You are on page 1of 16

Inorganica Chimica Acta 547 (2023) 121368

Contents lists available at ScienceDirect

Inorganica Chimica Acta


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

Spectral studies, crystal structures, DNA binding, and anticancer potentials


of Pd(II) complexes with iminophosphine ligands: Experimental and
computational methods
Mahsa Pooyan a, Zahra Shariatinia a, *, Fahimeh Mohammadpanah b, Khodayar Gholivand b, *,
Peter C. Junk c, Zhifang Guo c, Mohammad Satari d, Vahid Noroozi Charandabi e
a
Department of Chemistry, Amirkabir University of Technology (Tehran Polytechnic), P.O.Box:15875-4413, Tehran, Iran
b
Department of Chemistry, Faculty of Science, Tarbiat Modares University, P.O. Box 14115-175, Tehran, Iran
c
College of Science & Engineering, James Cook University, Townsville, Queensland, 4811, Australia
d
Department of Biology, Faculty of Sciences, Malayer University, Malayer, Iran
e
Kimia Andish Laboratory, KAGI Company, Tehran, Iran

A R T I C L E I N F O A B S T R A C T

Keywords: Herein, we studied the synthesis and the biological properties of palladium(II) complexes derived from imino­
Iminophosphine ligands phosphine ligands DPPB-3-hydroxybenzohydrazone (DPPB-HBH), DPPB-4-phenylthiosemicarbazone (DPPB-
Molecular docking PTSC), and DPPB-thiosemicarbazone (DPPB-TSC) (DPPB = 2-(diphenylphosphino)benzaldehyde) using theo­
Palladium complexes
retical and experimental methods. The molecular structures of [PdCl(DPPB-HBH)]Cl, [Pd(DPPB-HBH)2]
DFT calculations
DNA binding
Cl2⋅2CH3CN, [PdCl(DPPB-PTSC)], and [PdCl(DPPB-TSC)] were characterized by various spectroscopic tech­
Cytotoxicity niques and X-ray crystallographic for [PdCl(DPPB-HBH)]Cl, [Pd(DPPB-HBH)2]Cl2⋅2CH3CN, and [PdCl(DPPB-
PTSC)], confirming the distorted square-planar geometry for the palladium atoms. Complexes with the gen­
eral formula [PdCl(Ligand)x] (x = 1) were employed to evaluate biological activities such as cytotoxicity against
the MDA-MB-231 breast cancer cells and DNA binding ability. The results revealed the high cytotoxicity of the
palladium complexes compared to the reference drug cisplatin (IC50 = 24.59 µg/mL), as well as their moderate
binding ability to DNA through electrostatic interactions and groove binding. It was also found that [PdCl(DPPB-
HBH)]Cl has the highest activity among others (IC50 = 9.3 µg/mL). Their interaction mode with DNA was
investigated using molecular docking and NCI calculations, which showed complete agreement with the
experimental results. Furthermore, quantum chemistry calculations were used to analyze the structure–activity
relationship and introduce reactive sites, which determined the cause of the difference in the reactivity of the
samples.

1. Introduction of structure and biology so that they cover a wide range of biological
applications such as anticancer, antibacterial, and antifungal activities
Despite significant advances in medical science, cancer is still [5–9]. Palladium complexes are much more reactive and sensitive to
considered the most common disease in the world [1]. Organoplatinum hydrolysis than their platinum counterparts. However, their low cost,
drugs such as cisplatin, oxaliplatin, and carboplatin are among the most high solubility, and greater tendency to interact with DNA prompted
widely employed anticancer drugs, but due to side effects such as renal researchers to improve the properties of the complexes and increase
and hepatotoxicity, dose limitation, and drug resistance, their long-term their stability by choosing appropriate ligands [5,10–13]. Iminophos­
treatment is not recommended [2–4]. Hence much research is being phines are a group of polyfunctional chelating ligands that contain both
done in the field of introducing non-platinum complexes as alternative hard and soft donor atoms in their structure [14–16]. These ligands
drugs to overcome these limitations. Palladium complexes are a group of stabilize the construction of the complex by forming chelate rings
these candidates that have attracted the attention of researchers due to around the central metal due to weak π-accepting and strong σ-donating
their similar coordination chemistry with platinum complexes in terms characteristics and also increase the reactivity of the complex by

* Corresponding authors.
E-mail addresses: shariati@aut.ac.ir (Z. Shariatinia), gholi_kh@modares.ac.ir (K. Gholivand).

https://doi.org/10.1016/j.ica.2022.121368
Received 18 September 2022; Received in revised form 23 December 2022; Accepted 23 December 2022
Available online 28 December 2022
0020-1693/© 2022 Elsevier B.V. All rights reserved.
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

creating asymmetry in the metal orbitals [17,18]. MHz); δ = –16.65 ppm.


Hydrazones and thiosemicarbazones derived from 2-(diphenyl­
phosphino)benzaldehyde are a group of iminophosphine ligands that 2.2.2. Synthesis of DPPB-PTSC
contain PNO, and PNS donor atoms, respectively. These compounds A solution of 2-(diphenylphosphino) benzaldehyde (0.43 mmol)
show different biological activities depending on the coordinated metal dissolved in 25 mL of hot methanol with a few droplets of glacial acetic
ion and coordination geometry and the number of donor atoms [19–22]. acid was poured into a hot methanolic solution of 4-phenylthiosemicar­
In addition, their metal complexes are effective on a wide range of bazide (0.43 mmol). After refluxing for 18 h, finally, slow solvent
cancer cells, and some of them exhibit similar and even higher cyto­ evaporation yielded a yellow powder, which was rinsed by diethyl ether
toxicity than the reference drug cisplatin [19,23–25]. Although the after extraction from the remaining solvent.
biological properties of palladium complexes of these derivatives have (2-(diphenylphosphino) benzaldehyde 4-phenylthiosemicarbazone)
been investigated in a limited way, the results indicate their significant (DPPB-PTSC)
antifungal and anticancer activities [5,21,22,26–29]. (C6H5)2P(o-C6H4-CH = N-NH-C(S)-NHC6H5) (DPPB-PTSC). M.p.
Considering the background presented on the anticancer potential of 206 ◦ C. FT-IR (cm− 1): ν(N–H)ph = 3443w, br, 3302w, br, ν(N–H)hyd =
palladium complexes and the interesting coordination behavior of the 3135w, br, ν(C–H)arom. = 2990w, ν(C– – N) = 1541s, ν(C–P) = 1438m
iminophosphines, in this paper, we investigate the structural properties ν(C–– S) = 1190s. 1H NMR (300.13 MHz, ppm); δ=: 11.93 (s, 1H, N–H),
and biological activity of palladium complexes derived from the 9.89 (s, 1H, NH-C6H5), 8.77 (d, 1H, JP-H = 4.67, CH = N), 6.80–8.32 (m,
condensation reactions of 2-(diphenylphosphino)benzaldehyde with 3- aromatic hydrogens). 31P(1H) and 31P NMR (121.49 MHz); δ = –14.17
hydroxybenzohydrazide, 4-phenylthiosemicarbazide, and thio­ ppm.
semicarbazide. The structures of the molecules were characterized using
various spectroscopic techniques and X-ray crystallography. Given that 2.2.3. Synthesis of DPPB-TSC
the main target of most anticancer drugs is DNA, the tendency of the Synthesis of DPPB-TSC was accomplished according to the literature
candidate palladium complexes to interact with DNA was assessed by method [15].
fluorescence and UV–vis spectroscopic methods. Also, their cytotoxicity (2-(diphenylphosphino) benzaldehyde thiosemicarbazone) (DPPB-
was evaluated against triple-negative breast cancer cells (MDA-MB- TSC)
231). Furthermore, to create a logical relationship between structure (C6H5)2P(o-C6H4-CH = N-NH-C(S)–NH2) (DPPB-TSC). M.p. 225 ◦ C.
and activity and to gain insight into the reason for the difference in the FT-IR (cm− 1): ν(NH2)asym = 3442w, br, ν(NH2)sym = 3306w, br, ν(N–H) =
biological potential of compounds, DFT and MEP studies were also 3166w, br, ν(C–H)arom. = 3016w, ν(C– – N) = 1574s, ν(C–P) = 1436m,
conducted. – S) = 1218s. 1H NMR (300.13 MHz, ppm); δ=: 11.51 (s, 1H,
1366s, ν(C–
N–H), 8.64 (d, 1H, JP-H = 4.61, CH = N), 8.17–8.14 (q, 2H, NH2) 6.74
2. Experimental –7.68 (m, 14H, Ph-P). 31P(1H) and 31P NMR (121.49 MHz); δ = –14.26
ppm.
2.1. Materials and methods
2.3. Synthesis procedure of complexes
All compounds used in this research were > 95 % pure because they
were provided by Sigma-Aldrich and Merck Companies. Materials used A solution of [PdCl2(COD)] (0.0125 mmol for [PdCl(DPPB-HBH)]Cl
in biological tests such as DNA, MTT (3–(4,5–dimethyl thiazole–2–yl)– and [Pd(DPPB-HBH)2]Cl2⋅2CH3CN, 0.06 mmol for [PdCl(DPPB-PTSC)]
2,5–diphenyltetrazolium bromide), ethidium bromide (EB) were ob­ and [PdCl(DPPB-TSC)]) in acetonitrile (10 mL) was poured into aceto­
tained from Sigma Corporation, but FBS (fetal bovine serum) and high nitrile solutions (20 mL) of ligands in 1:1 (for [PdCl(DPPB-PTSC)] and
glucose DMEM medium from GiBCO Company (Grand Island, NY, USA). [PdCl(DPPB-TSC)]) and 1:2 M ratios of metal to ligand (for [PdCl(DPPB-
Cell Bank at Pasteur Institute of Iran (Tehran) provided the MDA-MB- HBH)]Cl and [Pd(DPPB-HBH)2]Cl2⋅2CH3CN). Then, these solutions
231 cell line. were placed under ambient conditions for several days. Finally, slow
Electrothermal apparatus was utilized to determine melting points. solvent evaporation obtained suitable crystals for X-ray crystallography
The spectrophotometer Perkin–Elmer (Lambda 25) and Nicolet 510P determination.
spectrometer were applied to take UV–Vis and FT- IR spectra, respec­
tively. Bruker Avance DRX (300.13 MHz for 1H NMR spectra and 121.49 2.3.1. [PdCl(DPPB-HBH)]Cl
MHz for 31P NMR spectra) was used to acquire NMR spectra in d6–DMSO Yield = 45 % (yellow color). M.p. 250 ◦ C. FT-IR (cm− 1): ν(O–H)asym
using 85 % H3PO4 and tetramethylsilane as standards for 31P- and 1H = 3427w, br, ν(O–H)sym = 3122w, br ν(N–H) = 3059w, br, ν(C–H)arom. =
NMR spectra, respectively. 2929w, 2861w, ν(C– – N) = 1582s, ν(C–P) = 1437m, ν(C– – O) = 1648s. 1H
NMR (300.13 MHz, ppm); δ=: 9.08 (s, 1H, N–H), 9.99 (s, 1H, O–H),
2.2. Synthesis procedures of ligands 8.45 (d, 1H, CH = N), 6.98–7.70 (m, aromatic hydrogens). 31P (1H) and
31
P NMR (121.49 MHz); δ = 35.30 ppm.
2.2.1. Synthesis of DPPB-HBH
A mixture of 0.43 mmol (0.127 g) of 2-(diphenylphosphino) benz­ 2.3.2. [Pd(DPPB-HBH)2]Cl2⋅2CH3CN
aldehyde within ethanol (20 mL) and some droplets of glacial acetic acid Red; yield = 15 %. M.p. 265 ◦ C. 1H NMR (300.13 MHz, ppm); δ=:
were poured into the ethanolic solution (20 mL) of 3-hydroxybenzohy­ 9.09 (d, 1H, O–H), 8.49 (d, 1H, CH = N), 7.72–6.91 (m, 36Harom, Ph).
31
drazide (0.43 mmol, 0.065 g) at room temperature. After refluxing for P (1H) and 31P NMR (121.49 MHz); δ = 33.32 ppm.
8 h, a white solid was obtained through slow evaporation of the solvent,
which was subsequently separated and washed with diethyl ether. 2.3.3. [PdCl(DPPB-PTSC)]
(2-(diphenylphosphino) benzaldehyde 3-hydroxybenzohydrazone) M.p. 286 ◦ C. FT-IR (cm− 1): = ν(N–H)ph = 3405w, br, 3262w, br,
(DPPB-HBH) ν(N–H)hyd = 3049w, br, ν(C–H)arom. = 2924w, ν(C– – N) = 1490s,
(C6H5)2P(o-C6H4-CH = N-NH-C(O)(m-C6H4OH)) (DPPB-HBH). M.p. ν(C–P) = 1430m ν(C– – S) = 1180s. 1H NMR (300.13 MHz, ppm); δ=:
130 ◦ C. FT-IR (cm− 1): ν(O–H)asym = 3403w, br, ν(O–H)sym = 3225w, br 8.67 (s, 1H, N–H), 8.54 (d, 1H, NH-C6H5), 8.46 (d, 1H, CH = N),
ν(N–H) = 3058w, br, ν(C–H)arom. = 2971w, 2928w ν(C– – O) = 1645s, 7.09–7.69 (m, aromatic protons).
1
ν(C– N) = 1588s, ν(C P) = 1443m. H NMR (300.13 MHz, ppm); δ=:
– –
11.58 (s, 1H, N–H), 11.93 (s, 1H, O–H), 9.22 (d, 1H, CH = N), 2.3.4. [PdCl(DPPB-TSC)]
6.53–7.64 (m, aromatic hydrogens). 31P(1H) and 31P NMR (121.49 M.p. 290 ◦ C. FT-IR (cm− 1): ν(NH2)asym = 3400w, br, ν(NH2)sym =

2
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

3281w, br, ν(N–H) = 3145w, br, ν(C–H)arom. = 2925w, ν(C–– N) = 1470s, Table 1
ν(C–P) = 1394m, ν(C– – S) = 1154s. 1H NMR (300.13 MHz, ppm); δ=: Crystal data and structural refinement parameters of complexes. [Pd(DPPB-
9.16 (s, 1H, N–H), 7.66 (m, 2H, NH2), 8.36 (d, 1H, CH = N), 6.04–7.97 HBH)2]Cl2•2CH3CN, [PdCl(DPPB-HBH)]Cl, and [PdCl(DPPB-PTSC)].
(m, 14H, Ph-P). 31P (1H) and 31P NMR (121.49 MHz); δ = 22.54 ppm. Compound [Pd(DPPB-HBH)2] [PdCl(DPPB-HBH)] [PdCl(DPPB-
Cl2•2CH3CN Cl PTSC)]
2.4. X-ray crystallographic details Empirical C56H48Cl2N6O4P2Pd C26H21Cl2N2O2PPd C26H22ClN3PPdS
formula
Formula 1108.24 601.72 581.34
Single crystals of complexes [PdCl(DPPB-HBH)]Cl, [Pd(DPPB-
weight
HBH)2]Cl2⋅2CH3CN, and [PdCl(DPPB-PTSC)] were dipped into a viscous T (K) 100(2) 100(2) 100(2)
hydrocarbon oil and mounted onto a loop to determine their structures Crystal system, Monoclinic, C2/c Triclinic, P-1 Monoclinic, P21/
using Australian Synchrotron over MX1 macromolecular beamlines at space group c
100 K and λ = 0.71073 Å. All diffractions were indexed and integrated a(Å) 22.326(5) 9.5020(19) 9.5520(19)
17.221(3) 15.332(3) 13.663(3)
by the XDS software suite [30]. SHELXT was applied to solve data
b(Å)
C(Å) 15.132(3) 17.398(4) 18.321(4)
refining against F2 via the full-matrix least-squares method with α 90 74.67(3) 90
SHELXL-2018 in Olex2 [31,32]. The riding model theoretically deter­ β 118.39(3) 87.18(3) 100.11(3)
mined the positions of all H-atoms. The CCDC numbers of Pd(II) com­ γ 90 86.34(3) 90
plexes are 1,580,485 (for [PdCl(DPPB-HBH)]Cl), 1,580,470 (for [Pd V 5118(2) 2438.0(9) 2353.9(8)
Z 4 4 4
(DPPB-HBH)2]Cl2⋅2CH3CN), and 1,580,469 (for [PdCl(DPPB-PTSC)]). Dcalc (Mg/ 1.438 1.639 1.640
Also, the crystallographic information is available freely on http m− 3(− |-))
s://www.ccdc.cam.ac.uk/data_request/cif (Cambridge Crystallo­ Absorption 0.583 1.073 1.079
graphic Data Centre). The crystal structure of the complexes and crystal coefficient
(mm− 1)
data and refinement parameters are presented in Fig. 2 and Table 1,
F(000) 2272 1208 1172
respectively. Crystal size 0.11 × 0.05 × 0.02 0.10 × 0.04 × 0.02 0.09 × 0.03 ×
(mm) 0.01
2.5. DNA binding studies θ rang for data 1.9–24.99 1.21–24.99 1.87–24.99
collection (◦ )
Limiting − 26 ≤ h ≤ 26 − 11 ≤ h ≤ 11 − 11 ≤ h ≤ 11
2.5.1. Electronic absorption titration method indices − 20 ≤ k ≤ 20 − 18 ≤ k ≤ 18 − 15 ≤ k ≤ 15
Electronic absorption and fluorescence spectroscopy techniques − 17 ≤ l ≤ 17 − 20 ≤ l ≤ 19 − 21 ≤ l ≤ 21
examined the tendency of Pd (II) complexes to interact with ct-DNA. All Total 30,906 19,591 28,137
reflections
tests were accomplished under ambient conditions within buffer Tris-
Unique 4501 [R(int) = 7895 [R(int) = n/a] 4014[R(int) =
HCl solution composed of NaCl (50 mM in deionized water) + Tris- reflections 0.0605] 0.0351]
HCl (5 mM) at pH 7.2, adjusted by concentrated hydrochloric acid. The (Rint)
ct-DNA standard solution concentration was measured by its absorbance Completeness 99.7 % (θ = 25.00) 91.9 % (θ = 25.00) 96.9 % (θ =
(260 nm) via the Beer-Lambert equation (molar absorption coefficient to θ(%) 25.00)
Data/ 4501/0/324 7895/6/ 616 4014/0/299
was 6600 M− 1cm− 1). This solution was used within three days. The
restraints/
constant 48 µM concentration of each complex ([PdCl(DPPB-HBH)]Cl, parameters
[PdCl(DPPB-PTSC)], and [PdCl(DPPB-TSC)]) within Tris-HCl buffer/ Refinement Full-matrix least Full-matrix least Full-matrix least
DMSO (less than 1 % final volume) mixture was titrated by different ct- method squares on F2 squares on F2 squares on F2
Goodness-of-fit 1.050 1.064 1.061
DNA concentrations (9.94–170 µM) to perform the electronic absorption
(GOF) on F2
titration assays. Various ct-DNA concentrations were added into each Final R indices R1 = 0.0308, wR2 = R1 = 0.0666, wR2 R1 = 0.0241,
complex solution with a fixed concentration so that the ratio of [DNA]/ [I ˃ 2σ(I)] 0.0807 = 0.1718 wR2 = 0.0644
[complex] = r (r = 0.2, 0.4, 0.6, 0.8, 1, 1.2, 1.4, 1.6, 1.8, 2, 2.2, 2.4). At R indices (all R1 = 0.0341, wR2 = R1 = 0.1009, wR2 R1 = 0.0260,
each titration step, an equal amount of DNA was added to the control data) 0.0825 = 0.1928 wR2 = 0.0656
Largest 0.840 and − 0.56 1.66 and –1.34 0.47 and − 0.82
and test solutions to remove the DNA adsorption. All absorption spectra
difference in
were recorded after 7 min of incubation at room temperature. Then, the peak and
Wolfe-Shimer formula (Eq. (1)) was used to estimate Kb binding con­ hole (e Å− 3)
stant [33]. In this equation, εb, εf, εa are related to extinction coefficients
of sample completely bound onto DNA, non-bound complex, and in DNA
presence (Aobsd/[Pd] = εa), respectively. (5 µM EB, 100 µM ct-DNA, buffer Tris-HCl with pH 7.2). Fluorescence
intensity changes of the test solutions were recorded after incubation for
[DNA] [DNA] 1 7 min at room temperature from 500 to 700 nm using 471 nm excitation
= + (1)
|(εa − εf )| |(εb − εf )| Kb |(εb − εf )| radiation. The extent of EB-DNA fluorescence quenching by complexes
was calculated using the Stern–Volmer formula (Eq. (3)) [35], where F
Graphical plots of [DNA]/|(εa–εf)| against [DNA] afford Kb values
and F0 reveal fluorescence intensity in quencher (complex) presence and
which can be obtained using the ratio of the slope (1/|(εb–εf)|) to the
absence, respectively, Ksv exhibits linear Stern-Volmer quenching con­
intercept (1/Kb|(εb–εf)|) of the linear plots. Gibbs free energy values for
stant, [Q] represents complex concentration. Ksv value was estimated
DNA binding were obtained using the following equation (Eq. (2)) [34].
using the slope of F0/F vs [Q] linear plot.
ΔG0b = − RT ln Kb (2)
F0 /F = 1 + Ksv[Q] (3)

2.5.2. Competitive binding studies


Since DNA and synthesized complexes do not show fluorescence 2.6. Cytotoxicity activity evaluation
activity in free form and in complex form, to determine the model of
interaction of compounds with ct‑DNA, the ethidium bromide (EB) 2.6.1. Cell culture
fluorophore was used as a fluorescence probe in this spectroscopy MDA-MB-231 triple-negative breast cancer (TNBC) cell line was
technique. EB-DNA competition experiments were performed by adding purchased from the Cell Bank of Pasteur Institute of Iran. MDA-MB-231
different concentrations (20–200 µM) of complexes to EB-DNA solution cells were cultured in DMEM (high glucose) medium supplemented with

3
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

1 % penicillin–streptomycin (GIBCO) as antibiotics and 10 % fetal representation of the interactions of the docked molecules was depic­
bovine serum (FBS) at 37 ◦ C in a humidified incubator containing 5 % ted using the LIGPLOT program (v.1.4.5) [41]. Also, non-covalent in­
CO2. After 2–3 days, all culture media were renewed, considering MDA- teractions between DNA and complexes were investigated through the
MB-231 cellular growth rates. non-covalent interaction (NCI) method [42] using Multiwfn software
[43]. This method, which is known as the reduced density gradient
2.6.2. MTT cytotoxicity assays (RDG) method [44], is one of the popular methods for studying weak
Cytotoxicity activities of Pd(II) complexes towards the MDA-MB-231 interactions, which can differentiate between attractive (such as
cell line (a typical TNBC line) were evaluated in vitro using the MTT π-stacking) and repulsive interactions [43]. Calculating RDG grid data
assay. MTT is a water-soluble tetrazolium salt that is reduced in viable for large systems takes a long time, so NCI analysis is performed based
cells by the enzyme Succinate dehydrogenase in the mitochondrial on promolecular density for large systems [45]. Promolecular density in
membrane to an insoluble purple formazan product. Since formazan the term is the electron density which is constructed almost by super­
absorption intensity has a direct relationship with alive cells’ density, it posing the electron density of the atoms in their free state, known as the
is possible to measure the cytotoxicity potential of the compounds. Promolecular approximation. Best docked poses were used to conduct
MDA-MB-231 cells were cultured into 96-well plates with 7000 cells/ these studies, and the data obtained from the Multiwfn program were
well density using 100 µl DMEM medium including 10 % FBS, which was visualized by VMD (Virtual Molecular Dynamic) software [46].
then incubated for 12 h under air containing 5 % CO2 at 37 ◦ C [36].
Subsequently, culture environment was changed by 100 µl of new so­ 2.8. Computational methods
lution composed of various (2.5, 5, 7.5, 10, 25, 50, 100 µg/mL) com­
plexes concentrations. The drugs were dissolved within DMSO and All quantum-chemical calculations were performed using density
diluted with culturing solution to desired concentrations. Each dosage functional theory (DFT) at the B3LYP (Becke, 3-parameter, Lee­
was poured into wells separately, and this was repeated triple times. –Yang–Parr) level [47,48] by the Gaussian 09 program package [49].
Control was a medium containing 1 % DMSO. After 24 and 48 h incu­ Basis sets of LANL2DZ [50] and 6–311 + G** were utilized to describe
bation, 10 µl of 5 mg/mL MTT dissolved in 90 µl of the medium was Pd atoms and main group elements (such as C, N, O, S, Cl, P, and H
added to each well and incubated for another 4 h. After the desired time, atoms), respectively, in all computations. Also, water solvent effects
the medium and unreacted MTT were discarded and 100 µl DMSO was were investigated in calculations using the self-consistent reaction field
added to each well to dissolve the formed formazan crystals. Finally, (SCRF) method based on the polarizable continuum model (PCM)
absorbances of specimens were recorded at 540 nm by BioTek ELx808 [51,52]. After the geometry optimization of the complexes, frequency
microplate reader. For comparison, the anticancer drug cisplatin was calculations were performed to confirm that the optimized structures are
used as a reference drug. IC50 values were calculated to evaluate the true minima. The absence of imaginary vibrational frequencies indi­
cytotoxicity effects of the Pd(II) complexes using logarithmic plots of cated that the optimized structures correspond to the energy minima.
viability percentage against drug concentration, at which IC50 is the Natural Bond Orbital (NBO) [53] calculations were also done to extract
concentration that causes 50 % of cell death. The following equation the natural atomic charges and analyze the frontier molecular orbitals.
determined the cellular viability percentage. The Chemcraft program (version 1.8) [54] was applied to draw the
frontier molecular orbitals. Also, using the energies of HOMO and LUMO
mean OD of the treated cells
Cellular viability percentage = orbitals, global reactivity descriptors were calculated by the methods
mean OD of the untreated cells (control)
reported in [55]. Studies of electron density distribution, atomic con­
× 100 tributions of HOMO and LUMO orbitals, and molecular electrostatic
potential (MEP) were completed using Multiwfn software (version 3.7)
2.6.3. Statistical study [43], known as the multifunctional wave function analysis program. The
Results were extracted as mean ± standard deviation of 3 individual outputs were visualized by the Visual Molecular Dynamics (VMD) pro­
experimental data. Using one-way ANOVA, statistical analysis was gram [46]. All wave function files were generated by Gaussian 09
performed utilizing GraphPad Prism 5 program. The P-value < 0.05 package. Fukui functions analysis was employed to determine electro­
represented a significant value. philic and nucleophilic sites. Single-point calculations of palladium
complexes in N, N-1, and N + 1 states (N = number of electrons in the
2.7. Protocol of molecular docking stable situation)were conducted at the same level of theory used for
optimization.
Molecular docking simulations were performed using the AutoDock
4.2 program [37]. The structures of all three complexes were optimized 3. Results and discussion
as described in the Computational Methods section. The crystal structure
of the B-DNA dodecamer with the sequence d(CGCGAATTCGCG)2 was 3.1. Synthesis
extracted from the Protein Data Bank (PDB ID: 1BNA) [38]. All water
and ligand molecules were removed from the macromolecule before the Hydrazone ligand DPPB-HBH was synthesized through a condensa­
molecular docking simulation to prepare the receptor. Then, Gasteiger tion reaction between 2-(diphenylphosphino) benzaldehyde with 3-
charges were calculated for each atom, and non-polar hydrogens were hydroxybenzohydrazide. Ligands DPPB-PTSC and DPPB-TSC were ob­
merged into carbon atoms. In all dockings, the grid box covered the tained according to published procedures [15,56]. Complexes [PdCl
entire macromolecule with a grid spacing of 0.375 Å. All calculations (DPPB-HBH)]Cl, [Pd(DPPB-HBH)2]Cl2⋅2CH3CN, [PdCl(DPPB-PTSC)],
were done by the Lamarckian genetic algorithm method in 100 runs and [PdCl(DPPB-TSC)] were prepared via the substitution reaction of
[39]. Other parameters were set based on the software’s default settings. PdCl2(COD) (COD = 1,5-cyclooctadiene) with ligands DPPB-HBH,
Results that differ by less than 0.5 Å in positional root–mean–square DPPB-PTSC, and DPPB-TSC. From the reaction of DPPB-HBH with
deviation (RMSD) were clustered together. The optimized clusters were PdCl2(COD), a mixture of [PdCl(DPPB-HBH)]Cl and [Pd(DPPB-HBH)2]
ranked based on the lowest binding energy. Finally, the conformation Cl2⋅2CH3CN complexes was obtained. Of these two molecules, only
with the lowest free energy of binding and the highest probability was [PdCl(DPPB-HBH)]Cl was used to evaluate biological activities due to its
selected as the desired conformation for the investigation of in­ dominant percentage and similar coordination model with [PdCl(DPPB-
teractions. The analysis of the resulting structural files and graphical PTSC)] and [PdCl(DPPB-TSC)] compounds. In complexes [PdCl(DPPB-
visualization of their surfaces were performed by AutoDockTools and PTSC)] and [PdCl(DPPB-TSC)], the coordination of the ligands to
Discovery Studio Visualizer [40]. In addition, the two-dimensional palladium cation is in a tridentate form through the P, N, and S atoms.

4
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

While the DPPB-HBH ligand acts as a P, N bidentate and P, N, O tri­ delocalization of electron density among the C(S)-N-N– – C fragment,
dentate donor ligand in the [Pd(DPPB-HBH)2]Cl2⋅2CH3CN and [PdCl which leads to partial double bond nature for C–S bonds. It was inferred
(DPPB-HBH)]Cl complexes, respectively, see Scheme 1. These com­ earlier from the lack of ν(S–H) vibrations in the range of 2500–2600
plexes are all stable to air and moisture, and their diverse structural cm− 1 but ν(N–H) stretching occurrence within 3049–3145 cm− 1 [62].
characteristics are investigated with X-ray crystallography analysis and Also, UV–vis spectra of Pd(II) samples display-three absorption peaks
UV–vis, FT-IR, 31P-NMR, and 1H-NMR, spectroscopy techniques. within regions 202–220, 251–298, and 390–415 nm, as shown in Fig. S1.
The appeared bands around 202–220 and 258–298 nm correspond to
3.2. Spectroscopic studies intra-ligand transitions (π-π* plus n-π* transitions of the aromatic rings
and C– – N groups) [63]. The 390–415 nm bands are assigned to charge

Comparing FT-IR spectra of complexes and corresponding ligands transfer (CT) transitions [62,64]. The UV–vis spectra of ligands and Pd
provided considerable evidence regarding ligand coordination mode to (II) complexes and their data are given in Figs. S1, S2, and Table S1.
1
central metal. Table 2 summarizes FT-IR and NMR spectroscopy results. H and 31P-NMR spectra of complexes and their corresponding li­
There are three characteristic absorption bands in the ranges of gands are recorded in DMSO‑d6. The 31P NMR spectra of the ligands
3145–3049, 1582–1490, and 1437–1394 cm− 1 within FT-IR spectra of show signals in the − 16.65 to − 14.17 ppm range due to the phosphorus
Pd(II) complex samples, which are assigned to stretching frequencies of atoms resonance of triphenylphosphine. The peaks are firmly moved
N–H, C– – N, C–P groups, respectively [57]. The shifts of the C–
– N and downfield upon complexation, indicating phosphorus atoms are
C P bands by 6–64 cm
– − 1
in the direction of lower wavenumbers involved in coordination. 1H NMR spectra of ligands exhibit signals
compared to those of the corresponding ligands confirm the participa­ belonging to the hydrogen atom of N-NH-C = X (X = O or S) group at δ
tion of azomethine nitrogen atoms and triphenylphosphine phosphorus 11.51–11.93 ppm [65], but these signals appear in the regions of δ
to coordinate metal cation [21]. The stretching vibration of the C– –O 8.67–9.16 ppm in the spectra of complexes. These shifts support the
bond in the [PdCl(DPPB-HBH)]Cl complex appears at 1648 cm− 1 [58]. It coordination of sulfur and oxygen atoms in corresponding ligands onto
is moved to a higher frequency than the free ligand. This shift proves central metal ions [66]. For each Pd(II) complex, a doublet appears near
that the oxygen atom has been coordinated as carbonyl (i.e., the 8.36–8.45 ppm due to the coupling of the azomethine proton to the
hydrazonic ligand is connected in the keto-form) [59]. phosphorus atom of triphenylphosphine, which is transferred to more
The characteristic bands of C(S) appear at around 1033–1093 cm− 1 shielded regions compared to those of the free ligands (9.22–8.64 ppm).
and 1098–1154 cm− 1 in thiosemicarbazone complexes. These bands Thus, this result confirms that the azomethine N atom has participated
show the values between stretching modes of single bond C–S (ν(C–S)) in complexation [59]. Furthermore, in the spectra of the complexes,
and stretching modes of C– – S double bond (ν(C–S)) [60,61] due to the there are a series of overlapping multiplets in the range of 6.04–7.97

Scheme 1. Chemical structures and synthesis process of Pd(II) complexes.

5
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

Table 2
Some spectroscopic data of Pd(II) complexes.
Name X ν(N–H) ν(C–
–O) ν(C–
–N) ν(C–P) ν(CS) ν(C–H)arom δ(N–H) δ(CH = N) Ref.

DPPB-HBH O 3058w, br 1645s 1588s 1443m – 2971w 11.58 9.22 –


[PdCl(DPPB-HBH)]Cl O 3059 w, br 1648s 1582s 1437m – 2929w 9.08 8.45 –
DPPB-PTSC S 3135w, br – 1541s 1438m 1072w 2990w 11.93 8.77 [56]
[PdCl(DPPB-PTSC)] S 3049w, br – 1490s 1430m 1093w, 1033w 2924w 8.67 8.46 –
DPPB-TSC S 3166w, br – 1534s 1436m 1090w 3016w 11.51 8.64 [15]
[PdCl(DPPB-TSC)] S 3145w, br – 1470s 1435m 1098w, 1154w 2925w 9.16 8.36 [15]

ppm assigned to H atoms in aromatic rings of ligands. lies between the C–S single bond (d = 1.82 Å) and the C– – S double bond
(d = 1.56 Å) in the [PdCl(DPPB-PTSC)] crystal structure [67], which
indicates the CS bond has a partial double bond character. Moreover, the
3.3. Crystal structure analysis of Pd(II) complexes
C–N bond (d = 1.299(3) Å) in this structure is shorter compared to the
classical C–N bond (d = 1.471 Å) [68]. All these shreds of evidence
To confirm the coordination structure of the Pd(II) complexes, suit­
confirm the delocalization of electron density in the thiosemicarbazone
able single crystals of [PdCl(DPPB-HBH)]Cl, [Pd(DPPB-HBH)2]
chain. In the crystal structure of [PdCl(DPPB-HBH)]Cl, the C– – O dis­
Cl2•2CH3CN and [PdCl(DPPB-PTSC)] were obtained by slow evapora­
tance (d = 1.257(10) and 1.262(9) Å) is close to the classical C– – O bond
tion of their acetonitrile solution. Two different types of crystals were
(d = 1.264 Å) [68], indicating the iminophosphine ligand is coordinated
obtained from the reaction of the DPPB-HBH ligand with PdCl2(COD),
in the keto form.
including thin yellow crystals ([PdCl(DPPB-HBH)]Cl) and cubic red
The [PdCl(DPPB-HBH)]Cl crystal structural analysis displays that the
crystals ([Pd(DPPB-HBH)2]Cl2•2CH3CN) (Fig. 1). The type of coordi­
structure contains two molecules and two chloride ions in its asym­
nation structure of every single crystal was determined by X-ray crys­
metric unit, as shown in Fig. 2. The DPPB-HBH tridentate ligand coor­
tallography. Fig. 2 and Table 1 illustrate the ORTEP plots and the
dination to central cation forms one five- plus one six-membered chelate
crystallographic information of these complexes, respectively. Also,
ring (Cg1 = Pd1-O1-C8-N2-N1, Cg7 = Pd2-O3-C34-N4-N3, Cg2 = Pd1-
Table S2 lists relevant bond lengths and bond angles. Crystallographic
P1-C2-C1-C7-N1, and Cg8 = Pd2-P2-C28-C27-C33-N3) around the Pd
data show that [PdCl(DPPB-PTSC)] and [Pd(DPPB-HBH)2]Cl2•2CH3CN
(II) centers. In the crystal structure of [PdCl(DPPB-HBH)]Cl, molecules
crystals are crystallized within monoclinic crystal systems with P21/c
A and B are connected one by one through N2-H2A⋯Cl4, O4-H4A⋯Cl4,
and C2/c space groups, whereas single crystals of [PdCl(DPPB-HBH)]Cl
O2-H2⋯Cl3, and N4-H4B⋯Cl3 hydrogen bonds forming zigzag chains
belong to the triclinic system and P-1 space group.
along the a-axis (Table S3). Adjacent zigzag chains are connected
In all structures, the central metal reveals distorted square-planar
through CH⋯π interactions (C43-H43⋯Cg4 and C30-H30⋯Cg10) and
coordination geometry. The three coordination positions of the [PdCl
create two-dimensional layers in the ab plane (Fig. 3A and Table S4).
(DPPB-HBH)]Cl and [PdCl(DPPB-PTSC)] complexes are coordinated
Complex [Pd(DPPB-HBH)2]Cl2•2CH3CN contains half of the mole­
by the PNO and PNS donor atoms of the corresponding ligands,
cule, acetonitrile solvent, and a chloride ion in its asymmetric unit, and
respectively, and the chloride ion occupies the fourth position. In [Pd
the other half, together with a solvent molecule and another chloride
(DPPB-HBH)2]Cl2•2CH3CN, four coordination positions are occupied by
ion, are generated by the inversion center. In this complex, the two
the P and N atoms of two iminophosphine ligands (DPPB-HBH), Fig. 2.
DPPB-HBH ligands are attached to the palladium atom as bidentate li­
According to Table S2, the cis bonding angles around Pd are greater or
gands through the phosphorus atoms of triphenylphosphine and the
less than 90◦ , and both trans ones are tinier than 180◦ . Hence, these
iminic nitrogen atoms, which form 2 six-member rings with Pd(II) cation
displacements confirm the deviation of the central metal ions from the
(Fig. 2). In this crystal structure, chloride ions outside the coordination
ideal square-planar geometry. The C–S bond distance (d = 1.753(2) Å)
sphere generate zigzag chains along the c-axis by creating N2-H2A⋯Cl1
and O2-H2⋯Cl1 hydrogen bridges between adjacent molecules (Fig. 3B
and Table S3).
The chains are joined together via CH⋯π connections between the
hydrogens of the phenyl rings of one chain of triphenylphosphine
fragment (H18) and the 4‑hydroxyphenyl rings of another chain. In
addition, other C–H⋯π intermolecular interactions such as C3-
H3⋯Cg6 and π⋯π interactions (Cg4⋯Cg4) between molecules of a
chain stabilize the crystal structure of [Pd(DPPB-HBH)2]Cl2•2CH3CN in
the bc-plane; see Fig. 3B and Table S4.
[PdCl(DPPB-PTSC)] crystal contains one molecule in the asymmetric
unit, and each molecule links to its neighboring molecule through C25-
H25⋯Cl1 (dH25⋯Cl1 = 2.93(10) Å) non-classical hydrogen bonding
interactions; hence, infinite one-dimensional polymeric chains are
extended along the a-axis. The one dimensional chains are inter­
connected together through C17-H17⋯N2, C12-H12⋯Cl1, C6-H6⋯Cl1,
and C19-H19⋯S1 contacts (dH17⋯N2 = 2.61(2), dH12⋯Cl1 = 2.79(8),
dH6⋯Cl1 = 2.82(9), and dH19⋯S1 = 2.84(9) Å) and weak π⋯π
stacking interactions (C23-H23⋯Cg4) to form the 2D network along ab-
plane, see Fig. 3C, Tables S3 and S4.

3.4. DNA binding studies

Fig. 1. Image indicating a mixture of [PdCl(DPPB-HBH)]Cl and [Pd(DPPB- 3.4.1. Electronic absorption titration method
HBH)2]Cl2•2CH3CN crystals at 50X magnification. Before starting the experiment, to measure the stability of the

6
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

Fig. 2. ORTEP diagrams of complexes [Pd(DPPB-HBH)2]Cl2•2CH3CN, [PdCl(DPPB-HBH)]Cl, and [PdCl(DPPB-PTSC)] (The thermal ellipsoids have been shown by
50% probability).

complexes, their absorption spectrum was measured in Tris-HCl buffer obtained to be − 24.66, − 19.84, and − 22.56 KJ.mol− 1 at 25 ◦ C for
solution containing 1 % DMSO after 24 h. The slight changes observed in compounds [PdCl(DPPB-HBH)]Cl, [PdCl(DPPB-PTSC)], and [PdCl
the absorption spectrum of the complexes prove their stability in the (DPPB-TSC)], respectively, indicating the spontaneous interactions of
Tris-HCl buffer environment (Fig. S3). the complexes with DNA [34].
Since DNA is one of the main targets of anticancer medications, the
interaction study of Pd(II) complexes with DNA and the model of their 3.4.2. Competitive binding using ethidium bromide (EB)
binding to each other provide valuable information. Consequently, the To further elucidate the DNA binding modes of Pd(II) complex
interactions between the DNA and synthesized complexes are evaluated samples, the EB displacing technique was performed by monitoring the
through the electron absorption titration technique by monitoring the variations of EB-DNA fluorescence intensity. Notably, the EB fluores­
absorbance variations. Fig. 4 depicts changes in absorption spectra for cence probe shows high emission intensity upon coupling with DNA,
the [PdCl(DPPB-HBH)]Cl, [PdCl(DPPB-PTSC)] and [PdCl(DPPB-TSC)] which changes with adding another DNA-binding compound [79].
samples in the presence and absence of various ct-DNA dosages. Elec­ Emission reductions may occur when the compound can replace EB in
tronic spectra of the free complexes illustrate three absorptions within the DNA strand, or it can accept excited electrons of the EB intercalator
ranges of 202–220, 251–298 (intra-ligand transitions), and 390–415 nm [80]. The fluorescence spectra of the EB-DNA solution are taken with the
(CT transitions). By adding different ct-DNA doses, absorption bands presence and absence of diverse palladium complexes concentrations.
related to intra-ligand transitions exhibit hyperchromism and blue shift Investigation of the effects of the complexes on the photo­
(Fig. 4). These spectral changes are not significant for the CT bands. The luminescence intensity shows that with increasing concentrations of
observed hyperchromic effects with blue-shift suggest effective in­ [PdCl(DPPB-HBH)]Cl, [PdCl(DPPB-PTSC)], and [PdCl(DPPB-TSC)] the
teractions of Pd(II) complexes with DNA helix structure, which is likely emission intensity gradually decreases. The observed quenching in­
to be due to the weak interactions with nucleobases in minor and major dicates the interactions of palladium complexes with DNA helix strands,
grooves of DNA or electrostatic interactions with the phosphate back­ as seen in Fig. 5. Quenching constant values (Ksv) as a measure of the
bone of the DNA helix [67,69,70]. The intrinsic binding constants (Kb) tendency of Pd(II) complexes to bind to DNA is calculated as 0.70 × 103,
between the complexes and ct-DNA are calculated from the ratio of slope 0.349 × 103, and 0.60 × 103 for [PdCl(DPPB-HBH)]Cl, [PdCl(DPPB-
to the Y-intercept of the linear plots of [DNA]/|(εa-εf)| vs [DNA] (Fig. 4). PTSC)], and [PdCl(DPPB-TSC)], respectively, using the Stern–Volmer
Kb values were found to be 2.1 × 104, 0.3 × 104, and 0.9 × 104 M− 1 for equation. The DNA-binding ability follows the order of [PdCl(DPPB-
[PdCl(DPPB-HBH)]Cl, [PdCl(DPPB-PTSC)], and [PdCl(DPPB-TSC)] HBH)]Cl˃ [PdCl(DPPB-TSC)]˃ [PdCl(DPPB-PTSC)]. These observations
complexes, respectively, which indicates that all three complexes have confirm the results obtained by absorption titration studies (Table 3).
good binding affinity to ct-DNA, and [PdCl(DPPB-HBH)]Cl interacts Based on the Ksv values and the information extracted from the experi­
more strongly with DNA than the other complexes. ments, the quenching can be considered a result of the acceptance of
Compared to the classical intercalator ethidium bromide (Kb = 1.23 excited electrons during the photoelectron phenomenon. [PdCl(DPPB-
(±0.07) × 105 M− 1) [71] and metallointercalators whose binding con­ HBH)]Cl creates the most robust binding to the DNA, among other
stants have been reported in the range of 105 to 107 M− 1 [72–76], Pd (II) structures. It is most likely associated with hydroxyl and carbonyl
complexes exhibited a lower Kb value. These results indicate that these moieties in the [PdCl(DPPB-HBH)]Cl structure.
complexes can interact with DNA moderately and with less strength than
intercalators [77,78]. Values of intrinsic binding constants are presented 3.5. Cytotoxicity activities
in Table 3. Standard Gibbs free energy (ΔG0b ) for DNA binding is
Following outstanding data attained by DNA binding studies, the

7
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

Fig. 3. (A) Crystal packing of [PdCl(DPPB-HBH)]Cl; 2D representation of intermolecular interactions in the ab-plane, showing the molecules are connected through
hydrogen bonding and C–H⋅⋅⋅π interactions along the a-axis and in the direction of the ab-plane. (B) Side view of [Pd(DPPB-HBH)2]Cl2•2CH3CN crystal packing in
the bc-plane. (C) Representation of the 2D layer in [PdCl(DPPB-PTSC)] shows how chains are connected in the ab-plane. In all shapes, non-interacting hydrogens were
removed for simplicity.

cytotoxicity effects of the Pd(II) complexes were evaluated on the pro­ the tendency of molecules to bind to DNA. This agreement suggests that
liferation of MDA-MB-231 cells using the MTT method. For this purpose, the cytotoxic activities of Pd(II) complexes are possibly associated with
various Pd(II) complexes concentrations (100, 50, 25, 10, 7.5, 5.0, 2.5 interactions by DNA.
µg/mL) were added to these breast cancer cells at 24- and 48-h intervals.
IC50 values of tested samples are presented in Table 4, where the low
3.6. Molecular docking studies
values of IC50 indicate their performance at low concentrations. Analysis
of the results reveals that all three Pd(II) complexes substantially inhibit
According to the results of experimental studies and the desired
the growth of cells in a time- and dose-dependent manner, meaning with
abilities of Pd(II) complexes to bind to DNA, to obtain an overview of the
increasing concentration and treatment time, the mean cell viability
interaction modes of complexes with DNA, molecular docking simula­
decreases. As a result, they have a positive effect on inhibiting cell
tion was conducted, which is a powerful graphical tool for better un­
proliferation (Fig. 6).
derstanding non-covalent interactions between drug and
Results also verify that all three complexes have higher cytotoxic
macromolecule.
activities than the cisplatin reference drug (IC50 = 24.59 µg/mL) upon
The geometries of all three molecular structures were fully optimized
incubation for 24 h. This increasing trend of inhibitory potency of
in the water phase by density functional theory (DFT) with B3LYP
compounds continues to be observed with the prolongation of the
method using LANL2DZ and 6–311 + G** basis sets for palladium and
treatment time. However, only [PdCl(DPPB-HBH)]Cl shows higher in­
main group elements, respectively. Also, DNA crystal structure (1BNA.
hibition than the cisplatin (IC50 = 6.24 µg/mL) after incubation for 48 h.
pdb) was obtained from the Protein Data Bank.
According to these interpretations, [PdCl(DPPB-HBH)]Cl has the most
The results show that all the complexes interact well with the DNA
potent anti-proliferative activity among other compounds tested as it
and bind to nucleotide residues in the minor groove of the DNA double
shows IC50 amounts of 9.3 and 4.99 µg/mL after 24 and 48 h, respec­
helix by H-bonds, electrostatic and hydrophobic interactions. Among
tively, against selected cancer cells. It could correspond to the mole­
the docked molecules, [PdCl(DPPB-HBH)]Cl demonstrates the highest
cule’s hydroxyl and carbonyl functional moieties, making it more
binding affinity with a free binding energy of − 8.88 kcal/mol (with Ki
reactive than the two complexes containing C(S) group. In general, the
= 0.310 µM) compared to the other two molecules, which is in good
anticancer potentials of these samples change as [PdCl(DPPB-HBH)]Cl
agreement with the experimental results (Table S5). The most probable
> [PdCl(DPPB-TSC)] > [PdCl(DPPB-PTSC)], which is consistent with
and best binding positions of the complexes in the DNA double-helix are

8
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

Fig. 4. Absorption titration spectra of complexes [PdCl(DPPB-HBH)]Cl, [PdCl(DPPB-PTSC)], and [PdCl(DPPB-TSC)] (48 µM) within Tris-HCl buffer upon both the
presence and absence of different DNA concentrations (9.94–170 µM). Arrows display absorption intensity changes by varying DNA concentrations.

Table 3
Electronic absorption spectroscopic data plus DNA binding parameters of sam­
ples [PdCl(DPPB-HBH)]Cl, [PdCl(DPPB-PTSC)], and [PdCl(DPPB-TSC)].
Compound λmax (nm) Δλ Kb Ksv ΔG* (KJ.
(nm) (M− 1) (M− 1) mol− 1)
Free Bound

[PdCl(DPPB- 298 258 40 2.1 × 0.70 × –24.66


HBH)]Cl 104 103
[PdCl(DPPB- 251 259 8 0.3 × 0.349 × –19.84
PTSC)] 104 103
[PdCl(DPPB- 290 265 25 0.9 × 0.60 × –22.56
TSC)] 104 103

depicted in Fig. 7.
As shown in Fig. 8, the orientation of [PdCl(DPPB-HBH)]Cl is such
that it fits precisely in parallel between the DNA strands and involves
more residues during the interaction process, thus creating more in­
teractions with the target DNA functional groups. In complexes [PdCl
(DPPB-PTSC)] and [PdCl(DPPB-TSC)], on the other hand, the orienta­
tion is such that part of both molecules is located outside the minor
groove and naturally provides fewer sites for binding (Figs. 7 and 8).
Hence, this evidence verifies the importance of the presence of hydroxyl
and carbonyl groups in increasing the binding potential of [PdCl(DPPB-
HBH)]Cl. Eight interactions stabilize [PdCl(DPPB-HBH)]Cl in the DNA
chains. The two classical hydrogen bonds are composed of the interac­
tion of a hydroxyl oxygen atom with the oxygen atoms of residues DT7
Fig. 5. The EB-DNA emission spectra both in the presence and absence of
and DT19 in chains A and B, respectively (with bond lengths of 2.18 and
various complex concentrations ([PdCl(DPPB-HBH)]Cl, [PdCl(DPPB-PTSC)],
2.77 Å). Also, two non-classical hydrogen bonds are formed by con­ and [PdCl(DPPB-TSC)]). [DNA] = 100 µM, [EB] = 5 µM, [Complex] =
necting oxygen atoms in hydroxyl moiety to iminic hydrogen with res­ 20–200 µM. Arrows display the emission intensity changes.
idues DA18:C2 and DC9:O2 (Fig. 7).

9
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

Table 4 complexes, NCI analysis based on promolecular density was performed


IC50 values obtained from the MTT assay for investigated Pd(II) complexes and using VMD 1.9.3 and Multiwfn 3.7 software.
cisplatin against MDA-MB-231 cells after 24 and 48 h. The best binding mode obtained from molecular docking simulations
Tumor Time IC50 (µg/mL) was used for the NCI studies. Different interactions can be identified in
cell (h)
[PdCl(DPPB- [PdCl(DPPB- [PdCl Cisplatin
various regions according to the obtained color-filled NCI gradient iso­
HBH)]Cl PTSC)] (DPPB- surfaces and considering the scale bar [44,45,81]. As can be seen from
TSC)] Figs. 8 and S5, each interaction is shown in a different color. The red
MDA-MB- 24 9.3 17.47 15.35 24.59 color is due to the steric repulsion in the phenyl and phenol rings of the
231 48 4.99 11.45 9.87 6.24 complexes, and the green and blue colors are due to the presence of van
der Waals and H-bonds contacts among the DNA functional groups and
molecules, respectively. It is clear from the results that in all investigated
On the other hand, DG10 and DC11 residues participate in both complexes, a large surface of NCI gradient isosurface is related to van
electrostatic and hydrophobic interactions with [PdCl(DPPB-HBH)]Cl, der Waals interactions (such as π-stacking) [45,67,81,82], which is in
and in addition, hydrophobic interactions are observed with DA18, complete alignment with the results of docking simulation and experi­
DT19, DC9, DT8, DA17, and DG16 (Fig. S4). Information on bond mental studies.
length, type of interaction, and participating residues are presented in
Table 5. In [PdCl(DPPB-PTSC)] and [PdCl(DPPB-TSC)] complexes, hy­
drophobic intermolecular forces are the leading forces in the in­ 3.7. Computational studies
teractions with DNA double-helix, and fewer residues are involved
within electrostatic and H-bonds contacts. 3.7.1. Quantum chemistry descriptors
To better visualize non-covalent interactions (NCI) such as steric Studies of compounds’ structural and electronic properties provide
effect, van der Waals interactions, and hydrogen bonds in drug-DNA deep insights into the influences of functional groups on reactivity.
Therefore, to make a correlation between the structural properties and

Fig. 6. Cell viability of cells treated by various samples concentrations for 24 h (A) and 48 h (B).

10
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

Table 5
Details of the interactions between the docked Pd(II) complexes and the DNA
double helix.
Compound Nucleobase Interaction type Distance
residues (Å)

[PdCl(DPPB- A: DT7:O2 Hydrogen Bond 2.18


HBH)]Cl B: DT19:O2 Hydrogen Bond 2.77
B: DA18:C2 Hydrogen Bond (non- 2.72
classical)
A: DC9:O2 Hydrogen Bond (non- 3.33
classical)
A: DG10:OP1 Electrostatic 4.20
A: DC11:OP1 Electrostatic 3.25
A: DC11:OP1 Electrostatic 5.45
A: DG10:C5′ Hydrophobic π-alkyl 3.82
[PdCl(DPPB- B: DG22:C5′ Hydrogen Bond (non- 3.59
PTSC)] classical)
B: DG22:OP1 Electrostatic 4.09
B: DG22:OP1 Electrostatic 5.47
A: DG4 Hydrophobic π-π T- 4.90
shaped
[PdCl(DPPB- B: DC23:OP1 Hydrogen Bond 2.06
TSC)] B: DG22:OP1 Electrostatic 5.09
B: DC23:OP1 Hydrogen Bond; 2.13
Electrostatic

Fig. 7. Demonstration of the possible interactions between investigated com­


plexes and DNA; The various interactions are distinguished by colored dashes in the activity of the Pd(II) complexes, the electron density distribution, the
which the green and gray colors represent the classical and non-classical atomic contribution of HOMO and LUMO orbitals, and other quantum
hydrogen bonds, respectively, and the orange color illustrates the electro­ chemical parameters affecting the biological activity like electron af­
static interactions. (For interpretation of the references to color in this figure finity, ionization potential, softness, hardness, chemical potential,
legend, the reader is referred to the web version of this article.) electrophilicity index, and electronegativity, were selected and
computed by respective standard equations [83]. The DFT optimization
of geometrical structures was performed at the B3LYP level using a
6–311 + G** basis set for O, N, C, H, S, and Cl atoms while the LANL2DZ
basis set for Pd atom within the solvent phase (water).
The reactivity of compounds is due to the charge transfer processes
between electron acceptor and donor groups. It is reasonable to analyze
the compounds’ reactivity differences based on their LUMO and HOMO
energies. Diagrams of frontier molecular orbitals (FMOs) of all three
molecules with the energy level values and energy differences between
them are depicted in Fig. 9. Analysis of FMOs displays that, in all
complexes, the HOMO orbitals are mainly located on the palladium and
the atoms coordinated to it, and the LUMO orbitals are spread over the
hydrazide fragment and phenyl ring attached to the C– – N bond of the
ligands.
Frontier orbitals composition was carried out by the Hirshfeld
method using Multiwfn software. Fig. S6 exhibits the percentage con­
tributions of HOMO and LUMO. As can be seen, the atomic contributions
of the HOMO orbitals in [PdCl(DPPB-PTSC)] and [PdCl(DPPB-TSC)] are
such that the sulfur atoms participate to a greater extent (9.55 %), while
in the [PdCl(DPPB-HBH)]Cl, the oxygen atom contributes less to this
orbital (2.99 %). This situation is reversed for LUMOs, where 2.86 % is
localized on the sulfur atom and 4.29 % on the O atom.
Fig. 10 displays the electron density maps of the molecules. As
evident from Fig. 10 and the values of atomic charges presented in
Table S7, there are more electronic densities around the O atoms of
[PdCl(DPPB-HBH)]Cl than sulfur atoms of [PdCl(DPPB-PTSC)] and
[PdCl(DPPB-TSC)] molecules, which illuminate their abilities to interact
and participate in nucleophilic reactions. Table S6 lists the data related
to the quantum reactivity parameters. Energy gap values provide in­
formation about the chemical reactivity and kinetic stability of com­
pounds that are very close to each other for the complexes tested, and
Fig. 8. NCI plots for the most probable binding modes of [PdCl(DPPB-HBH)]Cl
they are consistent with the energy gap values of biologically active
(A), [PdCl(DPPB-PTSC)] (B), and [PdCl(DPPB-TSC)] (C) complexes to DNA. molecules. Other chemical reactivity parameters were calculated using
Surfaces represent various non-covalent interactions based on the color bar. LUMO and HOMO levels. The results illustrate that the electron transfer
process in terms of energy in the studied complexes is favorable due to
the positive values of chemical hardness (η) and the negative values of
energy change (ΔE) [84].

11
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

Fig. 9. Schematic representation of the frontier molecular orbitals of [PdCl(DPPB-HBH)]Cl, [PdCl(DPPB-PTSC)], and [PdCl(DPPB-TSC)] with the energy gap values.

The electrophilicity index (ω) is the most crucial parameter in bio­ suitable for electrophilic attack. The blue, green, and yellow regions are
logical activity studies. The samples with higher ω have more in­ electron deficient, neutral, and less rich in electrons, respectively.
teractions with biological molecules. According to the ω values, all three As can be seen in Fig. 11, the areas around the Pd atoms have the
complexes have high electrophilicity and are considered the strongest in most negative potential, and the most positive one is related to the
the classification of the electrophilic index [85]. attached hydrogen to the hydrazonic nitrogen, which is the most
The order of increasing electrophilicity of the complexes is [PdCl desirable site for the nucleophilic attack. Localized surface maxima and
(DPPB-HBH)]Cl > [PdCl(DPPB-PTSC)] > [PdCl(DPPB-TSC)]. It suggests minima in ESPs have been exhibited, respectively, with red and blue
that there may be more interactions between [PdCl(DPPB-HBH)]Cl and spheres. Comparison between the oxygen and sulfur atoms in the
macromolecules. structures of the complexes reveals that the electrostatic potential of the
Electronegativity (χ ) demonstrates the electron capture capability of oxygen is more negative (the red region) than that of the sulfur atom
the compound. High or low electronegativity values determine the (the yellow region), which makes it prone to participate in nucleophilic
complexes’ Lewis acidic or basic properties. Samples with higher χ reactions. Since the dominant color of the electrostatic potential in Pd
participate in the reaction as a Lewis acid. The term -χ defines the (II) complexes is blue, they can act as electrophilic agents interacting
chemical potential of a compound. A more negative chemical potential with macromolecules. The highlighted reactivity sites are clear evidence
in a combination shows that it is more difficult to lose the electron, of the reactivity of the complexes.
whereas it is easier to accept electrons. The given high values of elec­ The performance of Fukui calculations of atoms helps to understand
trophilicity and electronegativity and the negative chemical potential the interactions better. According to the Mulliken population, Fukui
are considerable. It concludes that these complexes tend to accept functions provide quantitative data about atomic electrophilicity and
electrons and act as electrophiles in the reaction with macromolecules. nucleophilicity for a compound [87–89]. The following formula esti­
The order of increasing electronegativity is [PdCl(DPPB-HBH)]Cl > mates these functions.
[PdCl(DPPB-PTSC)] > [PdCl(DPPB-TSC)], indicating that [PdCl(DPPB-
f + (→
r ) = qr(N + 1) − qr(N) The nucleophilic attack
HBH)]Cl has a greater affinity for electron acceptance among the
other two complexes. This result entirely approves those attained from
f − (→
r ) = qr(N) − qr(N − 1)The electrophilic attack
the experimental analyses.
f 0(→
r ) = (qr(N + 1) − qr(N − 1) )/2 The radical attack
3.7.2. Reactive sites of molecules
Fukui functions and molecular electrostatic potential (MEP) were In Fukui functions equations, qr is the calculated atomic charge from
investigated to assign reactive sites in molecules. MEP is a computa­ analyzing the Mulliken population at the rth atomic site in neutral (N),
tional method highlighting a molecule’s electrophilic or nucleophilic anionic (N + 1), and cationic (N-1) chemical species. The 0, − , and +
areas based on electron density [86]. The MEP schemes of the complexes signs indicate radical, electrophilic, and nucleophilic attacks, respec­
are plotted using the DFT approach (6–311 + G**/LANL2DZ basis sets), tively. Full details of Fukui calculations for all atoms of three molecules
in which different colors indicate various values of electrostatic poten­ are presented in Tables S8–S10 (see the Supplementary data). The dif­
tials. High electron density positions are displayed in red and are ference between the nucleophilic and electrophilic Fukui functions is

12
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

Fig. 10. Color-filled maps of the electron density of molecules.

defined as the dual descriptor Δf(r) [90]. structures of complexes [PdCl(DPPB-HBH)]Cl, [Pd(DPPB-HBH)2]
Cl2•2CH3CN, and [PdCl(DPPB-PTSC)] were determined by X-ray crys­
Δf (r) = f + (→
r ) − f − (→
r ) Dual descriptor
tallography. Analysis of crystal structures confirmed the distorted
When Δf(r) < 0, the site is suitable for the electrophilic attack, and square-planar geometry for the palladium centers, it was also found that
when Δf(r) > 0, the location is fit for nucleophilic attack. In general, the DPPB-HBH ligand could form two different palladium complexes,
according to the dual descriptor data of atoms, palladium atoms and acting as a P, N bidentate and P, N, O tridentate in the [Pd(DPPB-HBH)2]
iminic carbons can act as nucleophile sites in all three complexes. In Cl2•2CH3CN and [PdCl(DPPB-HBH)]Cl complexes, while the DPPB-
[PdCl(DPPB-HBH)]Cl, the oxygen atom of the carbonyl and hydroxyl PTSC ligand behaves as a tridentate ligand through the P, N, and S
groups and the hydrazonic nitrogen atom can be a good place for elec­ donor atoms in the [PdCl(DPPB-PTSC)]. Among the synthesized com­
trophilic attack, while H9 is more suitable for nucleophilic attack due to pounds, complexes [PdCl(DPPB-HBH)]Cl, [PdCl(DPPB-PTSC)], and
its binding to the electronegative nitrogen atom (N8). In [PdCl(DPPB- [PdCl(DPPB-TSC)] were employed to evaluate the biological activities
PTSC)] and [PdCl(DPPB-TSC)] complexes, hydrazonic nitrogen and due to their similar coordination configuration. The results of the DNA-
sulfur atoms are suitable sites for nucleophilic attack due to the partial binding studies showed that all three complexes have good binding af­
double-bond character of the C–S bond and the conjugation effect of finity to ct-DNA through electrostatic and groove binding modes. Also,
double bonds. the cytotoxicity assay of palladium(II) complexes in vitro against MDA-
MB-231 breast cancer cells revealed their high activity compared to
4. Conclusions the reference drug cisplatin (IC50 = 24.59 µg/mL). [PdCl(DPPB-HBH)]Cl
exhibited the best antiproliferative activity (IC50 = 9.3 µg/mL) and the
In this work, a series of palladium(II) complexes bearing imino­ highest DNA binding ability among the tested samples, which could be
phosphine ligands based on benzohydrazone and thiosemicarbazone due to the presence of carbonyl and hydroxyl moieties and their
derivatives of structural and biological significance were synthesized participation in the interaction with DNA that was further investigated
and characterized with different spectroscopic techniques. The by molecular docking and NCI studies. Chemical reactivity descriptors

13
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

Fig. 11. Molecular electrostatic maps of complexes ([PdCl(DPPB-HBH)]Cl (A), [PdCl(DPPB-PTSC)] (B), and [PdCl(DPPB-TSC)] (C)) along with their electrostatic
potential values.

were calculated to establish a logical relationship between structure and Vahid Noroozi Charandabi: Investigation, Formal analysis,
activity. Small energy gap values of Pd(II) complexes confirmed their Methodology.
visualization as biologically active molecules. Also, the high electro­
negativity and electrophilicity values of [PdCl(DPPB-HBH)]Cl justified Declaration of Competing Interest
its greater reactivity than others. Possible sites for electrophilic and
nucleophilic attacks were determined using MPE calculations and Fukui The authors declare that they have no known competing financial
functions. Considering the concordance of theoretical and experimental interests or personal relationships that could have appeared to influence
results, it seems that the present study can be effective in the targeted the work reported in this paper.
design of anticancer candidates.
Data availability
CRediT authorship contribution statement
Data will be made available on request.
Mahsa Pooyan: Investigation, Formal analysis, Methodology,
Writing – original draft. Zahra Shariatinia: Supervision, Conceptuali­ Acknowledgment
zation, Validation, Visualization, Project administration, Data curation,
Funding acquisition, Resources, Writing – review & editing. Fahimeh This work was supported by Iran National Science Foundation (INSF)
Mohammadpanah: Investigation, Formal analysis, Methodology. (INSF, Grant No. 99025423), and the authors thank all generous
Khodayar Gholivand: Conceptualization, Validation, Visualization, support.
Project administration. Peter C. Junk: Investigation, Formal analysis,
Methodology. Zhifang Guo: Investigation, Formal analysis, Methodol­
ogy. Mohammad Satari: Investigation, Formal analysis, Methodology.

14
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

Appendix A. Supplementary data [22] V. Radulović, A. Bacchi, G. Pelizzi, D. Sladić, I. Brčeski, K. Andjelković, Synthesis,
structure, and antimicrobial activity of complexes of Pt (II), Pd (II), and Ni (II) with
the condensation product of 2-(diphenylphosphino) benzaldehyde and
Supplementary data to this article can be found online at https://doi. semioxamazide, Monatshefte für Chemie/Chemical Monthly 137 (6) (2006)
org/10.1016/j.ica.2022.121368. 681–691.
[23] M. Milenković, A. Bacchi, G. Cantoni, S. Radulović, N. Gligorijević, S. Aranđelović,
D. Sladić, M. Vujčić, D. Mitić, K. Anđelković, Synthesis, characterisation and
References biological activity of Co (III) complex with the condensation product of 2-
(diphenylphosphino) benzaldehyde and ethyl carbazate, Inorg. Chim. Acta 395
[1] A.R. Kapdi, I.J. Fairlamb, Anti-cancer palladium complexes: a focus on PdX 2 L 2, (2013) 33–43.
palladacycles and related complexes, Chem. Soc. Rev. 43 (13) (2014) 4751–4777. [24] M. Milenković, A. Bacchi, G. Cantoni, J. Vilipić, D. Sladić, M. Vujčić,
[2] R.A. Alderden, M.D. Hall, T.W. Hambley, The discovery and development of N. Gligorijević, K. Jovanović, S. Radulović, K. Anđelković, Synthesis,
cisplatin, J. Chem. Educ. 83 (5) (2006) 728. characterization and biological activity of three square-planar complexes of Ni (II)
[3] C. Deegan, B. Coyle, M. McCann, M. Devereux, D.A. Egan, In vitro anti-tumour with ethyl (2E)-2-[2-(diphenylphosphino) benzylidene] hydrazinecarboxylate and
effect of 1, 10-phenanthroline-5, 6-dione (phendione),[Cu (phendione) 3](ClO4) 2⋅ monodentate pseudohalides, Eur. J. Med. Chem. 68 (2013) 111–120.
4H2O and [Ag (phendione) 2] ClO4 using human epithelial cell lines, Chem. Biol. [25] M. Milenković, A. Pevec, I. Turel, M. Vujčić, M. Milenković, K. Jovanović,
Interact. 164 (1–2) (2006) 115–125. N. Gligorijević, S. Radulović, M. Swart, M. Gruden-Pavlović, Synthesis,
[4] A. Karaküçük-İyidoğan, D. Taşdemir, E.E. Oruç-Emre, J. Balzarini, Novel platinum characterization, DFT calculation and biological activity of square-planar Ni (II)
(II) and palladium (II) complexes of thiosemicarbazones derived from 5-substitu­ complexes with tridentate PNO ligands and monodentate pseudohalides. Part II,
tedthiophene-2-carboxaldehydes and their antiviral and cytotoxic activities, Eur. J. Eur. J. Med. Chem. 87 (2014) 284–297.
Med. Chem. 46 (11) (2011) 5616–5624. [26] K. Adaila, M. Milenković, A. Bacchi, G. Cantoni, M. Swart, M. Gruden-Pavlović,
[5] M.K. Amir, S.Z. Khan, F. Hayat, A. Hassan, I.S. Butler, Anticancer activity, DNA- M. Milenković, B. Čobeljić, T. Todorović, K. Anđelković, Synthesis,
binding and DNA-denaturing aptitude of palladium (II) dithiocarbamates, Inorg. characterization, DFT calculations, and antimicrobial activity of Pd (II) and Co (III)
Chim. Acta 451 (2016) 31–40. complexes with the condensation derivative of 2-(diphenylphosphino)
[6] D. Anu, P. Naveen, N.P. Rath, M. Kaveri, Palladium (II) complexes containing benzaldehyde and Girard’s T reagent, J. Coord. Chem. 67 (22) (2014) 3633–3648.
substituted thiosemicarbazones. Synthesis, spectral characterization, X-ray [27] M.M. Đorđević, D.A. Jeremić, M.V. Rodić, V.S. Simić, I.D. Brčeski, V.M. Leovac,
crystallography, biomolecular interactions and in vitro cytotoxicity, J. Mol. Struct. Synthesis, structure and biological activities of Pd (II) and Pt (II) complexes with 2-
1206 (2020), 127703. (diphenylphosphino) benzaldehyde 1-adamantoylhydrazone, Polyhedron 68
[7] A. Garoufis, S. Hadjikakou, N. Hadjiliadis, Palladium coordination compounds as (2014) 234–240.
anti-viral, anti-fungal, anti-microbial and anti-tumor agents, Coord. Chem. Rev. [28] M. Milenković, G. Cantoni, A. Bacchi, V. Spasojević, M. Milenković, D. Sladić,
253 (9–10) (2009) 1384–1397. N. Krstić, K. Anđelković, Synthesis, characterization and antimicrobial activity of
[8] F. Huq, H. Tayyem, P. Beale, J.Q. Yu, Studies on the activity of three palladium (II) Pd (II) and Fe (III) complexes with ethyl (2E)-2-[2-(diphenylphosphino)
compounds of the form: Trans-PdL2Cl2 where L= 2-hydroxypyridine, 3- benzylidene] hydrazinecarboxylate, Polyhedron 80 (2014) 47–52.
hydroxypyridine, and 4-hydroxypyridine, J. Inorg. Biochem. 101 (1) (2007) 30–35. [29] V. Simic, S. Kolarevic, I. Brceski, D. Jeremic, B. Vukovic-Gacic, Cytotoxicity and
[9] E.A. Nyawade, N.R. Sibuyi, M. Meyer, R. Lalancette, M.O. Onani, Synthesis, antiviral activity of palladium (II) and platinum (II) complexes with 2-
characterization and anticancer activity of new 2-acetyl-5-methyl thiophene and (diphenylphosphino) benzaldehyde 1-adamantoylhydrazone, Turk. J. Biol. 40 (3)
cinnamaldehyde thiosemicarbazones and their palladium (II) complexes, Inorg. (2016) 661–669.
Chim. Acta 515 (2021), 120036. [30] W. Kabsch, Automatic processing of rotation diffraction data from crystals of
[10] F.F. Bobinihi, Group 10 dithiocarbamate complexes for biological applications and initially unknown symmetry and cell constants, J. Appl. Cryst. 26 (6) (1993)
as single source precursors to metal sulphide nanoparticles, North-West University 795–800.
(South Africa), 2019. [31] O.V. Dolomanov, L.J. Bourhis, R.J. Gildea, J.A. Howard, H. Puschmann, OLEX2: a
[11] A.C. Caires, Recent advances involving palladium (II) complexes for the cancer complete structure solution, refinement and analysis program, J. Appl. Cryst. 42
therapy, Anti-Cancer Agents in Medicinal Chemistry (Formerly Current Medicinal (2) (2009) 339–341.
Chemistry-Anti-Cancer Agents) 7(5) (2007) 484-491. [32] G.M. Sheldrick, Crystal structure refinement with SHELXL, Acta Crystallogr.
[12] S. Ray, R. Mohan, J.K. Singh, M.K. Samantaray, M.M. Shaikh, D. Panda, P. Ghosh, Section C Struct. Chem. 71 (1) (2015) 3–8.
Anticancer and antimicrobial metallopharmaceutical agents based on palladium, [33] A. Wolfe, G.H. Shimer Jr, T. Meehan, Polycyclic aromatic hydrocarbons physically
gold, and silver N-heterocyclic carbene complexes, J. Am. Chem. Soc. 129 (48) intercalate into duplex regions of denatured DNA, Biochemistry 26 (20) (1987)
(2007) 15042–15053. 6392–6396.
[13] M. Sebastian, N. Ninan, E. Elias, Nanomedicine and Cancer therapies, CRC Press, [34] D. Sabolová, M. Kožurková, T. Plichta, Z. Ondrušová, D. Hudecová, M. Šimkovič,
2012. H. Paulíková, A. Valent, Interaction of a copper (II)–Schiff base complexes with calf
[14] A. Bacchi, M. Carcelli, M. Costa, A. Fochi, C. Monici, P. Pelagatti, C. Pelizzi, thymus DNA and their antimicrobial activity, Int. J. Biol. Macromol. 48 (2) (2011)
G. Pelizzi, L.M.S. Roca, Stable alkynyl palladium (II) and nickel (II) complexes with 319–325.
terdentate PNO and PNN hydrazone ligands, J. Organomet. Chem. 593 (2000) [35] J.R. Lakowicz, G. Weber, Quenching of fluorescence by oxygen. Probe for
180–191. structural fluctuations in macromolecules, Biochemistry 12 (21) (1973)
[15] A. Bacchi, M. Carcelli, M. Costa, A. Leporati, E. Leporati, P. Pelagatti, C. Pelizzi, 4161–4170.
G. Pelizzi, Palladium (II) complexes containing a P, N chelating ligand Part II. [36] A. Mazloom-Jalali, Z. Shariatinia, I.A. Tamai, S.-R. Pakzad, J. Malakootikhah,
Synthesis and characterisation of complexes with different hydrazinic ligands. Fabrication of chitosan–polyethylene glycol nanocomposite films containing ZIF-8
Catalytic activity in the hydrogenation of double and triple CC bonds, nanoparticles for application as wound dressing materials, Int. J. Biol. Macromol.
J. Organomet. Chem. 535 (1–2) (1997) 107–120. 153 (2020) 421–432.
[16] M. Milenković, I.N. Shcherbakov, L.D. Popov, S.I. Levchenkov, S.A. Borodkin, G. [37] C.G. Ricci, P.A. Netz, Docking studies on DNA-ligand interactions: building and
G. Alexandrov, Synthesis, characterization and crystal structures of Ni (II) and Cu application of a protocol to identify the binding mode, J. Chem. Inf. Model. 49 (8)
(I) complexes with the condensation product of 2-(diphenylphosphino) (2009) 1925–1935.
benzaldehyde and 1-hydrazinophthalazine, Polyhedron 121 (2017) 278–284. [38] H.R. Drew, R.M. Wing, T. Takano, C. Broka, S. Tanaka, K. Itakura, R.E. Dickerson,
[17] S. Doherty, J.G. Knight, T.H. Scanlan, M.R. Elsegood, W. Clegg, Iminophosphines: Structure of a B-DNA dodecamer: conformation and dynamics, Proceedings of the
synthesis, formation of 2, 3-dihydro-1H-benzo [1, 3] azaphosphol-3-ium salts and National Academy of Sciences 78(4) (1981) 2179-2183.
N-(pyridin-2-yl)-2-diphenylphosphinoylaniline, coordination chemistry and [39] G.M. Morris, D.S. Goodsell, R.S. Halliday, R. Huey, W.E. Hart, R.K. Belew, A.
applications in platinum group catalyzed Suzuki coupling reactions and J. Olson, Automated docking using a Lamarckian genetic algorithm and an
hydrosilylations, J. Organomet. Chem. 650 (1–2) (2002) 231–248. empirical binding free energy function, J. Comput. Chem. 19 (14) (1998)
[18] T.F. Vaughan, D.J. Koedyk, J.L. Spencer, Comparison of the Reactivity of platinum 1639–1662.
(II) and platinum (0) complexes with iminophosphine and phosphinocarbonyl [40] D. Systemes BIOVIA, discovery studio modeling environment. Release 4.5, Dassault
ligands, Organometallics 30 (19) (2011) 5170–5180. Systemes 2015 San Diego, CA.
[19] B. Čobeljić, M. Milenković, A. Pevec, I. Turel, M. Vujčić, B. Janović, N. Gligorijević, [41] A.C. Wallace, R.A. Laskowski, J.M. Thornton, LIGPLOT: a program to generate
D. Sladić, S. Radulović, K. Jovanović, Investigation of antitumor potential of Ni (II) schematic diagrams of protein-ligand interactions, Protein Eng. Des. Sel. 8 (2)
complexes with tridentate PNO acylhydrazones of 2-(diphenylphosphino) (1995) 127–134.
benzaldehyde and monodentate pseudohalides, J. Biol. Inorg. Chem. 21 (2) (2016) [42] M. Vatanparast, Z. Shariatinia, Computational studies on the doped graphene
145–162. quantum dots as potential carriers in drug delivery systems for isoniazid drug,
[20] A. Kobayashi, D. Yamamoto, H. Horiki, K. Sawaguchi, T. Matsumoto, K. Nakajima, Struct. Chem. 29 (2018) 1427–1448.
H.-C. Chang, M. Kato, Photoinduced Dimerization Reaction Coupled with [43] T. Lu, F. Chen, Multiwfn: a multifunctional wavefunction analyzer, J. Comput.
Oxygenation of a Platinum (II)–Hydrazone Complex, Inorg. Chem. 53 (5) (2014) Chem. 33 (5) (2012) 580–592.
2573–2581. [44] R. Thomsen, M.H. Christensen, MolDock: a new technique for high-accuracy
[21] N. Malešević, T. Srdić, S. Radulović, D. Sladić, V. Radulović, I. Brčeski, K. molecular docking, J. Med. Chem. 49 (11) (2006) 3315–3321.
Anđelković, Synthesis and characterization of a novel Pd (II) complex with the [45] T. Wei, Z. Yang, M. Zhou, Y. Zhou, Comparison of biotransformation mechanisms
condensation product of 2-(diphenylphosphino) benzaldehyde and ethyl of 2, 4, 6-trinitrotoluene and its hydride-Meisenheimer metabolite by the old
hydrazinoacetate. Cytotoxic activity of the synthesized complex and related Pd (II) yellow enzyme family of flavoproteins, Energetic Mater. Front. 1 (3–4) (2020)
and Pt (II) complexes, Journal of inorganic biochemistry 100(11) (2006) 1811- 216–226.
1818.

15
M. Pooyan et al. Inorganica Chimica Acta 547 (2023) 121368

[46] W. Humphrey, A. Dalke, K. Schulten, VMD: visual molecular dynamics, J. Mol. [69] S. Tabassum, M. Ahmad, M. Afzal, M. Zaki, P.K. Bharadwaj, Synthesis and structure
Graph. 14 (1) (1996) 33–38. elucidation of a copper (II) Schiff-base complex: in vitro DNA binding, pBR322
[47] A.D. Becke, Becke’s three parameter hybrid method using the LYP correlation plasmid cleavage and HSA binding studies, J. Photochem. Photobiol. B Biol. 140
functional, J. Chem. Phys 98 (492) (1993) 5648–5652. (2014) 321–331.
[48] P.J. Stephens, F.J. Devlin, C.F. Chabalowski, M.J. Frisch, Ab initio calculation of [70] S. Tabassum, W.M. Al-Asbahy, M. Afzal, F. Arjmand, V. Bagchi, Molecular drug
vibrational absorption and circular dichroism spectra using density functional design, synthesis and structure elucidation of a new specific target peptide based
force fields, J. Phys. Chem. 98 (45) (1994) 11623–11627. metallo drug for cancer chemotherapy as topoisomerase I inhibitor, Dalton Trans.
[49] M. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, 41 (16) (2012) 4955–4964.
G. Scalmani, V. Barone, B. Mennucci, G. Petersson, Gaussian 09, revision D. 01, [71] A. Tarushi, E. Polatoglou, J. Kljun, I. Turel, G. Psomas, D.P. Kessissoglou,
Gaussian, Inc., Wallingford CT, 2009. Interaction of Zn (II) with quinolone drugs: structure and biological evaluation,
[50] P.J. Hay, W.R. Wadt, Ab initio effective core potentials for molecular calculations. Dalton Trans. 40 (37) (2011) 9461–9473.
Potentials for K to Au including the outermost core orbitals, J. Chem. Phys. 82 (1) [72] D.L. Arockiasamy, S. Radhika, R. Parthasarathi, B.U. Nair, Synthesis and DNA-
(1985) 299–310. binding studies of two ruthenium (II) complexes of an intercalating ligand, Eur. J.
[51] E. Cances, B. Mennucci, J. Tomasi, A new integral equation formalism for the Med. Chem. 44 (5) (2009) 2044–2051.
polarizable continuum model: Theoretical background and applications to [73] E.K. Efthimiadou, A. Karaliota, G. Psomas, Metal complexes of the third-generation
isotropic and anisotropic dielectrics, J. Chem. Phys. 107 (8) (1997) 3032–3041. quinolone antimicrobial drug sparfloxacin: Structure and biological evaluation,
[52] M. Cossi, V. Barone, B. Mennucci, J. Tomasi, Ab initio study of ionic solutions by a J. Inorg. Biochem. 104 (4) (2010) 455–466.
polarizable continuum dielectric model, Chem. Phys. Lett. 286 (3–4) (1998) [74] K. Ghosh, P. Kumar, N. Tyagi, U.P. Singh, V. Aggarwal, M.C. Baratto, Synthesis and
253–260. reactivity studies on new copper (II) complexes: DNA binding, generation of
[53] A.J. Foster, F. Weinhold, Natural hybrid orbitals, J. Am. Chem. Soc. 102 (24) phenoxyl radical, SOD and nuclease activities, Eur. J. Med. Chem. 45 (9) (2010)
(1980) 7211–7218. 3770–3779.
[54] https://www.chemcraftprog.com, ((accessed 18 October 2022)). [75] G. Psomas, A. Tarushi, E.K. Efthimiadou, Y. Sanakis, C.P. Raptopoulou, N. Katsaros,
[55] K. Gholivand, F. Mohammadpanah, M. Pooyan, R. Roohzadeh, Evaluating anti- Synthesis, structure and biological activity of copper (II) complexes with oxolinic
coronavirus activity of some phosphoramides and their influencing inhibitory acid, J. Inorg. Biochem. 100 (11) (2006) 1764–1773.
factors using molecular docking, DFT, QSAR, and NCI-RDG studies, J. Mol. Struct. [76] A.A. Sharfalddin, M.A. Hussien, Bivalence metal complexes of antithyroid drug
1248 (2022), 131481. carbimazole; synthesis, characterization, computational simulation, and biological
[56] A. Castiñeiras, R. Pedrido, Factors involved in the nuclearity of silver studies, J. Mol. Struct. 1228 (2021), 129725.
thiosemicarbazone clusters: Cocrystallization of two different sized tetranuclear [77] N. Biswas, S. Khanra, A. Sarkar, S. Bhattacharjee, D.P. Mandal, A. Chaudhuri,
silver (I) clusters derived from a phosphinothiosemicarbazone ligand, Inorg. Chem. S. Chakraborty, C.R. Choudhury, One new azido bridged dinuclear copper (II)
47 (13) (2008) 5534–5536. thiosemicarbazide complex: synthesis, DNA/protein binding, molecular docking
[57] K. Gholivand, Z. Shariatinia, M. Pourayoubi, 2J(P,C) and 3J(P,C) Coupling study and cytotoxicity activity, New J. Chem. 41 (21) (2017) 12996–13011.
Constants in Some New Phosphoramidates. Crystal Structures of CF3C(O)N(H)P(O) [78] W.-J. Lian, X.-T. Wang, C.-Z. Xie, H. Tian, X.-Q. Song, H.-T. Pan, X. Qiao, J.-Y. Xu,
[N(CH3)(CH2C6H5)]2 and 4-NO2-C6H4N(H)P(O)[4-CH3-NC5H9]2, Zeitschrift für Mixed-ligand copper (II) Schiff base complexes: the role of the co-ligand in DNA
anorganische und allgemeine Chemie 631(5) (2005) 961-967. binding, DNA cleavage, protein binding and cytotoxicity, Dalton Trans. 45 (22)
[58] M. Abdouss, S. Azodi-Deilami, E. Asadi, Z. Shariatinia, Synthesis of molecularly (2016) 9073–9087.
imprinted polymer as a sorbent for solid phase extraction of citalopram from [79] J. Olmsted III, D.R. Kearns, Mechanism of ethidium bromide fluorescence
human serum and urine, J. Mater. Sci. - Mater. Med. 23 (6) (2012) 1543–1552. enhancement on binding to nucleic acids, Biochemistry 16 (16) (1977) 3647–3654.
[59] A. Ghaffari, M. Behzad, M. Pooyan, H.A. Rudbari, G. Bruno, Crystal structures and [80] B.C. Baguley, M. Le Bret, Quenching of DNA-ethidium fluorescence by amsacrine
catalytic performance of three new methoxy substituted salen type nickel (II) Schiff and other antitumor agents: a possible electron-transfer effect, Biochemistry 23 (5)
base complexes derived from meso-1, 2-diphenyl-1, 2-ethylenediamine, J. Mol. (1984) 937–943.
Struct. 1063 (2014) 1–7. [81] S. Radisavljević, A.Đ. Kesić, D. Ćoćić, R. Puchta, L. Senft, M. Milutinović,
[60] S.D. Oladipo, B. Omondi, Mercury (II) N, N′ -diarylformamidine dithiocarbamates N. Milivojević, B. Petrović, Studies of the stability, nucleophilic substitution
as single-source precursors for the preparation of oleylamine-capped HgS reactions, DNA/BSA interactions, cytotoxic activity, DFT and molecular docking of
nanoparticles, Transit. Met. Chem. 45 (6) (2020) 391–402. some tetra-and penta-coordinated gold (iii) complexes, New J. Chem. 44 (26)
[61] C. Rao, R. Venkataraghavan, The C= S stretching frequency and the “− N− C= S (2020) 11172–11187.
bands” in the infrared, Spectrochim. Acta A: Mol. Spectrosc. 45 (1989) 299–305. [82] S.V. Baykov, A.S. Mikherdov, A.S. Novikov, K.K. Geyl, M.V. Tarasenko, M.
[62] Z. Shariatinia, Z. Nikfar, K. Gholivand, S. Abolghasemi Tarei, Antibacterial A. Gureev, V.P. Boyarskiy, π–π noncovalent interaction involving 1, 2, 4-and 1, 3,
activities of novel nanocomposite biofilms of chitosan/phosphoramide/Ag NPs, 4-oxadiazole systems: The combined experimental, theoretical, and database study,
Polym. Compos. 36 (3) (2015) 454–466. Molecules 26 (18) (2021) 5672.
[63] A.A. Ali, H. Nimir, C. Aktas, V. Huch, U. Rauch, K.-H. Schäfer, M. Veith, [83] K. Gholivand, F. Mohammadpanah, M. Pooyan, R. Roohzadeh, Evaluating anti-
Organoplatinum (II) complexes with 2-acetylthiophene thiosemicarbazone: coronavirus activity of some phosphoramides and their influencing inhibitory
synthesis, characterization, crystal structures, and in vitro antitumor activity, factors using molecular docking, DFT, QSAR, and NCI-RDG studies, J. Mol. Struct.
Organometallics 31 (6) (2012) 2256–2262. 131481 (2021).
[64] A.R. Latham, V. Hascall, H.B. Gray, The electronic structures and spectral [84] R.G. Parr, L.V. Szentpály, S. Liu, Electrophilicity index, J. Am. Chem. Soc. 121 (9)
properties of the square-planar dithiooxalate complexes of Nickel (II), Palladium (1999) 1922–1924.
(II), Platinum (II), and Gold (III), Inorg. Chem. 4 (6) (1965) 788–792. [85] L.R. Domingo, P. Pérez, The nucleophilicity N index in organic chemistry, Org.
[65] K. Gholivand, C.O. Della Védova, M.F. Erben, H.R. Mahzouni, Z. Shariatinia, Biomol. Chem. 9 (20) (2011) 7168–7175.
S. Amiri, Synthesis, spectroscopic study, X-ray crystallography and ab initio [86] M. Vatanparast, Z. Shariatinia, Revealing the role of different nitrogen
calculations of the two new phosphoramidates: C6H5OP(O)(NHC6H11)2 and [N functionalities in the drug delivery performance of graphene quantum dots: a
(CH3)(C6H11)]P(O)(2–C5H4N-NH)2, J. Mol. Struct. 874 (1) (2008) 178–186. combined density functional theory and molecular dynamics approach, J. Mater.
[66] R. Prabhakaran, P. Kalaivani, R. Huang, P. Poornima, V.V. Padma, F. Dallemer, Chem. B 7 (40) (2019) 6156–6171.
K. Natarajan, DNA binding, antioxidant, cytotoxicity (MTT, lactate dehydrogenase, [87] P.W. Ayers, R.G. Parr, Variational principles for describing chemical reactions: the
NO), and cellular uptake studies of structurally different nickel (II) Fukui function and chemical hardness revisited, J. Am. Chem. Soc. 122 (9) (2000)
thiosemicarbazone complexes: synthesis, spectroscopy, electrochemistry, and X- 2010–2018.
ray crystallography, J. Biol. Inorg. Chem. 18 (2) (2013) 233–247. [88] R.G. Parr, W. Yang, Density functional approach to the frontier-electron theory of
[67] X.-B. Fu, D.-D. Liu, Y. Lin, W. Hu, Z.-W. Mao, X.-Y. Le, Water-soluble DNA minor chemical reactivity, J. Am. Chem. Soc. 106 (14) (1984) 4049–4050.
groove binders as potential chemotherapeutic agents: synthesis, characterization, [89] W. Yang, W.J. Mortier, The use of global and local molecular parameters for the
DNA binding and cleavage, antioxidation, cytotoxicity and HSA interactions, analysis of the gas-phase basicity of amines, J. Am. Chem. Soc. 108 (19) (1986)
Dalton Trans. 43 (23) (2014) 8721–8737. 5708–5711.
[68] F.H. Allen, O. Kennard, D.G. Watson, L. Brammer, A.G. Orpen, R. Taylor, Tables of [90] C. Morell, A. Grand, A. Toro-Labbé, New dual descriptor for chemical reactivity,
bond lengths determined by X-ray and neutron diffraction. Part 1. Bond lengths in Chem. A Eur. J. 109 (1) (2005) 205–212.
organic compounds, J. Chem. Soc., Perkin Trans. 2 (12) (1987) S1–S19.

16

You might also like