
comparable to the C-H bond designations and the N-H bonds. Now, in terms of magnitude, the Cd-O
ot
bond has the highest magnitude, while the N66-H57 has the smallest magnitude of 1.013 Ǻ. Also, it can
be deduced that N-C, C-H and N-H bonds are single bonds, while Cd-O and Cd-N are double bonds.
tn
Table 2 presents the selected theoretical and experimental bond lengths of the synthesized
[Cd(nap)2(imd)3] complex.
rin
Table 2. Examined bond lengths of the complex via theoretical and experimental.
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
ed
v iew
re
er
pe
ot
tn
rin
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
3.3.1 Hirshfeld Surface Analysis
ed
Hirshfeld surface analysis is a computational approach used in drug design to understand the
intermolecular interactions between a therapeutic molecule and its target. [65]. This method relies on
the Hirshfeld partitioning of a crystal structure's electron density, enabling the visualization and
iew
quantification of individual atoms' contributions to intermolecular interactions [66]. The research
offers useful insights into the characteristics and intensity of several contacts, including electrostatic,
hydrogen, and van der Waals interactions such as hydrogen, van der Waals, and electrostatic
v
interactions. Moreover, it aids in discovering possible binding sites on the target protein that are not
re
easily visible in the crystallographic structure. It helps in defining the arrangement of molecules and
recognizing possible variations or combined crystal structures that may affect the drug's effectiveness
interactions with distances less than the sum of van der Waals radii are represented by red patches.
Van der Waals interactions are shown by white areas, whereas lengthier contacts are represented by
ot
blue areas [68]. Fig.5 shows that the red color represents N-H…O intermolecular interaction. The
distances from a location on the Hirshfeld surface to the closest atoms outside and within the surface
tn
are labeled as de and di, as seen in Fig.6. The interactions in the complex and their percentage
abundance are as follows: C…C 0.6 %, C…H 15%, C…O 0.1%, C…N 0.3%, H…C 12.9%, H…H
51.2%, H…N 1.7%, H…O 7.4%, N…H 1.9%, N…N 0.3%, N…O 0.2% and O…H 7.9%. The data
rin
obtained so far indicate that the H…H interaction is the prevalent force in the compound, suggesting
high selectivity, potency and binding specificity towards the target protein.
ep
Pr
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
ed
v iew
re
er
pe
ot
tn
rin
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
ed
v iew
re
er
pe
ot
tn
rin
ep
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
ed
3.4 Vibrational analysis
iew
Vibrational analysis is a significant spectroscopic technique used in numerous scientific domains like
chemistry, physics, and materials science. This method involves examining the oscillatory patterns
within molecules and materials to gain valuable insights into their arrangement, composition, and
v
behavior [69]. Vibrational spectroscopy provides scientists with precise instruments for identifying
re
and characterizing molecules, offering detailed information about their structural characteristics. By
analyzing these vibrational modes, scientists can infer details such as bond lengths, angles, and
symmetry within a molecule. The crucial role of vibrational analysis lies in unraveling the
er
complexities of chemical bonding within molecules, enabling the differentiation between different
types of chemical bonds (single, double, or triple) and the identification of functional groups. This
pe
information is fundamental for understanding chemical reactions and reactivity [70]. Vibrational
analysis is commonly utilized in drug research and development within the pharmaceutical sector. It
aids in assessing the purity of drug compounds, elucidating the structures of complex biomolecules,
ot
and monitoring the progress of chemical reactions in drug synthesis processes. Vibrational
spectroscopy also extends its application to biomedical research and diagnostics, proving invaluable
tn
in studying biological entities such as proteins, DNA, and lipids. Additionally, it serves as a reliable
tool for detecting abnormalities in tissues, thereby significantly contributing to disease diagnosis and
advancements in medical research [71]. The theoretical vibrational frequencies and assignments for
rin
C-H Vibration
Previous studies have recorded that the spectral range of 3000-3100 cm-1 encompasses aromatic and
Sp2 C-H stretching vibrations, whereas Sp3 C-H stretch vibrations typically occur in the range 3000 -
Pr
2850 cm-1 [72]. Notably, the aromatic and Sp2 C-H vibrations, as observed in this compound, span
3100.87-3314.88 cm-1. Similarly, the Sp3 C-H stretching vibrations occurred within the range
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
3030.72-3185.57 cm-1. Crucially, one can infer that the theoretically calculated C-H stretching
ed
vibrations match the values reported in previous researches.
N-H Vibrations
iew
The N-H stretching vibrations usually falls between 3500 - 3300 cm-1 in the infrared spectrum [73].
Observably, the theoretical N-H stretching vibrational frequency was observed around 3640.35-
3679.79 cm-1. The estimated frequencies somewhat exceed the stated values, although they are in
v
close alignment. This confirms the stable geometry conformation of the study complex.
re
C=C Vibrations
The documented literature indicates that the range for C=C stretching vibrations in IR spectroscopy
er
usually ranges from 1680 - 1600 cm-1 [74]. The theoretical C=C vibrations were observed around
1734.40 - 1656.24 cm-1. Notably, the theoretical values align with the reported values.
pe
C=O Vibrations
The stretching vibration frequency range for the C=O bond (also known as carbonyl stretching)
ot
typically falls between 1600 and 1800 cm⁻¹ [75]. In this present study, the C=O stretching vibrations
in this investigation falls within 1688.60 - 1681.14 cm-1. The calculated C=O vibrations fall within the
tn
C=N Vibrations
rin
The characteristic region for the C-N stretching vibrations in IR spectroscopy, as reported by existing
literature, typically falls between 1400 - 1200 cm-1. Observably, the C-N stretching vibrations were
ep
seen around 1559.45-1605.91 cm-1 [76]. The reported range for the C-N stretching vibrations is
slightly lower than the theoretical values. However, the result of the calculated vibrational frequencies
for all functional groups present in the study metal complex agrees with reported values from the
Pr
literature. Hence, the compound's selectivity, potency and high binding specificity towards the target
protein is affirmed.
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
3.5 Investigation of electronic property
ed
3.5.1 HOMO-LUMO and Quantum Descriptor Parameters
iew
system. Various molecular properties such as HOMO, LUMO, energy gap, ionization energy, electron
affinity, electronegativity, hardness, chemical potential, electrophilicity, and Fermi energy were
analyzed to understand the level of stability and reactivity of the studied complex [77]. The HOMO
energy is determined to be -7.60 eV, and the LUMO energy is determined to be -0.36 eV based on the
v
data in Table 3. The energy gap (Eg) is a parameter that measures the conductivity of a system, telling
re
us if the system is metallic, semi-metallic or possesses insulating properties [78]. The compound's
energy gap is determined to be 7.24 eV, indicating that it belongs to the category of a semiconductor
er
or an insulator. This indicates that the electron transfer from the HOMO to the LUMO is inefficient,
resulting in a system that is less reactive and more stable. The ionization potential (IP) is 7.6 eV,
pe
signifying that a significant quantity of energy is needed to ionize an electron from the atom. The
electron affinity (EA) is low at 0.36 eV, indicating that the system does not readily attract electrons.
The electronegativity (σ) of 0.13 eV is very low. The hardness (η), which is the degree to which a
system can withstand distortion, has a magnitude of 3.98 eV, indicating a moderate ability to
ot
withstand pressure and other factors. The electrophilicity (ω) with magnitude 1.64 eV is also very
tn
low. Lastly, the Fermi energy level (EFL) of 3.62 eV is explained with respect to the energy gap. The
gap between the Fermi energy level and the energy gap is around 4.35 eV, which is a very wide gap,
further confirming the limited electron transport separating the VBM and the CBM.
rin
Table 3: The HOMO energy, LUMO energy, energy gap (Eg), and quantum descriptor
characteristics in electron volts (eV) of the analyzed complex.
ep
[Cd(nap)2(imd)3] -7.60 -0.36 7.24 7.60 0.36 0.13 3.98 -3.62 1.64 3.62
Pr
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
3.5.2 Natural Bond Orbital (NBO) Analysis
ed
The analysis of Natural bond orbital (NBO) is commonly used to understand the computational
solution of the Schrödinger equation and simplify chemical bonding ideas. NBO analysis is the most
successful method for identifying inter- and intra-molecular orbital exchanges within a complex.
iew
Furthermore, NBO research offers a clear insight into the delocalization of electron density occurring
between filled donor (Lewis) and unfilled acceptor (non-Lewis) natural bond orbitals, highlighting a
more stable donor-acceptor interaction. [79]. The transitions were recorded based on the atomic
v
composition of the transition without taking into account their perturbation energy. Second order
re
perturbation energy refers to the energy needed in breaking a bond in a system. From Table S15, the
C-C donor to C-C acceptor possesses the highest perturbation of 30.99 kcal/mol. This single transition
can partially account for the fact that C-C transfer in both donors and acceptors are bonds that are
er
harder to break compared to other transitions. The transition of lone pair nitrogen donor to lone pair
cadmium has a disturbance of 20.7 kcalmol-1, which is seconds to the C-C transition. The transition
pe
LP(2)O20-LP*(6) Cd43 has a perturbation energy of 16.49 kcal/mol, followed by the transition from
O42 to the same cadmium atom at an almost similar energy of 14.61 kcal/mol. The least perturbation
energy is recorded in the transition from C-O donor to Cd lone pair, which is 4.72 kcal/mol. In
ot
conclusion, these transitions, bonds and energies are studied to give in-depth insight into the
Non-covalent interactions (NCI), or Reduced Density Gradient (RDG) analysis, is the study of the
forces and interactions between atoms in a molecule, including both intra and intermolecular
interactions [80]. The main examples of NCI include hydrogen bonding, van der Waals forces,
ep
hydrophobic interactions and electrostatic interactions [81]. In the case of our [Cd(nap)2(imd)3]
complex, both top and side views are captured to facilitate a more comprehensive visual examination.
Pr
The structural maps' green iso-surfaces show non-directional attraction, a symptom of weak
interactions brought on by van der Waals dispersion forces [82]. Strong contacts, primarily due to
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
steric repulsion, are represented by the red iso-surfaces and have a significant impact on the
ed
conformation and reactivity of ions and molecules. Analysis of Fig.S16 reveals that green surfaces are
dispersed throughout the compound, suggesting relatively weak forces holding it together. This
outcome is advantageous as such weak attraction facilitates the compound's easy dissolution in
iew
various biological systems, enabling faster diffusion into the bloodstream.
v
Understanding molecular interactions and locating potential sites for electrophilic and nucleophilic
assaults in a molecular system depend on the knowledge of the molecular electrostatic potential, or
re
MESP. [83]. This possible map is visually represented by a variety of colors. The red tint on the
MESP map indicates the most negative region, which is favorable to electrophilic attacks, according
er
to results in the literature [68]. Conversely, the blue color represents the most positive area,
encouraging nucleophilic assaults. Lighter shades of blue indicate slightly less positivity than deeper
pe
shades. Yellow regions on the map denote areas with less negative potential relative to the red areas.
Analysis of Fig.6 reveals that compound [Cd(nap)2(imd)3] exhibits greater nucleophilic tendencies
than electrophilic ones. The MESP color scheme aids in distinguishing between electrophilic and
ot
nucleophilic regions, clearly showing that the electrophilic portion of [Cd(nap)2(imd)3] is significantly
smaller than its nucleophilic counterpart. The observation is backed by data gathered from the FMO
tn
table, showing an estimated electrophilicity value of 1.64 eV, which is significantly low and aligns
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
ed
v iew
re
Fig.6: Molecular electrostatic potential map showing preferred locations for nucleophilic and
electrophilic attacks.
er
pe
3.6 Molecular docking (Antimicrobial and Anti-inflammation)
Molecular docking aims to identify the ligand's conformation that reduces the system's energy. A
scoring function is utilized, which is a mathematical formula that considers the many forms of
ot
interactions between the two molecules [58]. The scoring function is typically based on the principles
of quantum mechanics, but it can also be based on more empirical methods. Once the scoring function
tn
has been defined, the docking program will generate a number of possible conformations of the ligand
and then score each conformation [83]. The conformation with the lowest score is typically
rin
considered to be the most likely to bind to the receptor. Table S17 presents the docking outcomes,
revealing that when the structure [Cd(nap)2(imd)3] is docked with the ADP-ribosyl transferase binding
component protein (PDB: 6v1s), it shows a binding affinity of -9.8 kcalmol-1, forming eight hydrogen
ep
bonds, with serine having the shortest bond length. This notable abundance of hydrogen bonds may be
attributed to the compound's polarity and shape, suggesting promising antibacterial properties.
Pr
By contrast, the common antibiotic azithromycin creates three hydrogen bonds and has a binding
affinity of -9.5 kcal/mol. After examining the many interactions between [Cd(nap)2(imd)3] and 6v1s,
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
it was discovered that the stabilization of the ligand-protein complex is aided by hydrophobic pi-
ed
sigma, hydrophobic pi-alkyl contacts, and electrostatic pi-anion interactions with the amino acid
glutamine. Docking using [Cd(nap)2(imd)3] for the protein prostaglandin reductase 2 (PDB: 2zb7)
results in a binding affinity of -8.6 kcal/mol and six detected hydrogen bond interactions, with
iew
tyrosine A:100 exhibiting the shortest bond length (2.2931 Å). In comparison, ibuprofen creates two
H-bonds with a binding affinity of -7.6 kcal/mol. Additionally, electrostatic pi-anion interactions are
evident in the docking between [Cd(nap)2(imd)3] and prostaglandin reductase. The molecular docking
v
results strongly indicate that [Cd(nap)2(imd)3] exhibits high binding affinities with both ADP-ribosyl
re
transferase binding component and prostaglandin reductase 2. This suggests that [Cd(nap)2(imd)3]
holds promise as a potential candidate for treating bacterial infections and inflammation.
er
pe
ot
tn
rin
ep
Fig.7. 3D image of the studied complex and proteins obtained from the protein databank
Pr
3.7 FTIR
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
The FTIR spectra of the ligands and complex generated in the solid state are presented (Figs.S1-S3).
ed
Fig.8 represents the overlaid FTIR spectra of naproxen, imidazole and [Cd(nap)2(imd)3]. Some
characteristic spectra data (cm-1) are presented in Table 4. The stretching vibration of the carboxylic
group of free naproxen is represented by the medium band at 3384 cm-1 in its spectra. This band
iew
vanishes from the complex produced with the Cd(II) ion's spectra, signifying the carboxylate group's
deprotonation. The intense band at 1161 cm-1 corresponds to the ⱱ(C-O) stretching vibration of the
methoxy group (Ar-O-C) [84,85]. Two prominent bands at 1680 cm-1 and 1384 cm-1 were identified
v
as the vibrations of the C=O and C-O bonds in the carboxylic acid (-COOH) group. These bands in
re
the spectra of the complex, were replaced by strong bands at 1554 cm-1 due to the antisymmetric
ⱱas(CO2) and 1399 cm-1 assigned to the symmetric ⱱs(CO2) stretching vibrations of the naproxenato
ligand. The difference ∆v = [ⱱas(CO2) - ⱱs(CO2)] is a crucial tool for identifying the coordination
er
modes in a carboxylate ligand. As shown in Table 4, ∆v = 155 cm-1 corresponding to bidentate mode
of coordination, of the _COOˉ group [86,87]. This conclusion aligns with the findings from the X-ray
pe
investigations. The ⱱ(N-H) band at 3109 cm-1 in the unbound imidazole ligand shifted to 3126 cm-1
upon coordination in the complex. The binding of the imidazole ligands to the Cd(II) ion was verified
by the presence of a moderate band at 751 cm-1. This band is typically associated with the out-of-
ot
plane ρ(C-H) vibration [88,89]. The new bands observed at 650 cm-1 and 471 cm-1 in the spectra of
[Cd(nap)2(imd)3] were identified as the vibrations related to (Cd-N) and (Cd-O) bonds, respectively.
tn
These bands were not present in the free ligands. All assignments have been completed based on the
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
ed
v iew
re
er
Fig.8: Superimposed FTIR spectra of naproxen, imidazole and [Cd(nap)2(imd)3]
pe
ot
Table 4: Selected FTIR absorption bands (cm-1) of the ligands and complex
=(ⱱas - ⱱs)
In a DMSO solution, the ligands and the Cd(II) complex's absorption spectra were obtained between
200 and 700 nm. The spectrum of naproxen displays two pairs of bands at 264 and 274 nm, and 319
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
and 333 nm, which were identified as corresponding to the 𝝿 → 𝝿* and n → 𝝿* transitions,
ed
respectively. The peak at 275 nm in the free imidazole ligand corresponds to the 𝝿 → 𝝿* transition of
the aromatic ring. The electronic spectra of [Cd(nap)2(imd)3] displays absorptions that closely
resemble those of the individual ligands. A small change in the absorption peak is observed, possibly
iew
resulting from the ligands binding to the Cd(II) metal core. This verifies the lack of d - d transitions,
as anticipated for d10 ions such as Cd(II), where all d-electrons are paired up [87, 92].
3.9 NMR
v
re
NMR spectroscopy is used to analyze the behavior of the complex in solution, confirming the
formula and purity of the chemical. The ligands and complex’s 1H-NMR spectra were obtained in a
DMSO-d6 solution and are shown in Fig.S7-S9. The lack of a signal at 9.25 ppm, which corresponds
er
to the carboxylic hydrogen of the free naproxen in the 1H-NMR spectra of the complex, may indicate
the presence of the deprotonated ligand in the complex [93,94]. The additional signals of the
pe
naproxenato ligand exhibit upfield or downfield shifts after interacting with the Cd(II) ion, as
anticipated. Aromatic proton multiplets are seen between 7.71 and 7.81 ppm [95]. A singlet peak at
12.03 ppm is attributed to the N-H proton, while three more signals at 7.73, 7.24, and 7.01 ppm are
ot
The ligands and complex’s 13C-NMR spectra were obtained in a DMSO-d6 solution and are
tn
shown in Fig.S10-S12. The absorption signal of the carboxylic carbon (C14) shifted slightly from 175
ppm in the free ligand to 176 ppm in the complex, suggesting electron transfer from the carboxylate
rin
anion COO- to the cadmium metal center, resulting in de-shielding of the carboxylate carbon.
Additionally, the following carboxylate carbons have a noteworthy downfield chemical shift in
relation to naproxen: (C(12)) (CH-COO) (45.15 to 47.28 ppm), (C(13)) (CH3-C-COO) (18.94 to 20.63
ep
ppm) in the complex, indicating the interaction of the carboxyl group with cadmium due to
complexation [94].
Pr
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
3.10.1 Antioxidant activity
ed
The radical scavenging activity (RSA) of the compounds was screened using DDPH and
phosphomolybdate assays. In the DPPH assay (Table S18), the RSA was concentration-dependent,
with the complex performing better than the free ligands. This agrees closely with the results obtained
iew
for similar compounds [97-100]. The antioxidant ability of the tested compounds has been compared
holistically using their IC50 values (the sample concentration in µg/mL required to scavenge 50 % of
DPPH radical) (Fig. 9a). In comparison with ascorbic acid, the complex compared favorably with an
v
IC50 value of 24.41 µg/mL as against 20.55 µg/mL for the standard. The order of scavenging power
re
for the DPPH radical of the compounds is: nap < imd < [Cd(nap)2(imd)3] < AsA.
The result of the total antioxidant capacity of the compounds from the phosphomolybdate assay is
er
presented in Table S19. A similar trend to that of the DPPH assay was noticed, wherein the complex
appeared to have a better TAC value compared to the ligands. In the presence of the compound
pe
(antioxidants), a green coloration is observed, which is due to the formation of a phosphomolybdenum
(V) complex following the reduction of Mo(VI) to Mo(V) with an absorption maxima occurring at
765 nm [101-102]. Fig. 9b shows a comparison of the IC50 values of Ascorbic acid (37.59 µg/mL),
ot
[Cd(nap)2(imd)3] (112.09 µg/mL), naproxen (168.74 µg/mL) and imidazole (182.56 µg/mL) resulting
in a total antioxidant capacity order: AsA > [Cd(nap)2(imd)3] > nap > imd. Generally, the ascorbic
tn
acid standard possessed stronger antioxidant activity compared to the synthesized complex. However,
the complex performed far better than the parent ligands and can, therefore, be said to be a promising
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
ed
v iew
re
Fig. 9a&b: A Comparison of IC50 values of the ligands, complex and Ascorbic Acid
er
Standard
diameter of inhibition zones (mm). The results in Figs. 10a&b and Table 5 show that the free
naproxen was inactive while imidazole was moderately active against all the tested organisms in this
ot
study. However, the synthesized complex showed remarkable activity, especially against S. aureus
and B. anthracis, where higher activity was recorded in comparison to Ciprofloxacin. Against the
tn
fungal strains, the complex also presents a superior activity compared to the standard drug
Ketoconazole. Generally, the complex showed better activity compared to the unbound ligands. The
rin
improved activity of the complex when compared to the ligand could be due to the chelation effect
upon complexation with the metal ion. Chelation has been reported to increase the lipophilicity of
metal complexes causing their enhanced penetration of the membrane cells of the organisms
ep
[103,104].
Pr
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
Table 5: Antibacterial and antifungal activities of associated ligands and the complex
ed
Gram (+) Bacteria Gram (-) Bacteria Fungi
iew
imd 18 16 13 16 14 12
nap 0 0 0 0 0 0
[Cd(nap)2(imd)3] 28 26 20 18 22 24
Ciprofloxacin 26 22 32 26 ─ ─
Ketoconazole ─ ─ ─ ─ 20 16
v
KEY: (─) = Not Tested
re
er
pe
ot
4.0 Conclusion
The synthesis characterization with spectroscopic, physicochemical, DFT and single crystal X-ray
rin
diffraction studies of a novel Cd(II) complex bearing naproxen ligand have been carried out
successfully, and the structure has been established. Utilizing experimental and theoretical methods,
ep
including DFT computation and molecular docking, we characterized the complex's low reactivity,
high stability, and structural features, validated by close agreement between experimental and
theoretical data. Analysis of crystal packing interactions highlighted strong binding specificity
Pr
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
compared to standard drugs. These findings underscore [Cd(nap)2(imd)3]'s candidacy for future
ed
therapeutic applications targeting bacterial infections and inflammation. The biological capabilities of
the compound were assessed in vitro for its antioxidant, antibacterial, and antifungal properties. The
complex demonstrated enhanced scavenging of the DPPH free radical at varying concentrations
iew
compared to the free ligands, however its effectiveness was lower than that of the ascorbic acid
standard. A similar trend was obtained for the total antioxidant capacity. The in-vitro antimicrobial
testing against specific bacterial and fungal strains indicated that the naproxen ligand did not exhibit
v
any inhibitory effects. The combination exhibited significant antimicrobial activity against the studied
re
pathogens, particularly against S. aureus and B. anthracis, surpassing the activity of ciprofloxacin.
Hence, the compound shows promise as an antibacterial candidate that could help prevent diseases
TMS - Tetramethylsilane, FMO - Frontier Molecular Orbital, HOMO - Highest occupied molecular
ot
orbital, LUMO - Lowest unoccupied molecular orbital, VBM - Valence band maximum, CBM -
Conduction band minimum AsA-Ascorbic Acid, SC-XRD- Single crystal X-ray diffraction
tn
5.0 Declaration
Writing original draft. Joshua Ayoola Obaleye: Conceptualization, Visualization, Supervision and
ep
draft correction. Abdullahi Ola Rajee: Interpretation and discussion of results Rawlings A. Timothy
and Favour A. Nelson: Computational studies & analysis. Monu Joy: SC-XRD and structural
Pr
analysis.
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
The authors declare no conflicting interest in the course of this research.
ed
5.3 Acknowledgments
Abdulbasit A Aliyu is grateful to the Tertiary Education Trust Fund (TETFUND) of Nigeria for the
iew
Award of Bench Work Fellowship. The authors acknowledge the support of the World Bank STEP-B
Laboratory of the University of Ilorin, Nigeria and also the School of Chemistry and Physics,
University of Kwazulu-Natal, South Africa, for providing bench space. The authors would like to
v
acknowledge the Centre for High-Performance Computing (CHPC) in South Africa for providing
re
computational resources for this research project.
References
er
[1]. J.A. Obaleye, A. Lawal, A.O. Rajee, H.F. Babamale, F.B. Shittu. Synthesis, Characterization and
pe
Antimicrobial Activity of mixed Amodiaquine and Sulphadoxine mixed Ligands – metal
Complexes. NJBMB: 29 (2):(2014) 170-179.
[2]. A.E. Owen, E.C. Agwamba, M.E. Gideon, K. Chukwuemeka, E.U. Ejiofor, I. Benjamin, ...H.
Louis. Molecular structure, spectroscopy, molecular docking, and molecular dynamic studies of
ot
nickel-based complexes with H2S gas. RSC Adv., 12(47), (2022) 30365-30380.
[4]. A.O. Rajee, J.A. Obaleye, H. Louis, A.A. Aliyu, A. Lawal, I.O Amodu, A.L.E. Manicum. Single-
crystal X-ray, spectroscopy, quantum chemical calculations, and molecular docking investigation
rin
https://www.ncbi.nlm.nih.gov/books/NBK547742/ (2021).
[7]. H.F. Babamale, J.A. Obaleye, A. Lawal, A.O. Rajee, M.O. Bamigboye, M. Lawal, H.O. Sa’ad.
Mixed metal complexes of Artemether, Lumefantrine and Ascorbic acid: Synthesis,
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
characterization and biological activity studies. Centrepoint. (Science Edition). 25(2): (2019) 21-
32.
ed
[8]. A. Lawal, P.A. Ayanwale, A.O Rajee, M.O. Bamigboye, A.O. Saad, M. Lawal, H.F.
Babamale, G.G. Nnabuike, M.T. Yunus-Issa, S.A. Amolegbe. Preparation, Characterization
and Antibacterial Activity of Metal Complexes of Mixed Ligands of Nicotinamide and
iew
Metronidazole. Nigerian Journal of Biochemistry and Molecular Biology, 32(2): (2017) 82-87.
[9]. M. Lawal, J.A. Obaleye, R.N. Jadeja, V.K. Gupta, G.G. Nnabuike, M.O. Bamigboye ... A.T.
Raji. Mixed Pyrazinamide and 4-Dimethylaminopyridine Nickel (II) Complex: Characterization,
DFT, and Biological Potentials. J. Mol. Struct., (2023) 137171.
[10]. A.O. Rajee, A.A. Aliyu, H.F. Babamale, A. Lawal, S.O. Ayinla, W.A. Osunniran, J.A.
v
Obaleye. Preparation, characterization and antibacterial activity of Mn(II), Cu(II) and Zn(II)
re
complexes of methionine and 2,2-bipyridine co-ligands. JKCS, 13-1, (2020) 16-21.
[11]. N. Cristina, F.L. Adrián, F.L. Javier, C. Julia, L.C. José, L. Carlos. Synthesis, spectroscopic
studies and in vitro antibacterial activity of Ibuprofen and its derived metal complexes. Inorg.
Chem. Commun. 45, (2014) 61-65.
[12].
er
A.O. Rajee, H.F. Babamale, A.A. Aliyu, A. Lawal, S.O. Ayinla, W.A. Osunniran, A.A. Aliyu,
I. Musa. Synthesis, Structural Elucidation and Antimicrobial Activity of Metal (II) Polypyridyl
pe
Complexes of 2-Amino-4-(Methylthio) Butanoic Acid. FUJNAS. 2020; 9(1): 19-26.
[13]. H.K. Busari, L.A. Azeez, H.K. Aremu, S.O. Ayinla, L.A. Jinadu, J.A. Obaleye. Zinc and
Copper Complexes of 4-Methylbenzoic Acid and 2-Methylimidazole: Synthesis, Characterization,
Antimicrobial and Molecular Docking Studies. The Chemist, (2023) 121.
ot
[14]. Y.Q. Lin, X.M. Tian, B.X. Zhu, D.M. Chen, C. Huang. Five Porous Complexes Constructed
from a Racemic Ligand: Synthesis, Chiral Self-Assembly, Iodine Adsorption, and Desorption
Properties. Inorg. Chem, 62(30), (2023) 12099-12110.
tn
[15]. P. Zhou, R Shi, J.F. Yao, C.F. Sheng, H. Li. Supramolecular self-assembly of nucleotide–
metal coordination complexes: From simple molecules to nanomaterials. Coordin. Chem Rev,
292, (2015) 107-143.
rin
[16]. G.G. Nnabuike, S.N. Meena, A.R. Palake, K.M. Kodam, S. Salunke-Gawali, R.J. Butcher,
J.A. Obaleye. Zn (II) complexes with mefenamic acid: Synthesis, characterization, and anticancer
activity. J. Mol. Struct., 1294, (2023) 136432.
[17]. G.G. Nnabuike, S. Salunke-Gawali, A.S. Patil, R.J. Butcher, J.A. Obaleye, H. Ashtekar, B.
ep
Prakash Cobalt (II) complexes containing mefenamic acid with imidazole and pyridine based
auxiliary ligands: Synthesis, structural investigation and cytotoxic evaluation. J. Mol. Struct.,
1285, (2023) 135519.
Pr
[18]. G.A. Krishna, T.M. Dhanya, A.A. Shanty, K.G. Raghu, P.V. Mohanan. Transition metal
complexes of imidazole derived Schiff bases: Antioxidant / anti-inflammatory / antimicrobial /
enzyme inhibition and cytotoxicity properties. J. Mol. Struct. 1274, (2023) 134384.
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
https://doi.org/10.1016/j.molstruc.2022.134384
[19]. J.R. Vane. Inhibition of prostaglandin synthesis as a mechanism of action for aspirin-like
ed
drugs, Nat. New. Biol. 231 (25) (1971) 232–235.
[20]. N.M. Alorfi,. Pharmacological methods of pain management: Narrative review of medication
used. Int. J. Gen. Med, (2023) 3247-3256.
iew
[21]. L. Kuritzky, A. Weaver. Advances in rheumatology: coxibs and beyond. J. Pain Symptom
Manag, 25(2), (2003) 6-20.
[22]. A. Lawal, J.A. Obaleye, J.A. Synthesis, Characterization and antibacterial activity of aspirin
and paracetamol- metal complexes. Biokemistri 19: (2007) 9-15.
[23]. D. Kovala-Demertzi, M. Staninska, I. Garcia-Santos, A. Castineiras, M.A. Demertzis.
v
Synthesis, crystal structures and spectroscopy of meclofenamic acid and its metal complexes with
re
manganese(II), copper(II), zinc(II) and cadmium(II). Antiproliferative and superoxide dismutase
activity. J. Inorg. Biochem, 105(9), (2011) 1187–1195.
https://doi.org/10.1016/j.jinorgbio.2011.05.025
[24]. G. Malis, E. Geromichalou, G.D. Geromichalos, A.G. Hatzidimitriou, G. Psomas. Copper (II)
er
complexes with non–steroidal anti–inflammatory drugs: structural characterization, in vitro and in
silico biological profile. J. Inorg. Biochem, 224, (2021) 111563.
pe
[25]. S. Ramos-Inza, A.C. Ruberte, C. Sanmartin, A.K. Sharma, D. Plano. NSAIDs: Old
Acquaintance in the Pipeline for Cancer Treatment and Prevention─ Structural Modulation,
Mechanisms of Action, and Bright Future. J. Med. Chem, 64(22), (2021) 16380-16421.
[26]. F. Dimiza, A.N. Papadopoulos, V. Tangoulis, V. Psycharis, C.P. Raptopoulou, D.P.
ot
Kessissoglou, G. Psomas. Biological evaluation of cobalt (II) complexes with non-steroidal anti-
inflammatory drug naproxen. J. Inorg. Biochem, 107(1), (2012) 54-64.
[27]. M.S. Hasan, N. Das. A detailed in vitro study of naproxen metal complexes in quest of new
tn
complexes with the non-steroidal anti-inflammatory drugs naproxen and mefenamic acid:
Synthesis, structure, antioxidant capacity, and interaction with albumins and DNA. New J Chem,
42(20), (2018) 16666-16681.
[29]. P. Srivastava, R. Mishra, M. Verma, S. Sivakumar, A.K. Patra. Cytotoxic ruthenium (II)
ep
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
[31]. P.G. Row, P. Biswas, A. Dutta, P. Dastidar. Design, Synthesis, and Structural Insights of a
Series of Zn(II) − NSAID Based Coordination Complex Derived Metallogels and Their Plausible
ed
Applications in Self Drug Delivery. Ii. (2023) https://doi.org/10.1021/acs.cgd.2c01041
[32]. L. Zhang, X.M. Peng, G.L. Damu, R.X. Geng, C.H. Zhou. Comprehensive review in current
developments of imidazole‐based medicinal chemistry. Med. Res. Rev, 34(2), (2014) 340-437.
iew
[33]. J. Soni, A. Sethiya, N. Sahiba, D.K. Agarwal, S. Agarwal S. Contemporary progress in the
synthetic strategies of imidazole and its biological activities. Curr. Org. Synth, 16(8), (2019)
1078-1104.
[34]. L. Tabrizi, H. Chiniforoshan, P. McArdle, M. Ebrahimi, T. Khayamian. A novel bioactive
Cd(II) polymeric complex with mefenamic acid: Synthesis, crystal structure and biological
v
evaluations. Inorg. Chim. Acta, 432, (2015) 176–184. https://doi.org/10.1016/j.ica.2015.04.010
re
[35]. V. Bansal, A. Bharde, R. Ramanathan, S.K. Bhargava. Inorganic materials using
‘unusual’microorganisms. Adv. Colloid Interfac. 179, (2012) 150-168.
[36]. M. Bai, J.B. Zhang, L.H. Cao, Y.P. Li, D.Z. Wang. Zinc (II) and cadmium (II) metal
complexes with bis (tetrazole) ligands: synthesis and crystal structures. J. Chin. Chem. Soc-Taip.,
58(1), (2011) 69-74.
er
[37]. N. Sultana, M.S. Arayne, M. Afzal. Synthesis and antibacterial activity of cephradine metal
pe
complexes: part II complexes with cobalt, copper, zinc and cadmium. Pak. J. Pharm. Sci, 18(1),
(2005) 36-42.
[38]. M. Montazerozohori, S. Zahedi, M. Nasr-Esfahani, A. Naghiha. Some new cadmium
complexes: antibacterial/antifungal activity and thermal behavior. J. Ind. Eng. Chem, 20(4),
ot
(2014) 2463-2470.
[39]. M. Rezaee, M. Montazerozohori, R. Naghiha, E.P. Kokhdan. New tetradentate long chain
Schiff base and its cadmium (II) complexes: Antimicrobial, antioxidant and anticancer activities.
tn
S.K. Belus; E.I. Kozhukhova; V.N. Krasnoperova. Zinc(II) and cadmium(II) halide complexes
with caffeine: synthesis, X-ray crystal structure, cytotoxicity and genotoxicity studies, Inorg.
Chim. Acta. 487 (2019) 184e200. https://doi.org/10.1016/j.ica.2018.11.036.
[41]. Z. Saedi; E. Hoveizi; M. Roushani; S. Massahi; M. Hadian; K. Salehi. Synthesis,
ep
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
[45]. R.D.I.I. Dennington, T.A. Keith, J.M. Millam. GaussView, version 6.0. 16. Semichem Inc
Shawnee Mission KS (2016).
ed
[46]. M.E. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, D.J.
Fox. Gaussian 16. (2016).
[47]. E.D. Glendening, C.R. Landis, F. Weinhold. NBO 7.0: New vistas in localized and
iew
delocalized chemical bonding theory. J. Comput. Chem, 40(25), (2019) 2234-2241.
[48]. T. Lu, & F. Chen. Multiwfn: a multifunctional wavefunction analyzer. J. Comput. Chem,
33(5), (2012) 580-592.
[49]. M.A. Spackman, D. Jayatilaka. Hirshfeld surface analysis. Cryst. Eng. Comm, 11(1), (2009)
19–32.
v
[50]. S.L. Tan, M.M. Jotani, E.R. Tiekink Utilizing Hirshfeld surface calculations, non-covalent
re
interaction (NCI) plots and the calculation of interaction energies in the analysis of molecular
packing. Acta. Crystallogr. E., 75(3), (2019) 308-318.
[51]. B.J. McConkey, V. Sobolev, M. Edelman. The performance of current methods in ligand–
protein docking. Curr. Sci. India. (2002) 845-856.
[52].
er
E. Ferrero, N. Lo Buono, A. Horenstein, A. Funaro, F. Malavasi. The ADP-ribosyl cyclases-
the current evolutionary state of the ARCs. Front. Biosci-Landmrk, 19, (2014) 986-1002.
pe
[53]. K. Aktories, C. Mohr, G. Koch. Clostridium botulinum C3 ADP-ribosyltransferase. ADP-
Ribosylating Toxins, (1992) 115-131.
[54]. D. Studio. Dassault systemes BIOVIA, Discovery studio modelling environment, Release 4.5.
Accelrys Softw Inc, (2015) 98-104.
ot
[55]. S. Dallakyan. PyRx-python prescription v. 0.8. The Scripps Research Institute, 2010 (2008).
[56]. W.L. DeLano. Pymol: An open-source molecular graphics tool. CCP4 Newsl. Protein
Crystallogr, 40(1), (2002) 82-92.
tn
[57]. E. Nittinger, T. Inhester, S. Bietz, A. Meyder, K.T. Schomburg, G. Lange, Klein R. and Rarey
M. (2017). J. Med. Chem., 60, 4245–4257.
[58]. R. Dias, L.F.S. Macedo Timmers, R.A. Caceres, J. de Azevedo, W. Filgueira. Evaluation of
rin
molecular docking using polynomial empirical scoring functions. Curr. Drug Targets, 9(12),
(2008) 1062-1070.
[59]. B. Abhipsa, P. S. Gourav, P. Falguni, P. Biswaranjan. Modification of the Time of Incubation
in Colorimetric Method for Accurate Determination of the Total Antioxidants Capacity Using
ep
gabonensis and Irvingia wombolu). Adv. J. Food Sci. Tech., 17(5), (2019) 94–98.
https://doi.org/10.19026/ajfst.17.6034.
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
[61]. P. Prieto, M. Pineda, M. Aguilar. Spectrophotometric Quantitation of Antioxidant Capacity
through the Formation of a Phosphomolybdenum Complex: Specific Application to the
ed
Determination of Vitamin E 1. Anal. Biochem. (1999) 341, 337–341.
[62]. C. Valgas, S.M. De Souza, E.F.A. Smânia, A.S. Jr. Screening methods to determine
antibacterial activity of natural products, Braz. J. Microbiol. 38 (2007) 369–380.
iew
[63]. S. Magaldi, S. Mata-Essayag, C. Hartung de Capriles, C. Perez, M.T. Colella, C. Olaizola, Y.
Ontiveros. Well diffusion for antifungal susceptibility testing, Int. J. Infect. Dis. 8 (2004) 39–45.
[64]. C. Ngwang, F. Majoumo-Mbe, E.N. Nfor, M. Akongwi, H.O. Edet, E.A. Afu, ... & H. Louis.
Theoretical modelling of the structure, reactivity, and the application of Co (II), Cu (II), and Ni
(II) Schiff base complexes as sensor materials for phosgene (COCl2) gas. Chem. Phys. Impact, 7,
v
(2023) 100352.
re
[65]. A. Fatima, H. Arora, P. Bhattacharya, N. Siddiqui, K.M. Abualnaja, P. Garg, S. Javed. DFT,
molecular docking, molecular dynamics simulation, MMGBSA calculation and Hirshfeld surface
analysis of 5-sulfosalicylic acid. J. Mol. Struct., 1273, (2023) 134242.
[66]. P.R. Spackman, M.J. Turner, J.J. McKinnon, S.K. Wolff, D.J. Grimwood, D. Jayatilaka, M.A.
er
Spackman. CrystalExplorer: A program for Hirshfeld surface analysis, visualization and
quantitative analysis of molecular crystals. J. Appl. Crystallogr, 54(3), (2021) 1006-1011.
pe
[67]. J.J. Piña, D.M. Gil, H. Pérez. Revealing new non-covalent interactions in polymorphs and
hydrates of Acyclovir: Hirshfeld surface analysis, NCI plots and energetic calculations. Comput.
Theor. Chem, 1197, (2021) 113133.
[68]. T.C. Jeyakumar, R.A. Timothy, O.C. Godfrey, R. Rajaram, E.C. Agwamba, O.E. Offiong, H.
ot
Louis. Insight into the crystal structure analysis, vibrational studies, reactivities (MESP, HOMO-
LUMO, NBO), and the anticancer activities of ruthenium diazide [Ru (POP)(PPh3)(N3)2]
complex by molecular docking approach. Inorg. Chem. Commun, 158, (2023) 111714.
tn
[69]. C.R. Baiz, B. Błasiak, J. Bredenbeck, M. Cho, J.H. Choi, S.A. Corcelli, ... M.T. Zanni.
Vibrational spectroscopic map, vibrational spectroscopy, and intermolecular interaction. Chem.
Rev., 120(15), (2020) 7152-7218.
rin
[70]. M. Ahmed, W. Lu. Probing complex chemical processes at the molecular level with
vibrational spectroscopy and X-ray tools. J. Phys. Chem. Lett, 14(41), (2023) 9265-9278.
[71]. A.V. Vlasov, N.L. Maliar, S.V. Bazhenov, E.I Nikelshparg, N.A. Brazhe, A.D. Vlasova, ... &
V.I. Gordeliy. Raman scattering: From structural biology to medical applications. Crystals, 10(1),
ep
(2020) 38.
[72]. V.K. Ahluwalia. Infrared Spectroscopy. In Instrumental Methods of Chemical Analysis (pp.
179-231). (2023) Cham: Springer Nature Switzerland.
Pr
[73]. A.B.D. Nandiyanto, R. Oktiani, R. Ragadhita. How to read and interpret FTIR spectroscope
of organic material. Indones. J. Sci. Tech, 4(1), (2019) 97-118.
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
[74]. I. Fleming, D. Williams. Infrared and Raman Spectra. Spectroscopic methods in organic
chemistry, (2019) 85-121.
ed
[75]. B. Fang, T. Wang, X. Chen, T. Jin, R. Zhang, W. Zhuang, W. Modeling vibrational spectra of
ester carbonyl stretch in water and DMSO based on molecular dynamics simulation. J. Phys.
Chem. B, 119(38), (2015) 12390-12396.
iew
[76]. D.S. Volkov, O.B. Rogova, M.A. Proskurnin. Temperature dependences of IR spectra of
humic substances of brown coal. Agron. J., 11(9), (2021) 1822.
[77]. W. Emori, H. Louis, S.A. Adalikwu, R.A. Timothy, C.R. Cheng, T.E. Gber, ...A.S. Adeyinka.
Molecular modeling of the spectroscopic, structural, and bioactive potential of
tetrahydropalmatine: insight from experimental and theoretical approach. Polycycl. Aromat.
v
Comp, 43(7), (2023) 5958-5975.
re
[78]. M. Awais, I. Zeba, S.S.A. Gillani, M. Shakil, M. Rizwan. First-principles calculations to
investigate band gap of cubic BaThO3 with systematic isotropic external static pressure and its
impact on structural, elastic, mechanical, anisotropic, electronic and optical properties. J. Phys
Chem. Solids, 169, (2022) 110878.
[79].
er
H. Louis, E.A. Eno, R.A. Timothy, E.C. Agwamba, T.O. Unimuke, P.T. Bukie, ... O.E.
Offiong. Understanding the influence of alkyl-chains and hetero-atom (C, S, O) doped electron-
pe
acceptor fullerene-free benzothiazole for application in organic solar cell: first principle
perception. Opt. Quant. Electron, 54(11), (2022) 681.
[80]. S. Sarala, S.K. Geetha, S. Muthu, A. Irfan. Computational investigation, comparative
approaches, molecular structural, vibrational spectral, non-covalent interaction (NCI), and
ot
electron excitations analysis of benzodiazepine derivatives. J. Mol. Model, 27(9), (2021) 266.
[81]. M.C. Storer, C.A. Hunter. The surface site interaction point approach to non-covalent
interactions. Chem. Soc. Rev. (2022).
tn
[82]. R.A. Timothy, H. Louis, E.A. Adindu, T.E. Gber, E.C. Agwamba, O.E. Offiong, A.M.
Pembere. Elucidation of collagen amino acid interactions with metals (B, Ni) encapsulated
graphene/PEDOT material: Insight from DFT calculations and MD simulation. J. Mol. Liq, 390,
rin
(2023) 122950.
[83]. I.A. Guedes, C.S. de Magalhães, L.E. Dardenne. Receptor–ligand molecular docking.
Biophys. Rev-Ger, 6, (2014) 75-87.
[84]. X. Totta, A.G. Hatzidimitriou, A.N. Papadopoulos, G. Psomas. Nickel(II)–naproxen mixed-
ep
ligand complexes: synthesis, structure, antioxidant activity and interaction with albumins and
calf-thymus DNA. New J. Chem. (2017) https://doi.org/10.1039/c7nj00257b
[85]. M.A. Shaheen, S. Feng; M. Anthony, M.N. Tahir, M. Hassan M, S.M. Seo, S. Ahmad, M.
Pr
Iqbal, M. Saleem, C. Lu. Metal-Based Scaffolds of Schiff Bases Derived from Naproxen:
Synthesis, Antibacterial Activities, and Molecular Docking Studies. Molecules, 24, (2019) 1237;
https://doi.org/10.3390/molecules24071237
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
[86]. D. Kovala-Demertzi, D. Hadjipavlou-Litina,, A. Primikiri, M. Staninska, C. Kotoglou, M.A.
Demertzis. Anti-inflammatory, antiproliferative, and radical-scavenging activities of tolfenamic
ed
acid and its metal complexes. Chem. Biodiversity. 6: (2009) 948-960.
[87]. L. Tabrizi, H. Chiniforoshan, P. McArdle, M. Ebrahimi, T. Khayamian. A novel bioactive
Cd(II) polymeric complex with mefenamic acid: Synthesis, crystal structure and biological
iew
evaluations. Inorg. Chim. Acta, 432, (2015) 176–184. https://doi.org/10.1016/j.ica.2015.04.010
[88]. K. Nakamoto, Infrared and Raman spectra of inorganic and coordination compounds, Part B:
Applications in Coordination, Organometallic, and Bioinorganic Chemistry, Wiley, New Jersey,
6th edn, 2009.
[89]. G.G. Nnabuike, S. Salunke-Gawali, S. Archana, A.S. Patil, R.J. Butcher, J.A. Obaleye, H.
v
Ashtekar B. Prakash. Copper (II) complexes containing derivative of aminobenzoic acid and
re
nitrogen-rich ligands: Synthesis, characterization and cytotoxic potential. J. Mol. Struct. 1279,
(2023) 135002 https://doi.org/10.1016/j.molstruc.2023.135002.
[90]. K. Kafarska, D. Czakis-Sulikowska W.M. Wolf. NOVEL Co(II) AND Cd(II) COMPLEXES
WITH NON-STEROIDAL ANTI-INFLAMMATORY DRUGS Synthesis, properties and thermal
er
investigation. J. Therm. Anal. Calorim, Vol. 96 (2009) 2, 617–621
[91]. R. Smolkova, L. Smolko, V. Zelenak, J. Kuchar, R. Gyepes, I. Talian, J. Sabo, Z. Biscakova,
pe
M. Rabajdov. Impact of the central atom on human genomic DNA and human serum albumin
binding properties in analogous Zn(II) and Cd(II) complexes with mefenamic acid J. Mol. Struct.
1188 (2019) 42e50 https://doi.org/10.1016/j.molstruc.2019.03.078
[92]. A. Tarushi, P. Kastanias, C.P. Raptopoulou, V. Psycharis, D.P. Kessissoglou, A.N.
ot
[93]. Y.C. Chu, T.T. Wang, L.J. Wang, Q.Y. Luo, R. Jia, T.C. Hong, X.M. Wang, H.L. Zhu.
(2019). Synthesis, characterization, and biological evaluation of a novel Zn(II)-Naproxen
complex. Polyhedron, 163 (2019) 71–76
rin
[94]. J. Sharma, A.K. Singla, S. Dhawan. Zinc–naproxen complex: synthesis, physicochemical and
biological evaluation. Int. J. Pharmaceut., 260 (2003) 217–227.
[95]. C.N. Banti, C. Papatriantafyllopoulou, A.J. Tasiopoulos, S.K. Hadjikakou. New metalo-
therapeutics of NSAIDs against human breast cancer cells. Eur. J. Med. Chem, (2017)
ep
https://doi.org/10.1016/j.ejmech.2017.10.067
[96]. G.G. Nnabuike, S. Salunke-gawali, A.S. Patil, R.J. Butcher, J.A. Obaleye. Synthesis and
structures of tetrahedral zinc ( II ) complexes bearing indomethacin and nitrogen donor ligands.
Pr
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275
inflammatory drug naproxen. J. Inorg. Biochem, 107(1), (2012) 54–64.
https://doi.org/10.1016/j.jinorgbio.2011.10.014
ed
[98]. L. Tabrizi, H. Chiniforoshan, P. McArdle, P. Synthesis, crystal structure and spectroscopy of
bioactive Cd(II) polymeric complex of the non-steroidal anti-inflammatory drug diclofenac
sodium: Antiproliferative and biological activity. Spectrochim. Acta A, 136(PB), (2015) 429–436.
iew
https://doi.org/10.1016/j.saa.2014.09.053
[99]. L. Tabrizi, R. Golbang, H. Sadeghi, H. Chiniforoshan, P. Mcardle, B. Notash. Dinuclear
cadmium indomethacin and Lawsone complexes: synthesis, crystal structures, antiproliferative
and biological evaluations. J. Coord. Chem, 69(20), (2016) 3021–3034.
https://doi.org/10.1080/00958972.2016.1223845
v
[100]. M. Lazou, A.G. Hatzidimitriou, A.N. Papadopoulos, G. Psomas. Transition metal(II)
re
complexes with the non–steroidal anti–inflammatory drug oxaprozin: Characterization and
biological profile. J. Inorg. Biochem. Volume 243, June 2023, (2023) 112196
https://doi.org/10.1016/j.jinorgbio.2023.112196
[101]. D. Ahmed, M.M. Khan, R. Saeed. Comparative analysis of phenolics, flavonoids, and
er
antioxidant and antibacterial potential of methanolic, hexanic and aqueous extracts from
Adiantum caudatum leaves. Antioxidants-Basel, 4(2), (2015) 394–409.
pe
https://doi.org/10.3390/antiox4020394
[102]. M.S. Hasan, N. Das. A detailed in vitro study of naproxen metal complexes in quest of new
therapeutic possibilities. Alex. J. Med, 53(2), (2017) 157–165.
https://doi.org/10.1016/j.ajme.2016.06.003
ot
This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4937275