You are on page 1of 16

Journal of Molecular Structure 1276 (2023) 134759

Contents lists available at ScienceDirect

Journal of Molecular Structure


journal homepage: www.elsevier.com/locate/molstr

Structural and theoretical insights towards the understanding of the


effect on the conformation of ligand by complexation process
B.S. Chethan a, H.R. Rajegowda b, Riyaz ur Rahaman Khan b, N.K. Lokanath a,∗
a
Department of Studies in Physics, University of Mysore, Manasagangotri, Mysuru, Karnataka 570006, India
b
Department of Chemistry, Acharya Institute of Graduate Studies, Soladevanahalli, Bengaluru, Karnataka 560107, India

a r t i c l e i n f o a b s t r a c t

Article history: The ligand and the palladium complex were synthesized and characterized by the single crystal X-ray
Received 10 October 2022 diffraction method. The Hirshfeld surface analysis quantified the intermolecular interactions stabilizing
Revised 1 December 2022
the crystal packing in both compounds. DFT (density functional theory) calculations were used to opti-
Accepted 5 December 2022
mize the geometric structures and calculate the electronic properties of the synthesized compounds. The
Available online 5 December 2022
ligand and complex’s electronic structure and UV-Vis spectrum are interpreted by DFT and TD-DFT (time
Keywords: dependent-density functional theory) calculations. The NBO (natural bond orbital) analysis was employed
Palladium complex to study the origin and nature of the hyperconjugation interactions. The intramolecular interaction ener-
DFT gies associated with different topologies were calculated using the QTAIM (Quantum Theory of Atoms in
Hirshfeld surface Molecules), and the non-covalent interactions revealed the type and nature of the interactions exhibited
QTAIM by the compound. The structural and theoretical analysis revealed the change in the properties of the
Non-covalent interactions
ligand due to the complexation process.
© 2022 Elsevier B.V. All rights reserved.

1. Introduction The metal complexes with positively charged metal centers


favor binding to the negatively charged biomolecules (the con-
Synthesis of the Schiff based ligand and its corresponding metal stituents of the proteins and nucleic acids) [12]. Hence, the phar-
complexes is one of the most intensively developing fields of in- maceutical use of such metal complexes has excellent poten-
organic chemistry. The various properties exhibited by these com- tial. Due to the promising biological activity, such as anti-cancer
pounds, which are useful for various applications, stimulate further drugs, antimicrobial, antimycotic bacterial agents, and anti-HIV
research in the domain [1–5]. activities [13], the metal complexes are of great importance in
The Schiff base compounds exhibit significant biological activi- the field of the drug. Hence, an extensive understanding of the
ties, which play a key role in medicine and pharmaceutical fields crystal structure is required for the purposeful synthesis of the
[6–8]. The Schiff bases are highly sorted as ligands in transition new drugs [14]. Many metal complexes with chelating ligands are
metal chemistry due to their ability to stabilize the metals in found to be promising candidates for the synthesis of single-phase,
different oxidation states. The presence of the azomethine group stoichiometric-mixed, and hetero-metal-doped metal oxides [15],
(CH=N) in the ligand is mainly responsible for the high biologi- metal oxide nanoparticles [16], metal sulfide thin films [17] and
cal activity of the Schiff base ligands, as the nitrogen atom of the metal sulfide nanoparticles through thermal decomposition path-
azomethine group plays an important role in the formation of the ways [18].
hydrogen bonds with the active centers of the biological molecules Theoretical investigations are important to explore the funda-
[9,10]. In the formation of the metal complexes from the Schiff mental properties and to understand the physical and chemical
based ligands, the consideration of the role of the non-covalent in- properties of the ligands and metal complexes. Various compu-
teractions is essential as it plays a major role in the formation of tational chemistry tools, such as density functional theory (DFT),
the solid phase architecture, and also, the theoretical analysis pro- have proven vital in biological and chemical research. The under-
vides greater insights towards the understanding of the practical standing of compounds is further enhanced by the experimental
application of the formation of various crystalline motifs [11]. and theoretical investigations of the structure, spectroscopic, and
Hirshfeld surface analysis.
In this regard, it is interesting to consider the cases where the

Corresponding author. azomethine group acts as an important factor in creating non-
E-mail address: lokanath@physics.uni-mysore.ac.in (N.K. Lokanath). covalently bound supramolecular architectures. Thus, in this as-

https://doi.org/10.1016/j.molstruc.2022.134759
0022-2860/© 2022 Elsevier B.V. All rights reserved.
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

pect, we report the synthesis and in-depth studies of the lig- 100 mL of distilled water. The complex was extracted into chlo-
and synthesized from the condensation reaction between ethyl roform (100 mL × 2) from the reaction mixture. The organic layer
2-amino-4,5,6,7-tetrahydrobenzo[b]thiophene-3-carboxylate and 2- that separated was dried over anhydrous sodium sulfate. This chlo-
hydroxy-1-napthaldehyde and the corresponding palladium com- roform extract was concentrated to 10 mL on a rotary evaporator
plex through the single-crystal X-ray crystallography and various and then treated with n-hexane (15 mL). The resulting precipitate
computational methods. The main objective of this work is to of [PdLCl] was filtered, washed with n-hexane (2 × 10 mL), and
study the structural property changes during the complexation dried. The orange-red single crystals of [PdLCl] suitable for X-ray
process with the synthesized Schiff-base ligand. diffraction study were grown by slow evaporation of a mixture of
chloroform and n-hexane solution of [PdLCl] (0.23 g, 88%); (Melt-
2. Experiments ing point: 253 °C). 1 H–NMR (DMSO-d6 , 399.82 MHz) δ : 0.824–
0.859 (q, 3H, –CH3 ), 1.415 (bs, 2H, 16–H2 ), 2.149 (s, 2H, 17–H2 ),
2.1. Materials 2.329–2.386 (d, 2H, 18–H2 ), 2.983 (s, 2H, 15–H2 ), 3.814–3.849 (q,
2H, –OCH2 –), 6.862–6.885 (d, 1H, ArH, 3–H), 6.961–6.996 (d, 1H,
2-Hydroxy-1-naphthaldehyde, ethyl 2-amino-4,5,6,7-tetra- ArH, 6–H), 7.117–7.136 (d, 1H, ArH, 7–H), 7.444–7.464 (d, 1H, ArH,
hydrobenzo[b]thiophene-3-carboxylate and sodium tetrachloropal- 5–H), 7.594–7.618 (d, 1H, ArH, 8–H), 7.714–7.735 (d, 1H, ArH, 4–
ladate (Na2 PdCl4 ) were purchased from Sigma Aldrich Ind. Ltd, H), 8.412 (s, 1H, CH=N); 13 C{1 H}-NMR (DMSO-d6 , 100.54 MHz) δ :
and they were used as received. The NMR spectra were recorded 14.282 (–CH3 ), 22.205 (C–16), 22.6 (C–17), 24.2 (C–18), 26.5 (C–
using the Agilent NMR spectrometer at 400 MHz for 1 H NMR 15), 60.065 (–OCH2 –), 110.295 (C–1), 122.680 (C–8, C–3), 123.3 (C–
and 100 MHz for 13 C NMR. The signals are abbreviated: s, singlet; 6), 123.7 (C–12), 126.846 (C–7), 128.933 (C-5, C-13), 130.178 (C-10,
d, doublet; t, triplet; dd, doublet-doublet; m, multiplet and the C-4), 133.259 (C-9), 133.813 (C-14), 157.285 (C-11), 160.237 (C=O),
chemical shifts (δ ) are reported in ppm (parts per million). FTIR 162.195 (CH=N), 165.337 (C-2).
characterization was carried out on a PerkinElmer 100 FTIR spec-
trometer. UV-visible spectra were recorded using a PerkinElmer 2.3. Structural analysis
Lambda 35 spectrometer with 20 μg mol−1 concentration so-
lutions for all the compounds. Single crystal X-ray diffraction A suitable single crystal of the ligand and the palladium com-
intensity measurement using Mo-Kα radiation (λ = 0.71073 Å) at plex was carefully selected under a polarizing microscope. It was
room temperature. mounted at the tip of the goniometer head using adhesive. The
single crystal intensity data of the ligand (LH) and palladium com-
plex ([PdLCl]) were collected using Rigaku Xta-lab mini diffrac-
2.2. Synthesis tometer equipped with sealed-tube X-ray source (Mo-Kα radia-
tion = 0.71073 Å) operating at 50 kV and 12 mA. The data were
2.2.1. Preparation of the Schiff-base ethyl (E)−2- collected for different settings of φ (0°−90°), with a scan width of
(((2-hydroxynaphthalen-1-yl)methylene)amino)−4,5,6,7- 0.5° and an exposure time of 3 s, by keeping the sample to a de-
tetrahydrobenzo[b]thiophene-3-carboxylate (LH) tector distance of 50 mm. The complete data were processed using
The solution of ethyl 2-amino-4,5,6,7-tetrahydrobenzo[b]- the d∗ trek program [19], and the absorption correction was per-
thiophene-3-carboxylate (0.225 g, 1 mmol) in ethanol (15 mL) was formed using the multi-scan employed in CrystalClear-SM Expert
added drop-wise to a magnetically stirred solution of 2-hydroxy- [20]. Using the OLEX2 software [21], the crystal structure of the
1-naphthaldehyde (0.172 g, 1 mmol) in ethanol (15 mL) at room compounds was solved by direct methods (SHELXS) [22], and the
temperature. The resulting orange-red mixture was refluxed with position of all the non-hydrogen atoms was identified and refined
stirring for 3 h, during which the progress of the reaction was on F2 by a full-matrix least-squares procedure using anisotropic
monitored by TLC. After completion of the reaction, the solvent displacement parameters as provided in the SHELXL program pack-
was distilled under reduced pressure on a rotary evaporator. The age [23]. The non-hydrogen atoms are refined anisotropically, and
solid obtained is then dried at room temperature in a desicca- the hydrogen atoms were treated as horse-riding atoms using the
tor over anhydrous calcium chloride to afford ligand (LH) as an SHELXL default parameters. The geometrical calculations and the
orange-red crystalline solid. This solid was recrystallized in a 1:1 crystal packing diagrams were prepared by the crystallographic
mixture of dichloromethane and n-hexane, giving needle shape program PLATON and MERCURY 4.2.0 software, respectively [24].
crystals of ligand (0.36 g, 95%); (Melting point: 170 °C). 1 H–NMR
(CDCl3 , 399.82 MHz) δ : 0.628–0.663 (t, 3H, –CH3 ), 1.02–1.039 2.4. Hirshfeld surface analysis
(t, 2H, 16–H2 ), 1.052–1.082 (t, 2H, 17–H2 ), 1.939–1.967 (t, 2H,
18–H2 ), 2.016–2.045 (t, 2H, 15–H2 ), 3.635–3.688 (q, 2H, –OCH2 –), The Hirshfeld surface analysis and its associated two-
6.390–6.413 (d, 1H, ArH, 3–H), 6.577–6.595 (d, 1H, ArH, 6–H), dimensional fingerprint plots for both the ligand and the
6.739–6.760 (d, 1H, ArH, 7–H), 6.948–6.968 (d, 1H, ArH, 5–H), corresponding palladium complexes were generated using the
7.021–7.044 (d, 1H, ArH, 8–H), 7.314–7.336 (d, 1H, ArH, 4–H), CrystalExplorer 17.5 software [25] with the help of the inalter-
8.514 (s, 1H, CH=N); 13 C{1 H}-NMR (CDCl3 , 100.54 MHz) δ : 14.7 able refined crystallographic information file (CIF) as input. The
(–CH3 ), 25.89 (C–16), 26.67 (C–17), 27.26 (C–18), 31.51 (C–15), Hirshfeld surfaces were mapped over dnorm , shape index, and
61.13 (–OCH2 –), 109.94 (C–1), 119.62 (C–8), 119.85 (C–3), 120.69 curvedness. The dnorm (normalized contact distance) is described
(C–6), 123.59 (C–12), 123.89 (C–7), 128.09 (C-5), 129.72 (C-13), in terms of the van der Waals radii (vdW) of atoms, di (distance
131.16 (C-10), 132.97 (C-4), 136.09 (C-9), 136.25 (C-14), 153.38 between the Hirshfeld surface and the closed nucleus inside
(C-11), 154.01 (C=O), 163.86 (CH=N), 165.62 (C-2). the surface), and de (distance from the Hirshfeld surface to the
nearby nucleus outside the surface) parameters are represented
2.2.2. Preparation of the palladium complex [PdLCl] by the color gradient from red to blue. The dark red spots on the
To a stirred solution of Na2 PdCl4 (0.147 g, 0.5 mmol) in wa- dnorm surface indicate the short interatomic contacts, whereas the
ter (10 mL) added, a solution of LH (0.190 g, 0.5 mmol) pre- other intermolecular interactions appear as light-red spots. The
pared in acetone (20 mL) under vigorous stirring. The resulting blue color indicates the presence of non-interaction regions. The
mixture was stirred for 3 h at room temperature. After complete various intermolecular interactions exhibited by both compounds
consumption, the ligand was monitored by TLC and poured into can be understood by the dnorm surface, and the quantifying of

2
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 1. Scheme of the synthesized ligand and its palladium complex.

these interactions was done using the 2D fingerprint plots [26,27]. 3. Results and discussion
Through the Hirshfeld surface analysis, all the close contacts of
the ligands and the palladium complex can be visualized, provid- 3.1. Synthesis of ligand (LH) and its Pd(II) complex [PdLCl]
ing valuable insights toward a better understanding the crystal
packing behavior in the solid state. The Schiff base ligand and its complex [PdLCl] were synthe-
sized, as shown in Fig. 1. The LH is freely soluble in chloroform,
2.5. Computational studies dichloromethane, methanol, ethanol, acetone, acetonitrile, DMSO,
and DMF but insoluble in n-hexane, diethyl ether, THF, toluene, and
Quantum computational techniques are highly useful for theo- water. The orange-red solid of [PdLCl] was synthesized by a 1:1
retically investigating the compounds’ physical and chemical prop- stoichiometric reaction of the ligand with Na2 PdCl4 at room tem-
erties. The geometry optimization was accomplished within the perature. The complex [PdLCl] showed solubility in polar solvents
framework of the density functional theory (DFT) [28], using the such as chloroform, dichloromethane, acetone, acetonitrile, DMSO,
hybrid functional with non-local exchange due to Becke and the and DMF. The complex is stable to air and moisture in both solid
correlation functional due to Lee, Yang, and Parr (B3LYP) [29], as and solution state. The synthesis scheme of the ligand and the pal-
implemented in the Gaussian 16 package [30]. The calculations ladium complex is given in Fig. 1.
were performed with the mixed basis set 6-311+G(d,p) [31] for
non-metal atoms and LanL2DZ [32] for the Pd(II) metal center 3.2. 1 H and 13 C{1 H} NMR spectral data
without any symmetry constraint. The atomic coordinates of the
crystal structure were used as the initial guess for electronic The 1 H and 13 C{1 H} NMR spectra of ligand (LH) were recorded
structure optimization. The corresponding vibrational analyses per- in CDCl3 and the spectra are shown in Figs. S1 and S2 (supplemen-
formed for the optimized geometries revealed the achievement of tary file), respectively. In the 1 H NMR spectrum of LH, the signal
the local minima on the potential energy surface. for methylene protons (–OCH2 ) appeared as a quartet at δ , 3.64–
Using Koopmans’ approximation, the highest occupied and low- 3.69 ppm, and the methyl protons signal observed at an upfield
est unoccupied molecular orbitals (HOMO-LUMO), their energy region of δ , 0.63–0.66 ppm as a triplet and the other four methy-
gap, and the related global and local indices revealing the chemical lene protons gave peaks at δ , 1.02–1.04 (16–H2 ), 1.05–1.08 (17–H2 ),
properties of the molecules were calculated [33]. To investigate the 1.94–1.97 (18–H2 ), 2.02–2.04 (15–H2 ) ppm as a triplet. The azome-
charge distribution and the possible reactive sites, molecular elec- thine (CH=N) proton signal appeared as a singlet at downfield re-
trostatic potential (MEP) were carried out. Furthermore, the natu- gion δ , 8.51 ppm as reported earlier in the similar Schiff base lig-
ral bond orbital (NBO) calculations were performed to understand ands [38]. In 13 C{1 H} NMR spectra of LH, the methylene carbon
various hyperconjugative interactions that stabilize the molecular (–OCH2 ) gave a peak at δ , 61.1 ppm, and the other four methy-
structures. All the results for the DFT calculations were visualized lene carbons gave peaks at δ , 25.9 (C–16), 26.7 (C–17), 27.3 (C–18),
using the GaussView 5 software [34]. 31.5 (C–15) ppm. The carbon signal for –CH3 appears at the upfield
The obtained geometries were used at the same level of the- region of δ , 14.7 ppm, but the -C=N (imine) and –C=O (ester) car-
ory for the calculations for QTAIM [35], NCI [36], reactivity descrip- bon signals observed at the downfield field chemical shift region
tors, and aromaticity. The QTAIM topological properties of the elec- of δ , 163.9 and 154 ppm, which are in agreement with the chem-
tron densities at the bond critical points (BCPs) were computed. ical shifts of those carbons reported for related Schiff base ligands
Furthermore, the RDG (reduced density gradient) based NCI anal- [39]. The 1 H and 13 C{1 H} NMR spectra of [PdLCl] were recorded in
ysis was carried out, and the corresponding color-filled isosurfaces DMSO–d6 and are shown in Figs. S3 and S4, respectively. In the 1 H
graphs were drawn using the VMD (Visual Molecular Dynamics) NMR spectrum of [PdLCl], the signal for–OCH2 appeared as a quar-
visualizer [37]. tet at δ , 3.81–3.85 ppm, and the methyl protons signal observed at

3
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 2. ORTEP plot of the synthesized ligand.

Table 1
Analysis of intramolecular hydrogen interactions exhibited by the ligand.

Donor-
H···Acceptor D···H (Å) H···A (Å) D···A (°) D-H···A (°)

O1-H1···O2 0.82 2.55 3.174(3) 134.17


O1-H1···N1 0.82 1.84 2.57(3) 146.78
C11-H11···S1 0.93 2.64 3.065(3) 108.36

δ , 0.82–0.86 ppm as a triplet. The azomethine (CH=N) proton sig-


nal appeared as a singlet at downfield region δ , 8.41 ppm [39,40].
In the 1 H and 13 C NMR spectra of LH, the methylene carbon (–
OCH2 ) peaked at δ , 60.06 ppm, and the carbon signal for –CH3 ap-
peared at the upfield region of δ , 14.3 ppm. The -C=N (imine) and
ester (–C=O) carbon signals were observed at the downfield field
chemical shift region of δ , 162.2, and 160.2 ppm [39,40]. The other
proton and carbon signals in LH and [PdLCl] appeared at their char-
Fig. 3. The strong intramolecular interactions lead to the formation of the various
acteristic δ ppm.
synthons (S(6) and S(5)).

3.3. Crystal structure analysis


atoms resulting in the formation of a 1D molecular sheet and
The X-ray structure determination details of the ligand and its supramolecular ring motif of R2 2 (10) as shown in Fig. 4. Further,
palladium complex are given in Table S1 (supplementary file). due to the weak intermolecular interactions of the C11 atom with
the five-membered ring (C12/S1/C13/C17/C18) and six-membered
3.3.1. Crystal structure of the ligand ring (C7/C8/C9/C10/C5/C6) results in the formation of 2D architec-
The ORTEP plot of the Schiff base ligand formed by the ture along the crystallographic b-axis, which further stabilizes the
condensation reaction between the ethyl 2-amino-4,5,6,7- crystal structure (Fig. S6). The tabulation of the π ···π interactions
tetrahydrobenzo[b]thiophene-3-carboxylate and 2-hydroxy-1- exhibited by the ligand is given in Table 2. The torsional angle of
naphthaldehyde is given in Fig. 2. The Schiff base ligand is planar C12-N1-C11-C10 is found to be −174.71(2)°, indicating an anti-
with a dihedral angle of 15.06°, as shown in Fig. S5. The C11=N1 periplanar arrangement. The conformational assignments across
bond lengths are consistent with a double bond (1.292 Å), while the bonds are given in Table S2. The hydrogen bonds were found
the C9-O1 bond length (1.335 Å) is consistent with that of a single to participate in the short inter/intramolecular interactions, and
bond. further, the van der Waals (π ···π ) interactions also contribute to
The hydroxy group of the 2-hydroxy-1-naphthaldehyde the crystal packing (Tables 1 and 2).
interacts with the nitrogen of the ethyl 2-amino-4,5,6,7-
tetrahydrobenzo[b]thiophene-3-carboxylate through strong intra 3.3.2. Crystal structure of the [PdLCl] complex
O1-H1···N1 interactions resulting in the formation of the S(6) The synthesized Pd complex crystallizes in the triclinic space
synthon as shown in Fig. 3. Further, due to the interaction be- group P1̄ with Z = 2. The ORTEP of the compound is given in
tween C11-H11···S1 and C15-H15···O3 resulted in the formation Fig. 5. The coordination sphere around the central Pd-atom consists
of synthon S(5) (Fig. 3). The various intra molecular interactions of two O- atoms and one N- and Cl- atom. The coordination dis-
exhibited by the compound responsible for forming the synthons tances for Pd1 case of Pd1-O3 = 1.996(3) Å, Pd1-O1 = 1.961(3) Å,
are given in Table 1, along with its corresponding distances and Pd1-Cl1 = 2.304(15) Å and Pd1-N1 = 2.027(3) Å was observed. The
angles. The molecules are interconnected with each other along different Pd1-O1 and Pd1-O2 distance results in the polyhedron
the crystallographic c-axis through C15-H15B···O3 and C3-H3···O3 distortion closer to a tetragonal flattening rather than the elonga-

4
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 4. The intermolecular interactions result in the formation of a 1-D molecular sheet along the c-axis.

Table 2
The geometric features (distances in Å) of the π ···π interactions obtained for the ligand.

Cg(ring I)···Cg(ring J) Cg···Cg (Å) α (°) β (°) γ (°) Cg(I) ···perp (Å) Cg(J) ···perp (Å)

Cg(1)···Cg(3)i 3.866(2) 12.07(12) 15.0 26.9 −3.446(10) −3.7344


Cg(3)···Cg(3)ii 3.997(2) – 29.4 29.4 3.482(11) 3.482(11)

Symmetry transformations: i = 2-x, -y, 2-z


ii = 3-x, -y, 2-z. Cg(1) = Center of gravity of the ring (S1/C12/C13/C7/C18); Cg(3) = Center of gravity of the ring
(C5/C6/C7/C8/C9/C10).

Table 3
Various intra-molecular and inter-molecular hydrogen bond interactions result
in the formation of synthons.

Donor-
H···Acceptor D···H (Å) H···A (Å) D···A (Å) D-H···A (°)

C11-H11···S1 0.93 2.35 2.897(4) 117.00


C22-H22A···O2 0.97 2.39 2.782(8) 103.22
C23-H23···Cl1 0.98 2.44 3.407(11) 169.38

H23···Cl1 interaction. This CHCl3 solvent connects the other two


solvent molecules by Cl2 and Cl3 atoms along the (110) and (111)
directions respectively (Fig. 6). These solvent molecules act as a
bridge interconnecting two complex molecules resulting in the for-
mation of wave-like nature along the a-axis (Fig. S8). Further, the
Cg···Cg interaction results in the generation of another molecule
along the bc-plane resulting in the formation of a 2-D structure
which stabilizes the crystal structure (Fig. S9). Due to the rotation
of the C13-C14 bond, modification in the intramolecular interac-
tions in the complex was observed with the formation of two ad-
ditional S(5) synthons compared to that of the ligand, as shown in
Fig. 7. Various intramolecular interactions and the π ···π interac-
tions contributing towards the stabilization of the crystal structure
are given in Tables 3 and 4, respectively.
Fig. 5. ORTEP of the palladium complex drawn at 50% probability.

3.4. Fourier transform infrared spectra analysis

tion of the octahedra. Hence, the palladium coordination polyhe- The experimental IR spectra of the Pd(II) complex exhibited the
dron adopts the square planar structure [41]. characteristic absorption bands around 1542–1569 cm−1, which is
The internal angles in the coordination of Pd complex attributed to the v(C=N) and the absorption band observed in the
are N1-Pd1-O3 = 93.00(13)°, N1-Pd1-O1 = 92.79(13)°, O3-Pd1- region 1546.64 cm−1 corresponds to the C=N vibration. A shift was
Cl1 = 87.18(10)° and O1-Pd1-Cl1 = 87.31(11)°. The dihedral angle observed in the absorption for the metal complex 1589.32 cm−1 ,
formed by the mean planes through the two six-membered rings indicating the coordination of the Palladium to the imine nitro-
is 1.85° (Fig. S7). gen [42]. The C=N stretching band in the Schiff base ligand was
Interestingly, it is worth noting that compared to the ligand in observed at 1691.92 cm−1 . In comparison, this peak was shifted
its uncoordinated state, a rotation about the C13-C14 bond by 180° to a lower frequency for the palladium complex (1589.32 cm−1 ),
was observed to allow the oxygen (O3) atom to chelate with the indicating the coordination of the nitrogen atom to the central
palladium metal. Interestingly the crystallization solvent (CHCl3 ) Pd(II) atom [43,44]. The stretching of the phenolic OH group was
also crystallizes bound with the palladium metal complex by C23- observed at 3666.75 cm−1 . This band is absent in the spectra of

5
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 6. Intermolecular interactions of the complex via solvent molecule along (110) and (111) directions.

Fig. 7. Intramolecular interactions of the complex (the solvent molecule is omitted for clarity).

Table 4
The Cg···Cg interactions contribute towards the stability of the crystal structure.

Cg(I)···Cg(J) Cg···Cg (Å) α (°) β (°) γ (°) Cg(I)···perp (Å) Cg(J)···perp (Å)

Cg(1)···Cg(3) 3.972(3) 4.5(2) 31.0 31.5 3.386(2) 3.4050


Cg(1)···Cg(4) 3.654(3) 3.8(3) 22.0 20.2 −3.428(2) −3.387
Cg(1)···Cg(5) 3.777(3) 3.3(3) 21.6 24.8 −3.427(2) −3.512
Cg(2)···Cg(1) 3.927(3) 4.5(2) 31.5 31.0 3.4052(16) 3.3806
Cg(2)···Cg(3) 3.674(2) 1.85(16) 20.0 21.4 3.4207(16) −3.437
Cg(4)···Cg(3) 3.594(3) 3.1(3) 17.1 19.7 −3.384(3) −3.5182

the palladium complex, indicating the deprotonation of the pheno- ory) approach with the CPCM (Conductor-like Polarizable Contin-
lic OH followed by the coordination to the palladium ion. Further, uum Model) to simulate the presence of the solvent. The wave-
these results were found to be in collaboration with that the the- length (λ), oscillator strength (f), excitation energies (E), and as-
oretical FTIR results. The theoretical and experimental FTIR spectra signment for the electronic transitions are given in Table 5. The
of the ligand and the palladium complex are given in Figs. 8 and UV–Visible spectra of the ligand showed a major absorption peak
S10, respectively. at 431.50 nm (theoretical: 455.16 nm), which is mainly due to
the π → π ∗ transitions. Further, the theoretical spectra showed a
3.5. UV-Vis spectral analysis shoulder at 314.32 nm, which is attributed to the n→π ∗ transi-
tions. A shift in the absorption peak was observed for the metal
The UV-visible absorption spectra of both the ligand and the complex, and a major absorption peak at 529.91 nm (theoretical:
complex were recorded at room temperature with 20 μg ml−1 513.36 nm) can be ascribed to the π → π ∗ transitions. For the pal-
concentration solutions in DMSO for the wavelength range of 300– ladium complex, the electronic spectral results indicated that the
700 nm. The electronic spectra of the compounds were also cal- complex had diamagnetic behavior exhibiting square planar geom-
culated using TD-DFT (Time Dependent Density Functional The- etry [45,46]. The theoretical and experimental UV-absorption spec-

6
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 8. Theoretical FTIR spectra of the ligand and the complex obtained using the same level of theory.

Table 5 electron density and conformational changes in the complex com-


Assignment of the UV absorption bands of the ligand and the com-
pared to that of the ligand molecule. The 2D fingerprint plots of
plex obtained from the TD-DFT calculations.
the Hirshfeld surface of the ligand, palladium complex, and its rel-
Wavelength (λ) E (eV) Configuration ative contributions of different interactions overlapping in the full
(nm)
fingerprint plots are shown in Fig. 12. The fingerprint plots can
Ligand be decomposed to highlight particular atoms pair close contacts,
455.16 2.724 H→L (98.99%) where the complementary regions of one molecule act as a donor
353.05 3.5118 H-2→L (92.52%)
(de > di ) and the other as an acceptor (de < di ).
339.53 3.6516 H-3→L (40.35%)
Complex In both the ligand and its palladium complex, the H···H interac-
575.85 2.1531 H→L (23.28%) tions are found to have the highest contribution to the total Hir-
575.85 2.1531 H→L+1 (64.37%) shfeld surface. A decrease in the percentage of hydrogen interac-
483.80 2.5627 H→L (30.11%)
tions was observed for the complex (32.7%) compared to that of
the ligand (55.3%), this is due to the chelation of the palladium
metal with the ligand molecule. Similar results were observed for
tra of the ligand and the complex are given in Figs. 9 and S11, re- the C···H interactions (ligand: 16.3% and Pd complex: 8.0%) and
spectively. O···H interactions (ligand: 10.6% and Pd complex: 4.8%), where the
contribution of these contact types are found to be decreased, as
compared to that of the metal complex. Interestingly, the C···C in-
3.6. Hirshfeld surface analysis
teractions are found to be increased in the metal complex (7.2%)
compared to that of the ligand (6.8%) (Fig. 13).
The significant intermolecular interactions in the crystalline ar-
Due to the complexation of the palladium chloride with the
rangements of the ligands and the metal complex are evaluated
ligand and the presence of the chloroform (CHCl3 ) as a solvent
by the Hirshfeld surface (HS) analysis. The Hirshfeld surfaces of
molecule, leading to the formation of the H···Cl interaction con-
the ligand and its palladium complex are shown in Fig. 10, re-
tributing about 31.8% to the total Hirshfeld surface of the molecule.
vealing the surfaces that have been mapped over dnorm [(−0.0522
to 1.4154 Å) in a1 and (−0.1151 to 1.5663 Å) in a2], shape index
[(−1.00 to 1.00 Å) in b1 and (−1.00 to 1.00 Å) in b2] and curved- 3.7. Structure properties study based on the density functional theory
ness [(−4.00 to −0.40 Å) in c1 and (−4.00 to −0.40 Å) in c2]. A (DFT)
significant change in the position of the blue and red triangles,
which indicates the C–H···π interactions, was observed between Theoretical calculations were performed to optimize the molec-
the shape index plot of the ligand and the palladium complex, as ular structures of the ligand and the complex to evaluate the phys-
shown in Fig. 10(b1) and 10(b2) respectively. Further, the flat re- ical and chemical properties of the synthesized compounds. The
gions on the curvedness plot validate the existence of the π ···π in- absence of the imaginary frequencies of the compounds confirmed
teractions in both the ligand and the complex. These findings were the existence of optimized structures to their ground state (mini-
found to be in accordance with the structural analysis results. The mum energy).
de and di mapped Hirshfeld surface of the ligand and the palla- The primary research objective is to study the complexation
dium complex is shown in Fig. 11. Significant changes were ob- process of the ligand with Pd2+ cation and the resulting structural
served in the surface, which is attributed to the changes in the and energetic variations. The structures of both the ligand and the

7
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 9. UV-visible spectra of the ligand and the complex were obtained using TD-DFT calculations.

Fig. 10. Hirshfeld surface plot of the ligand (dnorm : (a1), shape index: (b1), curvedness: (c1)) and the palladium complex (dnorm : (a2), shape index: (b2), curvedness: (c2)).

8
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 11. The Hirshfeld surface is generated using the surface property de (ligand: d1, e1) and di (palladium complex: d2, e2).

complex were optimized at the B3LYP/6-311+G(d,p) level of theory. ness (η) value indicates the molecular reactivity and the high value
The comparison of the bond length, bond angles, and torsional an- of the ligand compared to that of the complex is in accordance
gles of the crystal structure and the optimized structures are given with the energy gap values. The chemical potential (μ) value re-
in Table S3. An excellent correlation coefficient was observed be- flects the charge transfer of the system from a higher chemical
tween the experimental and theoretical structures (Ligand: bond potential to one of the lower chemical potentials confirming the
length – 0.999, bond angle – 0.9999, torsion angle – 0.999, palla- charge transfer happening from the ligand (−4.0963 eV) to the
dium complex: bond length – 0.988, bond angle – 0.9941, torsion complex (−4.6355 eV). This is further substantiated by the higher
angle – 0.988). electronegativity (χ ) of the complex (4.6355 eV) compared to that
of the ligand (4.0963 eV), as electronegativity value refers to the
3.8. Study of the electronic properties power to attract the electrons from a donor.

To analyze and understand the chemical activity of the com-


pounds, the frontier molecular orbitals (FMO) are essential. The 3.9. Molecular electrostatic potential (MEP)
FMO of the ligand and the complex were computed using the func-
tional B3LYP and 6-311+G(d,p) basis set. The MEP studies were performed to understand the extent of
The frontier molecular orbitals and the energy diagrams of the the chemical reactivity and to explore the molecular structures’ re-
synthesized ligand and the palladium complex are given in Figs. lationship with the compounds’ physicochemical properties. With
14 and 15, respectively. The highest occupied molecular orbitals the help of the color scheme employed in the MEP diagram, the
(HOMO) have the electron donating tendency, whereas the lowest electrophilic (positive regions) and nucleophilic sites (negative re-
unoccupied molecular orbitals (LUMO) have the electron accept- gions) are identified, as represented by the blue and red colors,
ing tendency [47]. From Fig. 14, the plot of the HOMO and LUMO respectively. Identifying these reactive regions provides significant
of the ligand indicates that the orbitals are spread over the en- information regarding the charge distribution in the compounds
tire ligand molecule. In contrast, for the metal complex, the HOMO [56]. The MEP map of the ligand and the complex, along with the
and LUMO orbitals are shifted and mainly concentrated on the C–N values, are given in Fig. 16. Significant changes in the electrostatic
bond, where the electrons are partially transferred from the ni- potential between the ligand and the complex were observed. In
trogen and the oxygen of the ligand to the central metal atom the MEP plot of the ligand, the red region was found to be around
(Fig. 15) [48]. The HOMO-LUMO orbital energies, their energy gaps, the oxygen atom of the carboxylate group, signifying it to be a nu-
and the compound’s other local and global parameters are listed in cleophilic site. As the metal approached the ligand, a conforma-
Table 6. tional change was observed as the oxygen atom in the carboxylate
From Table 6, it is observed that the energy gap of the com- group coordinated with the metal atom. Further, the absence of
plex (2.8822 eV) is less compared to the ligand (3.1258 eV) [49,50]. the red region on the oxygen atom of the carboxylate group in the
This decrease in the energy gap of the complex is attributed to metal complex substantiates the significant changes in the elec-
the increased conjugated system compared to the ligand [51]. The tron density distribution in the complex compared to that of the
synthesized palladium complex showed less energy gap compared ligand. Further, the interaction of the chloroform (CHCl3 ) solvent
to other reported palladium complexes [52–55]. The global hard- with the chlorine atom coordinated with the metal atom resulted

9
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 12. 2D fingerprint plots of the synthesized ligand and its corresponding palladium complex.

Table 6
Electronic properties of the ligand and the Pd-complex.

Eg (eV) I (eV) A (eV) χ (eV) η (eV) σ (1/eV) μ (eV) ω (eV)


Ligand
HOMO-LUMO 3.1258 5.6592 2.5334 4.0963 1.5629 0.6398 −4.0963 5.3681
HOMO-1- 5.2798 6.5873 1.3075 3.9474 2.6399 0.3788 −3.9474 2.9513
LUMO+1
Pd complex
HOMO-LUMO 2.8822 6.0766 3.1943 4.6355 1.4411 0.6939 −4.6355 7.4552
HOMO-1- 3.9857 6.6736 2.6879 4.6808 1.9928 0.5018 −4.6808 5.4971
LUMO+1
Where, Eg – Energy gap, I - Ionization energy, A – Electron affinity, χ – Electronegativity, η -Global hardness, σ - Global softness, μ - Chemical
potential, and ω - Global electrophilicity.

in the formation of the electrophilic region, as shown by the red vides better insight into the chemical concepts regarding the bond-
colored area in the metal complex. ing involved in the structure from the wave functions with high
accuracy [57]. The synthesized compounds’ second-order stabiliza-
4. NBO analysis tion energies (E2) are given in Table S4.
The hyperconjugative interaction energies were described using
The natural bond orbital (NBO) analysis helps to investigate the second-order perturbation approach, and the possible intensive
the charge transfer in the molecular system and their interactions interactions are given in Table S4.
among various bonds. The NBO calculations were carried out by The NBO analysis of the metal complex indicated a consider-
investigating all the possible interactions between filled Lewis-type able donor-acceptor type of delocalization between the lone pair
NBOs and empty non-Lewis NBOs, along with estimating their en- of oxygen/nitrogen orbitals and the anti-lone pair of the palladium
ergy by second-order perturbation theory. The NBO analysis pro- orbital. Due to the interactions of the bonding and anti-bonding

10
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 13. Percentage contribution to the Hirshfeld surface by various intermolecular contacts present in the synthesized compounds.

gies 147.47, 73.91, 50.91 and 50.29 kcal/mol, respectively. From Ta-
ble S4, it is clear that the stabilization of the system due to the
overlap of the bonding and anti-bonding orbitals also results in
charge transfer interactions.

4.10. Natural hybrid orbital analysis

The symmetrically orthogonalized hybrid orbital obtained from


the natural atomic orbital (NAO) centered on a particular atom un-
der investigation through unitary transformation results in the nat-
ural hybrid orbitals (NHOs). The direction of each hybrid is speci-
fied in terms of the spherical polar angles theta (θ ) and phi (ϕ )
from the nucleus, and the bending of the bond from the direction
of the line joining the nuclei centers is expressed as the deviation
angle (Dev) [58]. Table S5 summarises the bending of the different
bonds in both the ligand and the complex. In ligand, the oxygen of
σ C14-O3, σ C15-O2, and σ C9-O1 is more bent away from the C–O
line of centers by 8.6, 2.5 and 3°, respectively. As the approach of
the palladium metal resulted in the conformational changes in the
complex compared to that of the ligand. The deviation of the se-
lected atoms in the complex is given in Table S5. The deviation
of the σ C14-O3 and σ C9-O1 in the complex is found to be de-
creased (Dev. 1.8 and 1.9°), whereas a little change in the deviation
was observed for the σ C15-O2 group (Dev. 2.8°). This indicates that
the natural hybrid orbitals are influenced by the approaching group
(palladium metal), which causes the conjugative effect of the steric
Fig. 14. Frontier molecular orbital distribution among the ligand molecule.
effect.

orbitals, the delocalization effects play a crucial role in the coordi- 4.11. QTAIM and NCI analysis
nation environments of the Pd(II) ions.
The interactions energies present in the coordination envi- The QTAIM analysis has proved to be an efficient technique to
ronment are found to be LP(4)Cl1 →LP∗ (1)Pd1 , LP(1)N1 →LP∗ (5)Pd1 , visualize and study the nature of bondings in molecular systems
LP(2)O1 → LP∗ (5)Pd1 , LP(4)Cl1 → LP∗ (6)Pd1 associated with the ener- based on the topology of the electron density at specific points

11
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 15. Changes in the frontier molecular orbital distribution of the ligand due to the formation of the complex.

Fig. 16. The molecular electrostatic potential map of the ligand and the complex with the color range and the scale.

12
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Table 7
QTAIM properties of the synthesized ligand and its complex.

Interactions ρ BCP  2 ρ BCP VBCP GBCP


(ligand) (a.u.) (a.u.) ε (a.u.) (a.u.) HBCP |V|/G

C13-H13···H19 0.01371 0.04966 0.40906 −0.00777 0.0101 −0.0023 0.7693


O3-H3···N5 0.05594 0.11655 0.03429 −0.05468 0.04191 0.01277 1.3047
C27-H27C···O3 0.00737 0.02427 0.03066 −0.00444 0.00525 0.81116 0.8457

Interactions ρ BCP  2 ρ BCP ε VBCP GBCP HBCP |V|/G


([PdLCl]) (a.u.) (a.u.) (a.u.) (a.u.)

C11-H11···S1 0.16461 0.06067 0.92565 −0.01148 0.01332 0.00184 0.8618


C11-H11···H4 0.01530 0.05931 0.45081 −0.00925 0.01204 0.00278 0.76827
C23-H23···Cl1 0.01841 0.05708 0.02322 −0.01097 0.01262 0.00165 0.86925
C23-Cl2···H15B 0.00469 0.01536 0.06012 −0.00208 0.00296 0.00088 2.3636

Electron density (ρ ), Laplacian of electron density ( 2 ρ BCP ), ellipticity (ε ), electron potential energy density (VBCP ), Lagrangian kinetic energy (GBCP ), total
electron energy density (HBCP ) at the corresponding bond critical point (BCP) with the estimated interaction energy (Eint ).

Fig. 17. Molecular graph of the ligand and the complex, indicating the changes of the critical points (CPs).

called bond critical points (BCPs). It allows the possibility of divid- der Waals type of interactions, as indicated by the green region in
ing the system into subsystems based on the zero-flux in the elec- Fig. 18. Further, the NCI isosurface analysis found a few interactions
tron density gradient field. The QTAIM data helps to classify the that were not identified by the AIM analysis.
bond/interaction types, characterized the bond/interaction nature, The presence of the small greens patches in between the re-
and distinguish the bond/interaction strengths. A typical covalent gions of C44-H47···O2 and C44-H46···O2 indicates the small elec-
(shared shell) bond is characterized by ρ (r) > 0.1,  2 ρ (r) < 0, and tron density regions. Further, the reddish green region present in
H(r) < 0 (a.u.), an ionic (closed shell) interaction given by ρ (r) > between the region C38-H40···O4 indicates both attractive and re-
0,  2 ρ (r) > 0 and H(r) > 0 (a.u.) and regarding the 0 < ρ (r) < pulsive interactions due to ring closure. The peaks in the region
0.1,  2 ρ (r) > 0 and H(r) < 0 suggests a dative or electron-transfer −0.05 a.u. are associated with strong, attractive interactions (O2-
bond [59,60]. H3···N5). The peaks in the region −0.01 to 0.005 a.u. and from 0.01
The Schiff base ligand scaffold coordinate to the Pd(II) metal ion to 0.038 a.u. are due to the weak attractive (van der Waals) and re-
results in the electron density redistribution at many critical points pulsive interactions [61].
in the ligand scaffold due to the polarizing effect of the Pd(II) ions.
The QTAIM properties of the ligand and the complex is given in
Table 7.
A significant change in the intermolecular interactions was ob- 4.12.2. Pd complex
served in both the ligand and the metal complexes, which pro- Due to the coordination of the metal with the ligand, a major
moted us to explore the contributions of the weak non-covalent shift in the electron density of the ligand was observed (Fig. 18).
forces in the crystal assembly and the importance of the C–H···π The chelation of the metal with the coordinating atoms resulted
and C–H···Cl interactions establish the organization of the extended in the formation of two chelating rings (Pd1-N1-C11-C10-C9-O1
structure (Fig. 17). and Pd1-N1-C12-C13-C14-O3), which are associated with the pres-
ence of the red region in the ring critical point (RCP) due to the
4.12. NCI isosurface analysis of ligand and the complex steric repulsion. The blue circular disc formed along the coordinate
bond between the Pd1-Cl1 indicates the presence of high electron
4.12.1. Ligand density (Fig. 18). Further, the green region present between the
The red region in the middle of the aromatic and cyclohexane C23-H23···Cl1, C15-H15B···Cl2, and C23-Cl3···O3 indicates the exis-
ring indicates steric repulsion. Interestingly a blue region is ob- tence of weak van der Waals interaction which is attractive in na-
served between the O2-H3···N5 (ligand), indicating a strong hydro- ture. The C23-H23···Cl1 and C15-H15A···Cl2 interactions show the
gen bond interaction due to the high concentration of the elec- presence of the hydrogen halogen interactions exhibited by the
tron density. This results in the formation of a red region (C14- compound. The 2D NCI plot of the complex is associated with a
C7-C10-O2-H3···N5). Due to the presence of the C16-O48 group in large number of peaks compared to that of the ligand 2D NCI plot
close proximity to the O2-H3···N5 results in the formation of van (Fig. 18).

13
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 18. 3D Isosurface and 2D scatter NCI plots of the ligand and the complex reveals various interactions.

Fig. 19. Electron localization function plot of the ligand and the complex, drawn with contour lines.

4.13. Electron localization function (ELF) analysis 4.13.1. Independent gradient model (IGM) analysis
The IGM is another electron-density based computational
The electron localization function is used to evaluate the elec- method which helps in revealing the qualitative and quantita-
tron localization between chemically bonded atoms [62]. The ELF tive study of the non-covalent interactions. The key feature of the
analysis provides an idea regarding the chemical structure, molec- IGM is that it furnishes an uncoupling scheme that automatically
ular bonding, reactivity, and the possibility of finding an electron extracts the signature of the interactions between two different
with the same spin in the neighborhood of a reference electron can molecules (δ ginter ), or inside these molecules (δ gintra ), or between
be studied. The color scheme of red to blue, along with the color two atoms (δ gpair ). The δ ginter term indicates the study of inter-
codes (0.9–1.0), indicates the localization and delocalization of the molecular interactions [63].
electrons, respectively. The electron localization function contour From Fig. 20, it can be observed that the presence of the green
map of the compounds is given in Fig. 19. The red regions are asso- regions indicates the presence of the vdW nature of the weak in-
ciated with hydrogen atoms indicating that the electrons are highly teractions. On the other hand, the IGM analysis of the complex
localized. In contrast, the atomic regions of other atoms (C, N, and showed the presence of a blue isosurface between the coordina-
Cl) are represented by the blue color indicating the delocalization tion bonds (O, N, O, Cl-Pd), revealing the presence of high electron
of the electrons. density and strong bonding interactions in this region [64].

14
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

Fig. 20. 3D isosurfaces of the ligand and the complex obtained by the IGM analysis.

5. Conclusion Riyaz ur Rahaman Khan: Formal analysis, Methodology. N.K.


Lokanath: Resources, Visualization, Writing – review & editing,
The synthesized ligand and its palladium complex were struc- Supervision, Project administration.
turally characterized in this study. The Schiff base ligand was coor-
dinated as a tridentate ligand through O (O1 and O2) and N (N1) Data Availability
donor atoms to the palladium metal atom.
The ligand is observed to be exhibiting a dihedral angle of Data will be made available on request.
15.06°, and due to the complexation with the palladium atom, a
conformational change was observed in the ligand resulting in the Acknowledgments
decrease of dihedral angle in the complex (1.85°). Interestingly, a
The authors Chethan B. S. and Lokanath N. K. are thankful to
rotation of the ligand by 180° was observed about the C13-C14
the DST-FIST (SR/FST/PSI-119/2019), National Single Crystal Diffrac-
bond, allowing the oxygen (O3) atom to chelate with the palladium
tometer Facility, DoS in Physics, CPEPA, IOE and DST-PURSE, Vij-
metal, resulting in the formation of the additional S(5) synthon.
nana Bhavan, University of Mysore, Mysuru for providing the in-
Hydrogen-halogen interactions (C23-H23···Cl1) were observed be-
strumentation facility.
tween the crystallization solvent (CHCl3 ) and the complex. A sig-
nificant change in the electrostatic potential in the molecular elec- Supplementary materials
trostatic potential of the ligand was observed as the metal ap-
proached the ligand leading to a conformational change. The struc- Supplementary material associated with this article can be
tural conformational change was further corroborated by the de- found, in the online version, at doi:10.1016/j.molstruc.2022.134759.
creased deviation of the chelating atoms’ nuclei centers compared
to that of the ligand from the natural hybrid orbital analysis. References
The QTAIM analysis revealed the importance of the C–H···π and
C–H···Cl interactions in establishing the organization of the ex- [1] S.A. Dalia, F. Afsan, M.S. Hossain, M.N. Khan, C. Zakaria, M. Zahan, M. Ali, A
short review on chemistry of schiff base metal complexes and their catalytic
tended structure. The non-covalent interactions revealed the im- application, Int. J. Chem. Stud. 6 (2018) 2859–2866.
portance of the various interactions towards the crystal packing. [2] X. Liu, J.R. Hamon, Recent developments in penta-, hexa-and heptadentate
The blue isosurface between the coordinating bonds with the pal- Schiff base ligands and their metal complexes, Coord. Chem. Rev. 389 (2019)
94–118.
ladium metal indicated strong bonding interactions. This study em- [3] M.S. Hossain, P.K. Roy, C. Zakaria, M. Kudrat-E-Zahan, Selected Schiff base co-
phasizes understanding the effects on the conformation of the lig- ordination complexes and their microbial application: a review, Int. J. Chem.
and due to the complexation process and the importance of struc- Stud. 6 (1) (2018) 19–31.
[4] I. Kostova, L. Saso, Advances in research of Schiff-base metal complexes as po-
tural techniques, theories, and computational tools in achieving the tent antioxidants, Curr. Med. Chem. 20 (36) (2013) 4609–4632.
goal. The study of the changes in the conformation of the ligand [5] E. Ispir, The synthesis, characterization, electrochemical character, catalytic and
and also the changes in the intermolecular interactions due to the antimicrobial activity of novel, azo-containing Schiff bases and their metal
complexes, Dye. Pigments 82 (1) (2009) 13–19.
approach of the metal atom by experimental and theoretical may
[6] K. Divya, G.M. Pinto, A.F. Pinto, Application of metal complexes of Schiff bases
lead in the direction of the crystal packing pattern modification, as an antimicrobial drug: a review of recent works, Int. J. Curr. Pharm. Res. 9
which in turn facilitates the tuning of the physicochemical and (3) (2017) 27–30.
thermophysical properties of the compounds. [7] S. Shekhar, A.M. Khan, S. Sharma, B. Sharma, A. Sarkar, Schiff base metallo-
drugs in antimicrobial and anticancer chemotherapy applications: a compre-
hensive review, Emergent Mater. 5 (2021) 1–15.
Declaration of Competing Interest [8] S. Alyar, Ü.Ö. Özmen, Ş. Adem, H. Alyar, E. Bilen, K. Kaya, Synthesis, spectro-
scopic characterizations, carbonic anhydrase II inhibitory activity, anticancer
activity and docking studies of new Schiff bases of sulfa drugs, J. Mol. Struct.
There are no conflicts to declare. 1223 (2021) 128911.
[9] J.M. Mir, S.A. Majid, A.H. Shalla, Enhancement of Schiff base biological efficacy
by metal coordination and introduction of metallic compounds as anticovid
CRediT authorship contribution statement candidates: a simple overview, Rev. Inorg. Chem. 41 (4) (2021) 199–211.
[10] V. Rangaswamy, S. Renuka, I. Venda, Synthesis, spectral characterization and
B.S. Chethan: Writing – original draft, Conceptualization, antibacterial activity of transition metal (II) complexes of tetradentate Schiff
base ligand, Mater. Today Proc. 51 (2022) 1810–1816.
Methodology, Visualization, Software, Formal analysis, Data [11] A. Banerjee, D. Chowdhury, R.M. Gomila, S. Chattopadhyay, Recent advances
curation. H.R. Rajegowda: Formal analysis, Methodology. on the tetrel bonding interaction in the solid state structure of lead complexes

15
B.S. Chethan, H.R. Rajegowda, R.u.R. Khan et al. Journal of Molecular Structure 1276 (2023) 134759

with hydrazine based bis-pyridine Schiff base ligands, Polyhedron 216 (2022) novel chiral acyclic tellurated Schiff base ligands, New J. Chem. 42 (8) (2018)
115670. 6264–6273.
[12] A. Aragón-Muriel, Y. Liscano, Y. Upegui, S.M. Robledo, M.T. Ramírez-Apan, [41] W.P. Crisp, R.L. Deutscher, D.L. Kepert, Six-and eight-co-ordinate complexes of
D. Morales-Morales, J. Oñate-Garzón, D. Polo-Cerón, In vitro evaluation of the titanium tetrachloride, J. Chem. Soc. A: Inorganic, Physical, Theoretical. (1970)
potential pharmacological activity and molecular targets of new benzimida- 2199–2201.
zole-based schiff base metal complexes, Antibiotics 10 (6) (2021) 728. [42] M.M. Tamizh, K. Mereiter, K. Kirchner, B.R. Bhat, R. Karvembu, Synthesis, crys-
[13] D.M. Yufanyi, H.S. Abbo, S.J. Titinchi, T. Neville, Platinum (II) and ruthenium tal structures and spectral studies of square planar nickel (II) complexes con-
(II) complexes in medicine: antimycobacterial and Anti-HIV activities, Coord. taining an ONS donor Schiff base and triphenylphosphine, Polyhedron 28 (11)
Chem. Rev. 414 (2020) 213285. (2009) 2157–2164.
[14] Q.H. Tran, T.T. Doan, A novel study on curcumin metal complexes: solubil- [43] M.K. Biyala, K. Sharma, M. Swami, N. Fahmi, R.V. Singh, Spectral and biocidal
ity improvement, bioactivity, and trial burn wound treatment in rats, New J. studies of palladium (II) and platinum (II) complexes with monobasic biden-
Chem. 44 (30) (2020) 13036–13045. tate Schiff bases, Transit. Met. Chem. 33 (3) (2008) 377–381.
[15] H. Lu, D.S. Wright, S.D. Pike, The use of mixed-metal single source precur- [44] S.Y. Li, C.J. Chen, P.Y. Lo, H.S. Sheu, G.H. Lee, C.K. Lai, H-bonded metallomeso-
sors for the synthesis of complex metal oxides, Chem. Commun. 56 (6) (2020) gens derived from salicyladiminates, Tetrahedron 66 (32) (2010) 6101–6112.
854–871. [45] B. Bosnich, An interpretation of the circular dichroism and electronic spectra
[16] T. Palacios-Hernández, G.A. Hirata-Flores, O.E. Contreras-López, M.E. Mendoza- of salicylaldimine complexes of square-coplanar diamagnetic nickel (II), J. Am.
-Sánchez, I. Valeriano-Arreola, E. González-Vergara, M.A. Méndez-Rojas, Syn- Chem. Soc. 90 (3) (1968) 627–632.
thesis of Cu and Co metal oxide nanoparticles from thermal decomposition of [46] R.M. Silverstein, G.C. Bassler, Spectrometric identification of organic com-
tartrate complexes, Inorg. Chim. Acta 392 (2012) 277–282. pounds, J. Chem. Educ. 39 (11) (1962) 546.
[17] V. Flores-Romero, O.L. García-Guzmán, A. Aguirre-Bautista, I.D. Rojas-Montoya, [47] J. Chang, S.Z. Zhang, Y. Wu, H.J. Zhang, Y.X. Sun, Three supramolecular trinu-
V. García-Montalvo, M. Rivera, O. Jiménez-Sandoval, M.Á. Muñoz-Hernández, clear nickel (II) complexes based on Salamo-type chelating ligand: syntheses,
S. Hernández-Ortega, Zinc (ii) and cadmium (ii) complexes,[M (i Pr 2 P (X) NC crystal structures, solvent effect, Hirshfeld surface analysis and DFT calculation,
(Y) NC 5 H 10-κ 2-X, Y) 2](X and Y= O, S), as single source precursors for Transit. Met. Chem. 45 (4) (2020) 279–293.
metal sulfide thin films, New J. Chem. 44 (25) (2020) 10367–10379. [48] V. Torabi, H. Kargar, A. Akbari, R. Behjatmanesh-Ardakani, H. Amiri Rudbari,
[18] H.U. Islam, A. Roffey, N. Hollingsworth, W. Bras, G. Sankar, N.H. De Leeuw, M. Nawaz Tahir, Nickel (II) complex with an asymmetric tetradentate Schiff
G. Hogarth, Understanding the role of zinc dithiocarbamate complexes as base ligand: synthesis, characterization, crystal structure, and DFT studies, J.
single source precursors to ZnS nanomaterials, Nanoscale Adv. 2 (2) (2020) Coord. Chem. 71 (22) (2018) 3748–3762.
798–807. [49] G.S. Priyadarshini, V. Edathil, G. Selvi, Non linear optical properties of potent
[19] J. Pflugrath, The finer things in X-ray diffraction data collection, Acta Crystal- quinoline based schiff bases, Mater. Today Proc. 51 (2022) 1746–1750.
logr. Sect. D Biol. Crystallogr. 55 (10) (1999) 1718–1725. [50] C.U. Ibeji, K. Ukogu, M.T. Kelani, F.E. Ani, N.L. Obasi, S.A. Ogundare,
[20] C.S. Expert, Rigaku corporation, Tokyo, Japan (2011). G.E. Maguire, H.G. Kruger, Synthesis, Crystal structure, photolumines-
[21] O.V. Dolomanov, L.J. Bourhis, R.J. Gildea, J.A. Howard, H. Puschmann, OLEX2: a cence properties and quantum mechanics studies of two schiff bases of
complete structure solution, refinement and analysis program, J. Appl. Crystal- 2-amino-p-cresol, J. Mol. Struct. 1262 (2022) 133046.
logr. 42 (2) (2009) 339–341. [51] U. Solmaz, H. Arslan, Spectral, crystallographic, theoretical, and catalytic activ-
[22] G.M. Sheldrick, Phase annealing in SHELX-90: direct methods for larger struc- ity studies of the PdII complexes in different coordination modes of benzoylth-
tures, Acta Crystallogr. Sect. A Found. Crystallogr. 46 (6) (1990) 467–473. iourea ligand, J. Mol. Struct. 1269 (2022) 133839.
[23] A. Spek, Structure validation in chemical crystallography, Acta Crystallogr. Sect. [52] A.S. Al-Janabi, T.A. Yousef, M.E. Al-Doori, R. Bedier, B.M. Ahmed, Palladium
D 65 (2) (2009) 148–155. (II)-salicylanilide complexes as antibacterial agents: synthesis, spectroscopic,
[24] C.F. Macrae, I. Sovago, S.J. Cottrell, P.T. Galek, P. McCabe, E. Pidcock, M. Platings, structural characterization, DFT calculations, biological and in silico studies, J.
G.P. Shields, J.S. Stevens, M. Towler, Mercury 4.0: from visualization to analysis, Mol. Struct. 1246 (2021) 131035.
design and prediction, J. Appl. Crystallogr. 53 (1) (2020) 226–235. [53] M. Akram, M. Adeel, M. Khalid, M.N. Tahir, H.A.R. Aliabad, M.A. Ullah, J. Iqbal,
[25] M. Turner, J. McKinnon, S. Wolff, D. Grimwood, P. Spackman, D. Jayatilaka, A.A. Braga, Highly efficient one pot palladium-catalyzed synthesis of 3, 5-bis
M. Spackman, CrystalExplorer17, The University of Western Australia, 2017 (arylated) pyridines: comparative experimental and DFT studies, J. Mol. Struct.
Australia. 1213 (2020) 128131.
[26] P.R. Spackman, M.J. Turner, J.J. McKinnon, S.K. Wolff, D.J. Grimwood, D. Jayati- [54] S. Ahmad, S. Nadeem, A. Anwar, A. Hameed, S.A. Tirmizi, W. Zierkiewicz, A. Ab-
laka, M.A. Spackman, CrystalExplorer: a program for Hirshfeld surface analysis, bas, A.A. Isab, M.A. Alotaibi, Synthesis, characterization, DFT calculations and
visualization and quantitative analysis of molecular crystals, J. Appl. Crystal- antibacterial activity of palladium (II) cyanide complexes with thioamides, J.
logr. 54 (3) (2021) 1006–1011. Mol. Struct. 1141 (2017) 204–212.
[27] J.J. McKinnon, A.S. Mitchell, M.A. Spackman, Hirshfeld surfaces: a new tool [55] N. Kystaubayeva, F. Durap, T. Zharkynbek, V. Yu, M. Aydemir, N. Meriç, A. Za-
for visualising and exploring molecular crystals, Chem. Eur. J. 4 (11) (1998) zybin, А.Y. Ten, K.S. Rafikova, N.E. Binbay, Synthesis, and characterization of
2136–2141. palladium (II) and platinum (II) complexes with α –hydroxy [1-(2-ethoxyethyl)
[28] D.S. Sholl, J.A. Steckel, Density Functional theory: A Practical Introduction, John piperidin-4-yl] phosphonate and use of the palladium (II) complex as pre-cat-
Wiley & Sons, 2011. alyst in Suzuki-Miyaura cross-coupling reactions, J. Mol. Struct. 1270 (2022)
[29] B. Miehlich, A. Savin, H. Stoll, H. Preuss, Results obtained with the correlation 133912.
energy density functionals of Becke and Lee, Yang and Parr, Chem. Phys. Lett. [56] H. Kargar, M. Bazrafshan, M. Fallah-Mehrjardi, R. Behjatmanesh-Ardakani,
157 (3) (1989) 200–206. H.A. Rudbari, K.S. Munawar, M. Ashfaq, M.N. Tahir, Synthesis, characterization,
[30] M.E. Frisch, G. Trucks, H. Schlegel, G. Scuseria, M. Robb, J. Cheeseman, crystal structures, Hirshfeld surface analysis, DFT computational studies and
G. Scalmani, V. Barone, G. Petersson, H. Nakatsuji, Gaussian 16, Gaussian, Inc, catalytic activity of novel oxovanadium and dioxomolybdenum complexes with
Wallingford, CT, 2016. ONO tridentate Schiff base ligand, Polyhedron 202 (2021) 115194.
[31] R. Krishnan, J.S. Binkley, R. Seeger, J.A. Pople, Self-consistent molecular orbital [57] C.B. Sundaramurthy, C.P.K. Nataraju, L.N. Krishnappagowda, Design, synthesis,
methods. XX. A basis set for correlated wave functions, J. Chem. Phys. 72 (1) structural analysis and quantum chemical insight into the molecular structure
(1980) 650–654. of coumarin derivatives, Mol. Syst. Des. Eng. 7 (2) (2022) 132–157.
[32] F. Bobrowitcz, W. Goddard III, H.F. Schaefer III, Modern Theoretical Chemistry, [58] J. Foster, F. Weinhold, 111 Am. Chem. Soc. 102 (1980) 7211. Foster, A. J., & Wein-
Vol. 3, Plenum, New York, 1977 Ed. hold, F. (1980). Natural hybrid orbitals. Journal of the American Chemical Society,
[33] T. Koopmans, Über die Zuordnung von Wellenfunktionen und Eigenwerten zu 102(24), 7211-7218.
den einzelnen Elektronen eines Atoms, Physica 1(1-6) (1934) 104–113. [59] R.F. Bader, T. Nguyen-Dang, Quantum theory of atoms in molecules–Dalton re-
[34] R. Dennington, T. Keith, J. Millam, GaussView 5.0, 20, Gaussian, Inc., Walling- visited, in: Advances in Quantum Chemistry, Elsevier, 1981, pp. 63–124.
ford, 2008. [60] S. Kamaal, M. Usman, M. Afzal, A. Alarifi, A. Ali, R. Das, P. Lama, M. Ahmad, A
[35] R.F. Bader, H. Essén, The characterization of atomic interactions, J. Chem. Phys. new copper (II)-based layered coordination polymer: crystal structure, topol-
80 (5) (1984) 1943–1960. ogy, QTAIM analysis, experimental and theoretical magnetic properties based
[36] E.R. Johnson, S. Keinan, P. Mori-Sánchez, J. Contreras-García, A.J. Cohen, on DFT combined with broken-symmetry formalism (BS-DFT), Polyhedron 193
W. Yang, Revealing noncovalent interactions, J. Am. Chem. Soc. 132 (18) (2010) (2021) 114881.
6498–6506. [61] B. Chethan, N. Lokanath, Study of the crystal structure, H-bonding and nonco-
[37] W. Humphrey, A. Dalke, K. Schulten, VMD: visual molecular dynamics, J. Mol. valent interactions of novel cocrystal by systematic computational search ap-
Graph. 14 (1) (1996) 33–38. proach, J. Mol. Struct. 1251 (2022) 131936.
[38] M. Somashekar, P. Chetana, B. Chethan, H. Rajegowda, M. Cooper, Z. Ziora, [62] A. Savin, R. Nesper, S. Wengert, T.F. Fässler, ELF: the electron localization func-
N. Lokanath, P.S. Ganapathy, B. Srinatha, Synthesis and characterization of Zinc tion, Angew. Chem. Int. Ed. Engl. 36 (17) (1997) 1808–1832.
(II) complex with ONO donor type new phenylpropanehydrazide based lig- [63] K. Rayene, D. Imane, B. Abdelaziz, N. Leila, M. Fatiha, G. Abdelkrim,
and: crystal structure, Hirshfeld surface analysis, DFT, energy frameworks and G. Bouzid, L. Ismahan, H. Brahim, O. Rabah, Molecular modeling study
molecular docking, J. Mol. Struct. 1255 (2022) 132429. of structures, Hirschfield surface, NBO, AIM, RDG, IGM and 1HNMR of
[39] N. Turan, K. Buldurun, N. Çolak, İ. Özdemir, Preparation and spectroscopic thymoquinone/hydroxypropyl-β -cyclodextrin inclusion complex from QM cal-
studies of Fe (II), Ru (II), Pd (II) and Zn (II) complexes of Schiff base contain- culations, J. Mol. Struct. 1249 (2022) 131565.
ing terephthalaldehyde and their transfer hydrogenation and Suzuki-Miyaura [64] H. Alizadeh, M. Mirzaei, A.S. Saljooghi, V. Jodaian, M. Bazargan, J.T. Mague,
coupling reaction, Open Chem. 17 (1) (2019) 571–580. R.M. Gomila, A. Frontera, Coordination complexes of zinc and manganese
[40] H. Rajegowda, P.R. Kumar, A. Hosamani, R. Butcher, Synthesis, characteri- based on pyridine-2, 5-dicarboxylic acid N-oxide: DFT studies and antiprolifer-
zation and determination of absolute structures of palladium complexes of ative activities consideration, RSC Adv. 11 (59) (2021) 37403–37412.

16

You might also like