Journal Pre-proof Nickel complexes of aroylhydrazone ligand: synthesis, crystal structure and DNA binding properties
Ping Yang, Li-Lei Zhang, Zi-Zhou Wang, Dan-Dan Zhang, YaMin Liu, Qing-Shan Shi, Xiao-Bao Xie PII:
S0162-0134(19)30549-5
DOI:
https://doi.org/10.1016/j.jinorgbio.2019.110919
Reference:
JIB 110919
To appear in:
Journal of Inorganic Biochemistry
Received date:
17 August 2019
Revised date:
3 November 2019
Accepted date:
11 November 2019
Please cite this article as: P. Yang, L.-L. Zhang, Z.-Z. Wang, et al., Nickel complexes of aroylhydrazone ligand: synthesis, crystal structure and DNA binding properties, Journal of Inorganic Biochemistry (2019), https://doi.org/10.1016/j.jinorgbio.2019.110919
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2019 Published by Elsevier.
Journal Pre-proof
Nickel complexes of Aroylhydrazone ligand: synthesis, crystal structure and DNA binding properties
Ping Yang a, 1, Li-Lei Zhang b, 1, Zi-Zhou Wang c, Dan-Dan Zhang a, Ya-Min Liu a,
a
of
Qing-Shan Shi a, *, Xiao-Bao Xie a, *
Guangdong Institute of Microbiology, Guangdong Academy of Sciences,
ro
Guangzhou 510070, China.
-p
State Key Laboratory of Applied Microbiology Southern China.
re
Guangdong Provincial Key Laboratory of Microbial Culture
College of Chemistry and Chemical Engineering, Luoyang Normal University,
Luoyang 471000, China.
School of Chemistry and Chemical Engineering, Guangzhou University, 230 Wai
Jo ur
c
na
b
lP
Collection and Application. Guangdong Open Laboratory of Applied Microbiology.
Huan Xi Road, Guangzhou Higher Education Mega Center, Guangzhou 510006, China.
*
Corresponding author at: Guangdong Institute of Microbiology, Guangdong Academy of Sciences, Guangzhou 510070, China. E-mail:
[email protected] (Q-S. Shi);
[email protected] (X.-B. Xie). 1 Both the authors contributed equally to this work. 1
Journal Pre-proof
Highlights • Three novel nickel complexes of aroylhydrazone ligand (NiL1-3) were synthesized. • The ligands were coordinated by enolate form. • Complex NiL2 and NiL3 exhibit moderate binding affinity toward calf Thymus DNA.
of
• Complex NiL2 and NiL3 bind DNA through minor groove binding and
Jo ur
na
lP
re
-p
ro
intercalation.
2
Journal Pre-proof
Abstract In
this
work,
three
aroylhydrazone
ligands
((E)-2-hydroxy-N'-(1-(pyrazin-2-yl)ethylidene)benzohydrazide, (E)-3-hydroxy-N'-(1-(pyrazin-2-yl)ethylidene)benzohydrazide, (E)-4-hydroxy-N'-(1-(pyrazin-2-yl)ethylidene)benzohydrazide,
HL1; HL2; HL3)
and
and their
of
complexes with nickel (Ni(L1)2, NiL1; Ni(L2)2∙2DMF, NiL2; Ni(L3)2∙2DMF, NiL3) were prepared. The single crystal X-ray structures analysis of three compounds
ro
showed that they were neutral. The ligand adopts tridentate chelating mode. The
-p
nickel ion is six-coordinate with two O atoms and four N atoms from two ligands, and
re
forms an octahedral arrangement. The investigation of DNA binding ability by
lP
ultraviolet and fluorescence titrations showed that NiL2 and NiL3 exhibit moderate
na
binding affinity toward calf Thymus DNA. Spectroscopy, molecular docking, and molecular dynamics simulation indicated that NiL2 and NiL3 bind at the minor
Jo ur
groove of DNA through intercalation.
Graphical abstract
DNA binding of nickel aroylhydrazone complexes with different hydroxyl substitution positions was studied by ultraviolet spectra, fluorescence spectra, molecular docking, and molecular dynamics simulation.
3
Journal Pre-proof
Keywords
of
Pyrazine; Aroylhydrazone; Nickel complexes; Single-crystal structure; DNA binding 1 Introduction
ro
Metal complexes play an essential role in cancer treatment[1]. Various
-p
complexes from cisplatin (from the earliest discovery) to ruthenium complexes have
re
entered into their clinical trial stage[2][3][4]. These complexes have potential
lP
anticancer activity due to their strong DNA binding ability[5][6]. These complexes
na
can be tethered to DNA by covalent interaction or bound to DNA by non-covalent interaction, such as π-π stacking, electrostatic binding, H-bonding [7][8]. The
Jo ur
interaction of metal complexes with DNA is much dependent on the molecular structures of ligands. In earlier studies, the ligands o-phenanthroline and bipyridine were commonly used due to their planar structure[9][10], which allowed them to insert into DNA easily. Besides, thiosemicarbazone complexes have been shown to have significant anti-proliferative effects[11][12]. Aroylhydrazone is an excellent class of metal-chelating ligand [13]. The acyl hydrazone units of the aroylhydrazone have many interesting properties, such as the degree of rigidity, the conjugated π-system, and the protonated-deprotonated site on NH group [14][15]. Some aroylhydrazones have been used as iron chelators to treat 4
Journal Pre-proof
iron overload diseases or thalassemia[16][17]. Aroylhydrazone metal complexes showed various of biological activities, such as DNA/protein interaction[14][18], antioxidant[15][19], antibacterial[20][21] and anticancer[22][23]. Altering the structure and properties of aroylhydrazone compounds (such as electronegativity of heteroatom, ring size, and planarity) can affect its interaction with DNA[14][24].
of
Based on these considerations, three aroylhydrazone ligands were selected to study, which contain the methyl pyrazinylketone moiety and with different
ro
substitution positions of the hydroxyl groups on the benzene ring. The three ligands
-p
are close to planar and are potentially tridentate (N, N, O) chelators. The potential
re
H-bond donor (hydroxyl group) and the potential H-bond acceptor (uncoordinated
lP
pyrazyl N-atom) are favorably disposed to bind DNA. Moreover, the octanol-water
na
partition coefficients (log P) for ligands are about 1.9 [16], which indicates a good water solubility and lipid permeability. If they can be used as candidate drugs for the
Jo ur
treatment of cancer, they will be well absorbed by the human body. As such, we synthesized their nickel complexes and determined the crystal structure for the complexes. The binding ability and mode of the complexes with DNA were investigated using ultraviolet spectra, fluorescence spectra, molecular docking, and molecular dynamics simulation.
2 Materials and methods 2.1 Reagents and instrumentation Solvents and reagents were purchased from commercial companies and used 5
Journal Pre-proof
without further purification. Nuclear magnetic resonance (NMR) spectra were measured in dimethyl sulfoxide (DMSO) using a Bruker AVANCE III HD 600 (600 M). Mass spectra (MS) were measured using a Bruker maXis impact, equipped with an electrospray ion source (ESI), operated in a negative ion mode. Infrared (IR) spectra were recorded on a Bruker Tensor II using KBr pellets. Ultraviolet (UV)
of
spectra were measured using a Thermo Scientific NanoDrop One. Fluorescence spectra were measured using a PerkinElmer LS 45 fluorescence spectrometer.
ro
2.2 Synthesis
-p
(E)-2-hydroxy-N'-(1-(pyrazin-2-yl)ethylidene)benzohydrazide (HL1): According to a
re
method based on literature[25], a solution of 2-acetylpyrazine (1.22 g, 10 mmol) in
lP
ethanol (20 mL) was slowly added to a solution of 2-hydroxybenzohydrazide (1.52 g, 10 mmol) in methanol (20 mL). After the mixture solution was refluxed at 65 °C for 4
na
h, it was cooled down, suction filtered, and then washed with cold ethanol. The final product of 2.1 g with a yield of 82% was obtained by air-drying. MS, NMR, IR, and
Jo ur
UV data are shown in the Supporting information (Section S1).
(E)-3-hydroxy-N'-(1-(pyrazin-2-yl)ethylidene)benzohydrazide (HL2): A solution of 2-acetylpyrazine (1.22 g, 10 mmol) in ethanol (20 mL) was slowly added to a solution of 3-hydroxybenzohydrazide (1.52 g, 10 mmol) in ethanol (20 mL). The mixture solution was then refluxed at 65 °C for 4 h; after that, it was cooled down, suction filtered, and then washed with cold ethanol. The final product (1.65 g, 64% yield) was obtained by air-drying. MS, NMR, IR, and UV data are shown in the Supporting information (Section S1).
6
Journal Pre-proof (E)-4-hydroxy-N'-(1-(pyrazin-2-yl)ethylidene)benzohydrazide (HL3): According to a method based on literature[16], we made some modifications. 2-acetylpyrazine (1.22 g, 10 mmol) and 4-hydroxybenzohydrazide (1.52 g, 10 mmol) were mixed with 10 mL of DMSO. The solution mixture was refluxed at 75 °C for 4 h, and was after that cooled down, filtered, and washed with cold ethanol. Two grams of product (78% yield) were obtained by air-drying. MS, NMR, IR, and UV data are shown in the
of
Supporting information (Section S1).
ro
NiL1: A solution of Ni(NO3)2∙6H2O (2.91 mg, 0.01 mmol) in methanol (1 mL) was
-p
slowly added to a solution of HL1 (5.12 mg, 0.02 mmol) in N,N-dimethylformamide
re
(DMF) (1 mL). After three weeks of static volatilization at room temperature, single crystals of Ni(L1)2 were obtained with a yield (calculated based on metal ions) of
na
(Section S1).
lP
86%. Elemental analysis and IR data are shown in the Supporting information
Jo ur
NiL2: A solution of Ni(NO3)2∙6H2O (2.91 mg, 0.01 mmol) in methanol (1 mL) was slowly added to a solution of HL2 (5.12 mg, 0.02 mmol) in DMF (1 mL). After three weeks of static volatilization at room temperature, single crystals of Ni(L2)2∙2DMF were obtained with a yield (calculated based on metal ions) of 80%. Elemental analysis, IR, and UV data are shown in the Supporting information (Section S1).
NiL3: A solution of NiCl2∙6H2O (2.38 mg, 0.01 mmol) in methanol (1 mL) was slowly added into a solution of HL3 (5.12 mg, 0.02 mmol) in DMF (1 mL). After three weeks of static volatilization at room temperature, single crystals of Ni(L3)2∙2DMF were obtained with a yield (calculated based on metal ions) of 88%. 7
Journal Pre-proof Elemental analysis, IR, and UV data are shown in the Supporting information (Section S1).
2.3 Single-crystal X-ray diffraction Data for the complex of NiL1 was recorded with a CCD area detector and Oxford Gemini S Ultra to collect the diffraction data (Mo-Kα, λ = 0.071073 nm) at room temperature. Due to free solvent molecules in the lattice, data for the complexes
of
of NiL2 and NiL3 were recorded with a CCD area detector and Oxford Gemini S
ro
Ultra to collect the diffraction data (Cu-Kα, λ = 0.154178 nm) at room temperature.
-p
The SADABS program was used for absorption correction[26]. The crystal structure
re
was solved through the direct methods [27], and then, the SHELXTL program package [28][29] was used to conduct a full-matrix least-square structure refinement
lP
based on F2. The non-hydrogen atoms were all anisotropically refined. All H atoms
na
were placed in geometrically idealized positions, with C-H = 0.93 Å (sp2), C-H = 0.96 Å (sp3) and O-H = 0.82 Å[30]. Further details of the X-ray structural analyses
Jo ur
for complexes NiL1, NiL2, and NiL3 are given in Table 1. Selected bond lengths and angles for NiL1, NiL2, and NiL3 are listed in the Supporting information (Table S1).
2.4 UV titration Calf Thymus DNA (ctDNA) was resuspended in 10 mM Tris-HCl buffers, pH 7.0 containing 50 mM NaCl. The ratio of UV absorbance at 260 and 280 nm (A260/A280) of ctDNA stock solution was 1.9, which indicates that it is free of protein contamination. The DNA concentration was expressed as monomer units (nucleotides) and determined by UV absorption at 260 nm using a molar absorptivity of 6600 8
Journal Pre-proof L∙mol−1∙cm−1. The complex was dissolved in DMSO and then diluted with water. Due to lower water solubility of the complex, the concentration of DMSO in the final solution was kept lower than 5% (v/v) after the complexes were mixed with DNA. All measurements were performed at room temperature. For titration, 500 μL of 0.47 mM ctDNA solutions were mixed with 500 μL of H2O and different concentrations of the complex solutions (0.176 mM, 0.352 mM, 0.528 mM, 0.704 mM, 0.880 mM, 1.056
of
mM, 1.232 mM, 1.408 mM, 1.584 mM, and 1.760 mM). The absorbance (λmax = 260
ro
nm) of ctDNA was then measured. Comparatively, 500 μL of 0.088 mM NiL2 and
-p
0.176 mM NiL3 solutions were mixed with 500 μL of H2O and different
re
concentrations of ctDNA solution (0.047 mM, 0.141 mM, 0.235 mM, 0.329 mM, and
lP
0.423 mM). It is to study the effect of ctDNA in the absorption of the complexes. The
Jo ur
λmax = 389 and 309 nm.
na
absorption of NiL2 was measured at λmax = 383 nm, and that of NiL3 was measured at
2.5 Fluorescence titration
To further clarify the interaction of the complexes with DNA, steady-state competitive binding experiment was carried out. The GelStain GS101-02 dye was purchased from TransGen Biotech, Beijing, China. Dye (10×) was mixed with 4.7 mM ctDNA for 1 h and then diluted 1:10 to obtain working solution of ctDNA-dye complex (concentration = 0.47 mM). Then, 200 μL of the working solution of ctDNA-dye complex was mixed with 200 μL of H2O and various concentrations of the complex solutions (0.176 mM, 0.352 mM, 0.528 mM, 0.704 mM, 0.880 mM, and 9
Journal Pre-proof 1.056 mM). After that, the change of fluorescence emission spectra at λem = 600 nm (λex = 254) of the mixture solutions was recorded to study the effect of the complexes on the binding of ctDNA to dye. To further confirm the binding mode, the fluorescence titration of DAPI-DNA (DAPI is 4',6-diamidino-2-phenylindole) and methyl green-DNA by complexes NiL2 and NiL3 were carried out. DAPI and methyl green are minor and major groove’s binders
of
of DNA, respectively[31][32][33]. 200 μL ctDNA (0.47 mM) reacted with 8 μL DAPI
ro
staining solution (C1005) and 2 μL methyl green staining solution (C0115, Beyotime
-p
Biotech, Shanghai, China) for 5 minutes, respectively. Then the solution of
re
DAPI-DNA and methyl green-DNA were mixed with 200 μL H2O and complex
lP
solutions (0.088mM, 0.176 mM, 0.352 mM, 0.528 mM, 0.704 mM, and 0.880 mM)
na
respectively. The fluorescence emission spectra of λem = 454 nm (λex = 364) and λem =
Jo ur
677 nm (λex = 633) were recorded.
2.6 In silico studies
The initial conformations of NiL2 and NiL3 were obtained from the single crystal X-ray structure data. The crystal structure of human B-DNA (PDB ID: 1BNA) (The sequence is CGCGAATTCGCG) was downloaded from the RCSB Protein Data Bank[34][35]. The heteroatoms and water molecules were removed from the B-DNA before docking simulation. Molecular docking studies on ligands and B-DNA were carried out using the AutoDock Vina program[36]. During the process of docking simulation, the whole B-DNA crystal structure was considered active sites. Finally, 10
Journal Pre-proof
NiL2 and NiL3 were docked into the B-DNA to find and select the lowest energy among 20 docking modes given by cluster analysis, respectively. Molecular dynamics (MD) simulation of DNA-NiL2 / DNA-NiL3 complex was carried out using GROMACS 2019 software package[37]. The lowest binding energy conformation for each complex was employed as the initial conformation. The
of
AMBER99SB-ILDN force field was applied to describe the DNA receptor, whereas the universal force field was performed to describe ligands. Each system was
ro
immersed in a periodic water box of a cube shape. To hold the system at an
-p
electrically neutral state, we added appropriate numbers of sodium ions. The
re
simulation was started by energy minimization for 500 steps with the steepest descent
lP
method, followed by 100 ps restricted dynamics relaxation. Then, each system was
na
gradually heated from 0 K to 310 K within 200 ps. A 10 ns MD simulation with a time step of 2 fs was performed. During the simulation, the PME (Particle Mesh Ewald)
Jo ur
algorithm was employed to deal with the long-range electrostatic interactions. The velocity-rescaling method was used to control the temperatures of the DNA-NiL2 / DNA-NiL3 complex and the remaining atoms in this system.
3 Results and discussion 3.1 Spectroscopic characterization Compared with the raw materials’ spectra, the IR spectrum of HL1-HL3 showed new peaks at 1608, 1613, and 1609 cm-1 due to the stretching vibration of imine in hydrazone. For HL1, a peak at 752 cm-1 belongs to the out-of-plane bending vibration 11
Journal Pre-proof of the ortho-disubstituted benzene ring. For HL2, two peaks at 745 and 681 cm-1 belong to the out-of-plane bending vibration of the meta-disubstituted benzene ring. For HL3, a peak at 847 cm-1 belongs to the out-of-plane bending vibration of the para-disubstituted benzene ring. The carbonyl vibration at 1651, 1658, and 1656 cm-1 of HL1-HL3 disappeared in their complexes. The IR spectra of NiL1-NiL3 appeared new bands at 1596, 1580, and 1596 cm-1, which indicates that -NH-C=O was
of
adjusted to an enolate form -N=C-OH. The result is consistent with the single crystal
ro
X-ray analysis (see below). The IR spectra of these complexes are typical of other
-p
aroylhydrazone analogs[16].
re
The UV absorption peaks of HL1-HL3 in DMSO at room temperature appeared
lP
at 316 nm (ε = 2.34 × 104 L∙mol−1∙cm−1), 306 nm (ε = 4.67 × 104 L∙mol−1∙cm−1) and
na
309 nm (ε = 4.75 × 104 L∙mol−1∙cm−1), corresponding to π-π* transition. Similar peaks at 307 nm (ε = 3.25 × 104 L∙mol−1∙cm−1) and 315 nm (ε = 3.68 × 104 L∙mol−1∙cm−1)
Jo ur
were also observed in the spectra of NiL2 and NiL3, respectively (Due to poor solubility of NiL1, it was not tested). New peaks at 423 nm (ε = 1.33 × 104 L∙mol−1∙cm−1) and 423 nm (ε = 1.88 × 104 L∙mol−1∙cm−1) in the spectra of NiL2 and NiL3, respectively, were attributed to metal-to-ligand charge transfer (MLCT) transition[19]. Consequently, the complex solutions show a yellow-green color, which is different from the colorless ligand solutions. It also shows that the complexes are stable in solution.
3.2 Crystal structure 12
Journal Pre-proof
As shown in Table 1, the complexes NiL1-NiL3 belong to the orthorhombic system. The nickel center is six-coordinate with two O atoms and four N atoms from two aroylhydrazone ligands and forms an octahedron structure (Fig. 1d). In free aroylhydrazone ligand, the methyl group (C6) is in a syn position to the pyrazine ring nitrogen atom (N1)[16][25]. Here, the C6 and N1 of the complex is an anti-
of
conformation. Because the ligand is disposed to coordinate a metal ion through the pyrazinyl N-atom (N1), imine N-atom (N3), and carbonyl O-atom (O1) after rotation
ro
of the pyrazine group by 180° about C4–C5. The torsion angle of N1–C4–C5–C6 is in
-p
the ranges −175.0(5) to −179.4(5)°. In each case, the ligands bind Ni by a tri-dentate
re
chelating, and the NH group next to the carbonyl deprotonated to generate an enolate
lP
form of the ligand (Fig. 1). The structure is like that found for the related Co, Ni, and
na
Zn complex[38][39][40]. The oxygen atom (O2) on the phenolic group did not deprotonate and stuck to the outside of the coordination sphere to form intermolecular
Jo ur
hydrogen bonds, which is conducive to the growth and stability of crystals. Hydrogen bonds are shown in Table S2, Table S3, and Table S4. Since NiL1 contains a strong intramolecular hydrogen bond O2-H2A∙∙∙N4 (2.553(7) Å, 147.7°), its solubility is weak. The Ni-O and Ni-N bond lengths are in the ranges 2.080(2)-2.097(4) and 1.973(2)-2.144(5) Å (Table S1). Bond length is close to the statistic value of (N, N, O) tridentate chelated nickel complexes (N–Ni 2.093±0.06, N(middle)–Ni 1.973±0.041, O–Ni 2.107±0.083 Å) in CCDC (Sample is 311). The angle between coordination bonds was like that of nickel analogs [38] (Table S1). They deviate from 180° and 90°, indicating that the octahedron is a little distorted. 13
Journal Pre-proof
Table 1. Crystal data and structure refinement for NiL1-NiL3. NiL2
NiL3
Empirical formula
C26H22N8NiO4
C32H36N10NiO6
C32H36N10NiO6
Formula weight
569.22
715.42
715.42
Temperature (K)
298(2)
298(2)
298(2)
Wavelength (Å)
0.71073
1.54178
1.54178
Crystal system
Orthorhombic
Orthorhombic
Orthorhombic
Space group
Aba2
Pbcn
Pbcn
a (Å)
12.5534(7)
11.6963(2)
11.6656(3)
b (Å)
18.8700(13)
9.40620(10)
9.6490(3)
c (Å)
10.0629(8)
31.3167(5)
30.3106(8)
Volume (Å3) Z
2383.7(3)
3445.39(9)
3411.80(16)
4
4
4
Density (calculated) (Mg/m3) Absorption coefficient (mm-1)
1.586
1.379
1.393
1.302
1.315
F(000)
1176
1496
1496
Reflections collected
2400
Independent reflections, Rint
1542, 0.0269
Data / restraints / parameters Goodness-of-fit on F2 R1 [I > 2σ(I)]
ro -p
0.867
6435
2773, 0.0236
2697, 0.0197
1542 / 1 / 179
2773 / 0 / 226
2697 / 0 / 226
1.039
1.066
1.039
0.0359
0.0439
0.0377
0.0959
0.1189
0.1036
0.268 and -0.220
0.446 and -0.349
0.274 and -0.253
re
19218
lP
wR2 (all data)
Jo ur
na
Largest diff. peak and hole (e.Å-3)
of
NiL1
14
Journal Pre-proof
Figure 1 ORTEP drawing of complex NiL1 (a), NiL2 (b), NiL3 (c) and polyhedral of central nickel (d) with thermal ellipsoids 30% probability. For clarity, hydrogen atoms and solvent have been neglected. Symmetry code: A: -x, 1-y, z (in NiL1); -x, y, 3/2-z (in NiL2); -x, y, 1/2-z (in NiL3).
of
3.3 UV titration To investigate the binding affinity with the complex and DNA, UV titration was
ro
carried out. At a constant ctDNA concentration, the characteristic absorption of
-p
double-stranded DNA at 260 nm increased from the increasing concentration of the
re
complexes (Fig. 2a and Fig. 2b). At a constant complex concentration, the
lP
characteristic absorption of the complexes NiL2 and NiL3 at 383nm/389nm increased
na
from the increase in ctDNA concentration (Fig. 2c and Fig. 2d). Hyperchromicity has also been reported in the literature[41][42][43]. It appears that the complexes undergo
Jo ur
a distortion of the coordination sphere after binding to DNA resulting in an enhancing of the MLCT bands[44]. The characteristic absorption of the complex NiL3 at 309 nm decreased with the increase of ctDNA concentration (Fig. 2d). The hypochromism may be due to an increasing of electron cloud density around the complexes after binding to DNA, leading to a decreasing in the transition probability of π-π* intraligand[45][46][47]. Although the Ni(II) complex showed hypochromism of about 9.8% without any shift at 309 nm, the complexes interact with DNA most likely through an intercalation mode[19]. The DNA binding constants of NiL2 and NiL3 complexes, calculated by the equation: [DNA]/(εa − εf) = [DNA]/(εb − εf) + 1/Kb (εb − 15
Journal Pre-proof εf)[48][49], were 2 ×104 and 1×104 M−1. The binding constants have the same order of magnitude as that of the six-coordinated nickel analogs (~104) [50] but are lower than that of the four-coordinated nickel analogs (~105) [13]. Gibbs Free Energy of the binding of the complexes NiL2 and NiL3 with ctDNA were calculated to be −24.5 and −22.8 kJ/mol (−5.8 and −5.4 kcal/mol), using the equation: ΔG = −RTlnKb, where
of
R = 8.314 J·K−1·mol−1, and T = 298 K. The results indicate spontaneous binding
3
2
1
0 240
260
280
300
320
340
360
Absorbance
2.0
1.5
1.0
0.5
420
-p 3
2
1
0 240
440
260
280
300
320
340
360
380
400
420
440
Wavelength (nm)
0.423mM ctDNA 0.329mM ctDNA 0.235mM ctDNA 0.141mM ctDNA 0.047mM ctDNA 0mM ctDNA
(d) 3.5
0.423mM ctDNA 0.329mM ctDNA 0.235mM ctDNA 0.141mM ctDNA 0.047mM ctDNA 0mM ctDNA
3.0
Jo ur
(c) 2.5
400
na
Wavelength (nm)
380
4
2.5
Absorbance
Absorbance
4
5
Absorbance
5
re
(a)
1.760mM NiL3 1.584mM NiL3 1.408mM NiL3 1.232mM NiL3 1.056mM NiL3 0.880mM NiL3 0.704mM NiL3 0.528mM NiL3 0.352mM NiL3 0.176mM NiL3 0mM NiL3
(b)
lP
1.760mM NiL2 1.584mM NiL2 1.408mM NiL2 1.232mM NiL2 1.056mM NiL2 0.880mM NiL2 0.704mM NiL2 0.528mM NiL2 0.352mM NiL2 0.176mM NiL2 0mM NiL2
ro
between these compounds and ctDNA.
2.0 1.5 1.0 0.5 0.0 240 260 280 300 320 340 360 380 400 420 440 460 480 500
0.0 240 260 280 300 320 340 360 380 400 420 440 460 480 500
Wavelength (nm)
Wavelength (nm)
Figure 2 UV spectra: (a) titration of NiL2 against ctDNA; (b) titration of NiL3 against ctDNA; (c) titration of ctDNA against NiL2; and (d) titration of ctDNA against NiL3.
3.4 Fluorescence titration Further support concerning the binding of complexes to DNA was obtained using 16
Journal Pre-proof
a fluorescence titration to have a better insight into the binding effects. The ctDNA bound to dye emitted a robust orange light of 600 nm. However, the fluorescence intensity was weakened and quenched after the complex was added. This phenomenon suggests that the complexes can replace the dye on the DNA-dye (Fig. 3a and Fig. 3b). The quenching process conforms to the linear Stern-Volmer
of
equation[51][52]: I0/I = 1 + Ksqr, where I0 and I represent the fluorescence intensity without and with the addition of the complex, Ksq is the static quenching constant, and
ro
r is the concentration ratio of the complex and DNA. The Ksq of the complexes NiL2
analog
(E)-N'-(1-(3-methylpyrazin-2-yl)ethylidene)benzohydrazide
re
the
-p
and NiL3 were 1.17 and 0.68 (Fig. 3c and Fig. 3d). The Ksq value is close to that of nickel
lP
complexes[52]. The Ksq of HL2 and HL3 was 0.41 and 0.38 (Fig. S1). The results
na
suggested that the binding interaction of the Ni(II) complex with ctDNA was stronger than that of the free ligand. The formation of nickel complexes could enhance the
Jo ur
biological activity of free aroylhydrazone[23].
160 140
Intensity
120
180 160 140 120
Intensity
180
0mM NiL3 0.176mM NiL3 0.325mM NiL3 0.528mM NiL3 0.704mM NiL3 0.880mM NiL3 1.056mM NiL3
(b) 200
0mM NiL2 0.176mM NiL2 0.325mM NiL2 0.528mM NiL2 0.704mM NiL2 0.880mM NiL2 1.056mM NiL2
(a) 200
100 80
100 80
60
60
40
40
20
20 0
0 560
580
600
620
640
660
680
560
700
580
600
620
640
Wavelength (nm)
Wavelength (nm)
17
660
680
700
Journal Pre-proof
(d) 2.8 (c) 4.0
y = 0.6868x + 0.9893 R2 = 0.987
2.6
y = 1.1769x + 0.9988 R2 = 0.989
3.5
2.4 2.2
3.0
2.0
I0/I
I0/I
2.5
1.8 1.6
2.0
1.4 1.2
1.5
1.0 1.0
0.8 0.0
0.5
1.0
1.5
2.0
0.0
2.5
0.5
1.0
1.5
2.0
2.5
r
r
Figure 3 (a) and (b) are fluorescence titration of NiL2 and NiL3 against ctDNA-dye.
of
λex = 254 nm. (c) and (d) are the variation of fluorescence intensity ratio with the
ro
concentration ratio of NiL2 and NiL3 to DNA at 600 nm.
-p
The fluorescence titration of NiL2 and NiL3 to DAPI-DNA showed that the
re
fluorescence intensity decreases with the increase of complex concentration (Fig. 4a
lP
and Fig. 4b). But the fluorescence intensity of methyl green-DNA unchanged with the
na
increase of complex concentration (Fig. 4c and Fig. 4d). These results indicate that
200 180 160 140
Intensity
120 100 80 60
(b)
0mM NiL2 0.088mM NiL2 0.176mM NiL2 0.325mM NiL2 0.528mM NiL2 0.704mM NiL2 0.880mM NiL2
200
160 140 120 100 80 60
40
40
20
20
0 400
0mM NiL3 0.088mM NiL3 0.176mM NiL3 0.325mM NiL3 0.528mM NiL3 0.704mM NiL3 0.880mM NiL3
180
Intensity
(a)
Jo ur
NiL2 and NiL3 bind to the minor groove of DNA.
0
450
500
550
600
400
450
500
Wavelength (nm)
Wavelength (nm)
18
550
600
Journal Pre-proof
12
8
Intensity
(d)
0mM NiL2 0.088mM NiL2 0.176mM NiL2 0.325mM NiL2 0.528mM NiL2 0.704mM NiL2 0.880mM NiL2
10
6
12
8
6
4
4
2
2
0 660
680
700
720
0mM NiL3 0.088mM NiL3 0.176mM NiL3 0.325mM NiL3 0.528mM NiL3 0.704mM NiL3 0.880mM NiL3
10
Intensity
(c)
0 660
740
680
Wavelength (nm)
700
720
740
Wavelength (nm)
Figure 4 Fluorescence spectra of DAPI-DNA with NiL2 (a), DAPI-DNA with NiL3
ro
of
(b), methyl green-DNA with NiL2 (c), and methyl green-DNA with NiL3 (d).
-p
3.5 Molecular docking and molecular dynamics (MD) simulation
re
To better understand the binding mode and force of the complex and DNA,
lP
molecular docking and MD simulation were carried out using the structures of B-DNA (PDB ID: 1BNA) and aroylhydrazone complexes NiL2 and NiL3. The
na
docking data indicated that the least relative binding energies of DNA-NiL2 and
Jo ur
DNA-NiL3 were −8.6 and −8.8 kcal/mol. The calculated binding energy difference between DNA-NiL2 and DNA-NiL3 is small (0.2 kcal/mol), which is an agreement with the experimental difference (0.4 kcal/mol) obtained from the UV spectroscopic data. It indicated that the binding affinity of the two complexes to DNA is similar. After 10 ns MD simulation, the systems of DNA-NiL2 and DNA-NiL3 were equilibrated with no visible RMSD fluctuations (Fig. 5). Video of molecular movements can be seen in the supporting information. The results support that parallel-displaced π-π stacking interaction is the dominant force in stabilizing the combination of DNA and complexes (Fig. 6). It is consistent with the experimental data of spectra titration. NiL2 and NiL3 bind to the minor groove of DNA through 19
Journal Pre-proof inserting phenolic group into DC3:DG22 – DG4:DC21 and DG2:DC23 – DC3:DG22 base pairs [53][54][55]. The position of the hydroxyl group has little effect on the binding mode of nickel complexes with DNA. (a)
0.5
(b)
NiL2 DNA
NiL3 DNA
0.5
0.4
RMSD / nm
0.2
0.1
0.3
0.2
0.1
0.0
0.0 0
2000
4000
6000
8000
10000
0
2000
4000
6000
8000
10000
Time / ps
ro
Time / ps
of
RMSD / nm
0.4 0.3
Jo ur
na
lP
re
-p
Figure 5 MD simulation of NiL2 (a) and NiL3 (b) with B-DNA (PDB ID: 1BNA)
Figure 6 The lowest binding energy conformation of NiL2 (A) and NiL3 (B) bound to B-DNA (PDB ID: 1BNA). DA, light gray; DT, pink; DG, green; DC, gray.
4 Conclusion Three nickel aroylhydrazone complexes were synthesized, and their crystal structures were determined. The metal to ligand ratio of complexes is 1:2. The nickel center is a six-coordinate octahedral configuration. The acyl hydrazone unit of the ligand was converted into an enol structure. The ligands are tridentate chelate metal 20
Journal Pre-proof
ions. The spectroscopy, molecular docking, and molecular dynamics simulation results showed that the complex NiL2 and NiL3 bind to the minor groove of DNA through intercalation.
Acknowledgement
of
This work was supported by the National Natural Science Foundation of China (Grant 41701349), GDAS' Project of Science and Technology Development
-p
ro
(2018GDASCX-0912) and Nanyue Talent Fund (GDIMYET20180205).
re
Supplementary material
lP
Crystallographic data for the structures in this paper have been deposited with the Cambridge Crystallographic Data Centre, CCDC, 12 Union Road, Cambridge
na
CB21EZ, UK. Copies of the data can be obtained free of charge on quoting the
Jo ur
depository numbers CCDC-1906506, CCDC-1906519, and CCDC-1906530 for NiL1, NiL2 and NiL3 (Fax: +44-1223-336-033; E-Mail:
[email protected], http://www.ccdc.cam.ac.uk).
MS, NMR, IR, UV, and selected bond lengths and angles of compounds were shown in the Supporting information. The video is in another uploaded file.
References: [1]
M.A. Jakupec, M. Galanski, V.B. Arion, C.G. Hartinger, B.K. Keppler, Antitumour metal compounds: more than theme and variations, Dalt. Trans. (2008) 183–194. doi:10.1039/B712656P.
[2]
A.K. Bytzek, G. Koellensperger, B.K. Keppler, C. G. Hartinger, 21
Journal Pre-proof
Biodistribution of the novel anticancer drug sodium trans-[tetrachloridobis(1H-indazole)ruthenate(III)] KP-1339/IT139 in nude BALB/c mice and implications on its mode of action, J. Inorg. Biochem. 160 (2016) 250–255. doi:https://doi.org/10.1016/j.jinorgbio.2016.02.037. [3]
C.G. Hartinger, M.A. Jakupec, S. Zorbas-Seifried, M. Groessl, A. Egger, W. Berger, H. Zorbas, P.J. Dyson, B.K. Keppler, KP1019, A New Redox-Active
of
Anticancer Agent – Preclinical Development and Results of a Clinical Phase I
J.M. Rademaker-Lakhai, A Phase I and Pharmacological Study with
re
[4]
-p
doi:10.1002/cbdv.200890195.
ro
Study in Tumor Patients, Chem. Biodivers. 5 (2008) 2140–2155.
lP
Imidazolium-trans-DMSO-imidazole-tetrachlororuthenate, a Novel Ruthenium
na
Anticancer Agent, Clin. Cancer Res. 10 (2004) 3717–3727. doi:10.1158/1078-0432.CCR-03-0746. M.R. Gill, P.J. Jarman, S. Halder, M.G. Walker, H.K. Saeed, J.A. Thomas, C.
Jo ur
[5]
Smythe, K. Ramadan, K.A. Vallis, A three-in-one-bullet for oesophageal cancer: replication fork collapse, spindle attachment failure and enhanced radiosensitivity generated by a ruthenium(ii) metallo-intercalator, Chem. Sci. 9 (2018) 841–849. doi:10.1039/C7SC03712K. [6]
Q. Cao, Y. Li, E. Freisinger, P.Z. Qin, R.K.O. Sigel, Z.-W. Mao, G-quadruplex DNA targeted metal complexes acting as potential anticancer drugs, Inorg. Chem. Front. 4 (2017) 10–32. doi:10.1039/C6QI00300A.
[7]
A.C. Komor, J.K. Barton, The path for metal complexes to a DNA target, 22
Journal Pre-proof
Chem. Commun. 49 (2013) 3617–3630. doi:10.1039/C3CC00177F. [8]
K.E. Erkkila, D.T. Odom, J.K. Barton, Recognition and Reaction of Metallointercalators with DNA, Chem. Rev. 99 (1999) 2777–2796. doi:10.1021/cr9804341.
[9]
X.-H. Zou, B.-H. Ye, H. Li, J.-G. Liu, Y. Xiong, L.-N. Ji, Mono- and
of
bi-nuclear ruthenium(II) complexes containing a new asymmetric ligand 3-(pyrazin-2-yl)-as-triazino[5,6-f ]1,10-phenanthroline: synthesis,
-p
1423–1428. doi:10.1039/A900064J.
ro
characterization and DNA-binding properties, J. Chem. Soc. Dalt. Trans. (1999)
re
[10] Y. An, Y.-Y. Lin, H. Wang, H.-Z. Sun, M.-L. Tong, L.-N. Ji, Z.-W. Mao,
lP
Cleavage of double-strand DNA by zinc complexes of dicationic 2,2′-dipyridyl
na
derivatives, Dalt. Trans. (2007) 1250–1254. doi:10.1039/B616754C. [11] W.H. Morris, L. Ngo, J.T. Wilson, W. Medawala, A.R. Brown, J.D. Conner, F.
Jo ur
Fabunmi, D.J. Cashman, E.C. Lisic, T. Yu, J.E. Deweese, X. Jiang, Structural and Metal Ion Effects on Human Topoisomerase IIα Inhibition by α-(N)-Heterocyclic Thiosemicarbazones, Chem. Res. Toxicol. 32 (2019) 90–99. doi:10.1021/acs.chemrestox.8b00204. [12] J. Wang, Y. Gou, Z. Zhang, P. Yu, J. Qi, Q. Qin, H. Sun, X. Wu, H. Liang, F. Yang, Developing an Anticancer Copper(II) Multitarget Pro-Drug Based on the His146 Residue in the IB Subdomain of Modified Human Serum Albumin, Mol. Pharm. 15 (2018) 2180–2193. doi:10.1021/acs.molpharmaceut.8b00045. [13] R. Raj Kumar, R. Ramesh, Synthesis, molecular structure and electrochemical 23
Journal Pre-proof
properties of nickel(ii) benzhydrazone complexes: influence of ligand substitution on DNA/protein interaction, antioxidant activity and cytotoxicity, RSC Adv. 5 (2015) 101932–101948. doi:10.1039/C5RA19530F. [14] P. Krishnamoorthy, P. Sathyadevi, R.R. Butorac, A.H. Cowley, N.S.P. Bhuvanesh, N. Dharmaraj, Variation in the biomolecular interactions of
of
nickel(ii) hydrazone complexes upon tuning the hydrazide fragment, Dalt. Trans. 41 (2012) 6842–6854. doi:10.1039/C2DT30121K.
ro
[15] Y. Li, Z. Yang, M. Zhou, Y. Li, J. He, X. Wang, Z. Lin, Ni(II) and Co(II)
-p
complexes of an asymmetrical aroylhydrazone: synthesis, molecular structures,
re
DNA binding, protein interaction, radical scavenging and cytotoxic activity,
lP
RSC Adv. 7 (2017) 41527–41539. doi:10.1039/C7RA05504H.
na
[16] D.S. Kalinowski, P.C. Sharpe, P. V Bernhardt, D.R. Richardson, Structure– Activity Relationships of Novel Iron Chelators for the Treatment of Iron
Jo ur
Overload Disease: The Methyl Pyrazinylketone Isonicotinoyl Hydrazone Series, J. Med. Chem. 51 (2008) 331–344. doi:10.1021/jm7012562. [17] P. Bendova, E. Mackova, P. Haskova, A. Vavrova, E. Jirkovsky, M. Sterba, O. Popelova, D.S. Kalinowski, P. Kovarikova, K. Vavrova, D.R. Richardson, T. Simunek, Comparison of Clinically Used and Experimental Iron Chelators for Protection against Oxidative Stress-Induced Cellular Injury, Chem. Res. Toxicol. 23 (2010) 1105–1114. doi:10.1021/tx100125t. [18] S.P. Dash, A.K. Panda, S. Pasayat, R. Dinda, A. Biswas, E.R.T. Tiekink, Y.P. Patil, M. Nethaji, W. Kaminsky, S. Mukhopadhyay, S.K. Bhutia, Syntheses 24
Journal Pre-proof and structural investigation of some alkali metal ion-mediated LVVO2− (L2− = tridentate ONO ligands) species: DNA binding, photo-induced DNA cleavage and cytotoxic activities, Dalt. Trans. 43 (2014) 10139–10156. doi:10.1039/c4dt00883a. [19] Y. Li, Y. Li, Z. Yang, F. Meng, N. Wang, M. Zhou, Z. Xia, Q. Gong, Q. Gao,
of
Distinct supramolecular assemblies of Fe(iii) and Ni(ii) complexes constructed from the o-vanillin salicylhydrazone ligand: syntheses, crystal structures,
ro
DNA/protein interaction, and antioxidant and cytotoxic activity, New J. Chem.
-p
43 (2019) 8024–8043. doi:10.1039/C8NJ06530F.
re
[20] L. Pan, C. Wang, K. Yan, K. Zhao, G. Sheng, H. Zhu, X. Zhao, D. Qu, F. Niu,
lP
Z. You, Synthesis, structures and Helicobacter pylori urease inhibitory activity
na
of copper(II) complexes with tridentate aroylhydrazone ligands, J. Inorg. Biochem. 159 (2016) 22–28.
Jo ur
doi:https://doi.org/10.1016/j.jinorgbio.2016.02.017. [21] A. Jamadar, A.-K. Duhme-Klair, K. Vemuri, M. Sritharan, P. Dandawate, S. Padhye, Synthesis, characterisation and antitubercular activities of a series of pyruvate-containing aroylhydrazones and their Cu-complexes, Dalt. Trans. 41 (2012) 9192–9201. doi:10.1039/C2DT30322A. [22] Y. Gou, J. Li, B. Fan, B. Xu, M. Zhou, F. Yang, Structure and biological properties of mixed-ligand Cu(II) Schiff base complexes as potential anticancer agents, Eur. J. Med. Chem. 134 (2017) 207–217. doi:https://doi.org/10.1016/j.ejmech.2017.04.026. 25
Journal Pre-proof
[23] Y. Li, Z. Yang, M. Zhou, Y. Li, Synthesis and crystal structure of new monometallic Ni(ii) and Co(ii) complexes with an asymmetrical aroylhydrazone: effects of the complexes on DNA/protein binding property, molecular docking, and in vitro anticancer activity, RSC Adv. 7 (2017) 49404– 49422. doi:10.1039/C7RA10283F.
of
[24] P. Sathyadevi, P. Krishnamoorthy, R.R. Butorac, A.H. Cowley, N.S.P. Bhuvanesh, N. Dharmaraj, Effect of substitution and planarity of the ligand on
ro
DNA/BSA interaction, free radical scavenging and cytotoxicity of diamagnetic
re
9702. doi:10.1039/C1DT10767D.
-p
Ni(ii) complexes: A systematic investigation, Dalt. Trans. 40 (2011) 9690–
lP
[25] T. Xi-Shi, F. Yi-Min,
na
2-Hydroxy-N’-[1-(3-methylpyrazin-2-yl)ethylidene]benzohydrazide, Acta Crystallogr. Sect. E. 64 (2008) o707. doi:10.1107/S1600536808006697.
Jo ur
[26] L. Krause, R. Herbst-Irmer, G.M. Sheldrick, D. Stalke, Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination, J. Appl. Crystallogr. 48 (2015) 3–10. doi:10.1107/S1600576714022985. [27] G.M. Sheldrick, A short history of SHELX, Acta Crystallogr. Sect. A. 64 (2008) 112–122. doi:10.1107/S0108767307043930. [28] G. Sheldrick, SHELX, (2019). http://shelx.uni-ac.gwdg.de/. [29] G.M. Sheldrick, Crystal structure refinement with SHELXL, Acta Crystallogr. Sect. C. 71 (2015) 3–8. doi:10.1107/S2053229614024218. 26
Journal Pre-proof
[30] S. Øien, D.S. Wragg, K.P. Lillerud, M. Tilset, Di-μ-chlorido-bis[(2,2’-bipyridine-5,5’-dicarboxylic acid-κ2N,N’)chloridocopper(II)] dimethylformamide tetrasolvate, Acta Crystallogr. Sect. E. 69 (2013) m73–m74. doi:10.1107/S1600536812051422. [31] M.S. Haroone, Y. Hu, P. Zhang, R. Ma, X. Zhang, Q.M. Kaleem, M.B. Khan, J.
of
Lu, Fluorescence enhancement strategy for evaluation of the minor groove binder DAPI to complementary ssDNA sequence including telomere mimics in
ro
(ssDNA@DAPI/LDH)n ultrathin films, Dye. Pigment. 166 (2019) 422–432.
-p
doi:https://doi.org/10.1016/j.dyepig.2019.03.007.
re
[32] S.K. Kim, B. Nordén, Methyl green, FEBS Lett. 315 (1993) 61–64.
lP
doi:10.1016/0014-5793(93)81133-K.
na
[33] D. Prieto, G. Aparicio, P.E. Morande, F.R. Zolessi, A fast, low cost, and highly efficient fluorescent DNA labeling method using methyl green, Histochem.
Jo ur
Cell Biol. 142 (2014) 335–345. doi:10.1007/s00418-014-1215-0. [34] RCSB Protein Data Bank, (2019). http://rcsb.org. [35] P.W. Rose, B. Beran, C. Bi, W.F. Bluhm, D. Dimitropoulos, D.S. Goodsell, A. Prlić, M. Quesada, G.B. Quinn, J.D. Westbrook, J. Young, B. Yukich, C. Zardecki, H.M. Berman, P.E. Bourne, The RCSB Protein Data Bank: redesigned web site and web services, Nucleic Acids Res. 39 (2010) D392– D401. doi:10.1093/nar/gkq1021 %J Nucleic Acids Research. [36] O. Trott, A.J. Olson, AutoDock Vina: Improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading, 27
Journal Pre-proof
J. Comput. Chem. 31 (2010) 455–461. doi:10.1002/jcc.21334. [37] M.J. Abraham, T. Murtola, R. Schulz, S. Páll, J.C. Smith, B. Hess, E. Lindahl, GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers, SoftwareX. 1–2 (2015) 19–25. doi:https://doi.org/10.1016/j.softx.2015.06.001.
of
[38] D.-B. Dang, Y. Bai, C.-Y. Duan, Bis{N-(2-hydroxybenzoyl)-N’-[1-(2-pyridyl)ethylidene]hydrazinato}nickel(II),
-p
doi:10.1107/S1600536806032624.
ro
Acta Crystallogr. Sect. E. 62 (2006) m2290--m2292.
re
[39] X.-S. Tai, Y.-M. Feng, H.-X. Zhang, Bis[2-acetyl-3-methylpyrazine
lP
(2-hydroxybenzoyl)hydrazonato]zinc(II) monohydrate, Acta Crystallogr. Sect.
na
E. 64 (2008) m656. doi:10.1107/S1600536808009707. [40] X.-F. Hou, X.-L. Zhao, L. Zhang, W.-N. Wu, Y. Wang, Co(II)/Zn(II)/Cu(II)
Jo ur
Complexes Containing Hydrazone Ligand Bearing Pyrazine Unit: Syntheses, Crystal Structures and Fluorescence Properties, Chinese J. Inorg. Chem. 34 (2018) 201–205. doi:10.11862/CJIC.2018.005. [41] C. Hiort, P. Lincoln, B. Norden, DNA binding of .DELTA.and .LAMBDA.-[Ru(phen)2DPPZ]2+, J. Am. Chem. Soc. 115 (1993) 3448– 3454. doi:10.1021/ja00062a007. [42] C. Moucheron, A. Kirsch-De Mesmaeker, S. Choua, Photophysics of Ru(phen)2(PHEHAT)2+: A Novel “Light Switch” for DNA and Photo-oxidant for Mononucleotides, Inorg. Chem. 36 (1997) 584–592. 28
Journal Pre-proof
doi:10.1021/ic9609315. [43] Z. Chen, J. Zhang, S. Zhang, Oxidative DNA cleavage promoted by two phenolate-bridged binuclear copper(II) complexes, New J. Chem. 39 (2015) 1814–1821. doi:10.1039/C4NJ01623H. [44] V. Rajendiran, M. Murali, E. Suresh, M. Palaniandavar, V.S. Periasamy, M.A. Akbarsha, Non-covalent DNA binding and cytotoxicity of certain mixed-ligand
ro
(2008) 2157–2170. doi:10.1039/b715077f.
of
ruthenium(ii) complexes of 2,2′-dipyridylamine and diimines, Dalt. Trans.
-p
[45] P. Jaividhya, M. Ganeshpandian, R. Dhivya, M.A. Akbarsha, M. Palaniandavar,
re
Fluorescent mixed ligand copper(II) complexes of anthracene-appended Schiff
lP
bases: studies on DNA binding, nuclease activity and cytotoxicity, Dalt. Trans.
na
44 (2015) 11997–12010. doi:10.1039/C5DT00899A. [46] C. Rajarajeswari, M. Ganeshpandian, M. Palaniandavar, A. Riyasdeen, M.A.
Jo ur
Akbarsha, Mixed ligand copper(II) complexes of 1,10-phenanthroline with tridentate phenolate/pyridyl/(benz)imidazolyl Schiff base ligands: Covalent vs non-covalent DNA binding, DNA cleavage and cytotoxicity, J. Inorg. Biochem. 140 (2014) 255–268. doi:10.1016/j.jinorgbio.2014.07.016. [47] M. Ganeshpandian, S. Ramakrishnan, M. Palaniandavar, E. Suresh, A. Riyasdeen, M.A. Akbarsha, Mixed ligand copper(II) complexes of 2,9-dimethyl-1,10-phenanthroline: Tridentate 3N primary ligands determine DNA binding and cleavage and cytotoxicity, J. Inorg. Biochem. 140 (2014) 202–212. doi:10.1016/j.jinorgbio.2014.07.021. 29
Journal Pre-proof
[48] Y.-C. Huang, J. Haribabu, C.-M. Chien, G. Sabapathi, C.-K. Chou, R. Karvembu, P. Venuvanalingam, W.-M. Ching, M.-L. Tsai, S.C.N. Hsu, Half-sandwich Ru(η6-p-cymene) complexes featuring pyrazole appended ligands: Synthesis, DNA binding and in vitro cytotoxicity, J. Inorg. Biochem. 194 (2019) 74–84. doi:https://doi.org/10.1016/j.jinorgbio.2019.02.012.
of
[49] M. Alagesan, N.S.P. Bhuvanesh, N. Dharmaraj, Binuclear copper complexes: Synthesis, X-ray structure and interaction study with nucleotide/protein by
ro
in vitro biochemical and electrochemical analysis, Eur. J. Med. Chem. 78 (2014)
-p
281–293. doi:https://doi.org/10.1016/j.ejmech.2014.03.043.
re
[50] P. Krishnamoorthy, P. Sathyadevi, A.H. Cowley, R.R. Butorac, N. Dharmaraj,
lP
Evaluation of DNA binding, DNA cleavage, protein binding and in vitro
na
cytotoxic activities of bivalent transition metal hydrazone complexes, Eur. J. Med. Chem. 46 (2011) 3376–3387.
Jo ur
doi:https://doi.org/10.1016/j.ejmech.2011.05.001. [51] E. Ramachandran, V. Gandin, R. Bertani, P. Sgarbossa, K. Natarajan, N.S.P. Bhuvanesh, A. Venzo, A. Zoleo, A. Glisenti, A. Dolmella, A. Albinati, C. Marzano, Synthesis, characterization and cytotoxic activity of novel copper(II) complexes with aroylhydrazone derivatives of 2-Oxo-1,2-dihydrobenzo[h]quinoline-3-carbaldehyde, J. Inorg. Biochem. 182 (2018) 18–28. doi:https://doi.org/10.1016/j.jinorgbio.2018.01.016. [52] W.-W. Wang, Y. Wang, L. Zhang, Y.-F. Song, W.-N. Wu, Z. Chen, Syntheses, Crystal Structures and DNA-Bing Properties of Cu(II)/Ni(II) Complexes with 30
Journal Pre-proof
Acylhydrazone Ligand Bearing Pyrazine Unit, Chinese J. Inorg. Chem. 35 (2019) 563–568. doi:10.11862/CJIC.2019.037. [53] C. Icsel, V.T. Yilmaz, B. Cevatemre, M. Aygun, E. Ulukaya, Structures and anticancer activity of chlorido platinum(II) saccharinate complexes with monoand dialkylphenylphosphines, J. Inorg. Biochem. 195 (2019) 39–50. doi:https://doi.org/10.1016/j.jinorgbio.2019.03.008.
of
[54] H. Wang, W. Wang, W.J. Jin, σ-Hole Bond vs π-Hole Bond: A Comparison
-p
doi:10.1021/acs.chemrev.5b00527.
ro
Based on Halogen Bond, Chem. Rev. 116 (2016) 5072–5104.
re
[55] Y. Chen, Y.-B. Wang, Y. Zhang, W. Wang, Accurate calculations of the
lP
noncovalent systems with flat potential energy surfaces: Naphthalene dimer
na
and azulene dimer, Comput. Theor. Chem. 1112 (2017) 52–60.
Jo ur
doi:https://doi.org/10.1016/j.comptc.2017.04.021.
31
Journal Pre-proof Conflict of Interest
Jo ur
na
lP
re
-p
ro
of
There are no conflicts to declare.
32