Sie sind auf Seite 1von 6

Inorganic Chemistry Communications 103 (2019) 6–11

Contents lists available at ScienceDirect

Inorganic Chemistry Communications


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

Short communication

Synthesis, structure, and DNA-binding study of a novel Zn (II) complex with T


fleroxacin and 1,10-phenanthroline monohydrate
Yujie Zhu, Xinnuo Xiong, Zili Suo, Peixiao Tang , Qiaomei Sun, Xiaohui Ding, Hui Li
⁎ ⁎

School of Chemical Engineering, Sichuan University, Chengdu, Sichuan, China

GRAPHICAL ABSTRACT

ARTICLE INFO ABSTRACT

Keywords: A novel metal-based compound of zinc complexing with fleroxacin (flrx) was prepared in the presence of 1,10-
Zinc(II) complex phenanthroline (phen) monohydrate and characterized by various techniques. The crystal structure of the
Crystal structure complex ([Zn(flrx)(phen)(H2O)](NO3)·2H2O) was successfully disclosed through single-crystal X-ray diffraction,
Fleroxacin and Zn(II) connected with pyridone oxygen and the carboxylic oxygen atom of flrx by coordination bonds. The
DNA interaction
complex showed a square pyramid, and its structure was stabilized by intermolecular hydrogen bonding and π–π
Molecule docking
stacking. The interaction between the complex and calf thymus DNA (ctDNA) was studied. Results indicated that
[Zn(flrx)(phen)(H2O)](NO3)·2H2O bound to ctDNA through a static mechanism, and the hydrophobic force was
the main driving force in the binding process. Circular dichroism spectra, iodide quenching studies, and DNA
melting temperature experiments indicated that the complex bound to ctDNA in the groove. Molecular docking
further verified the groove binding mode and provided a visualized view for the interactions between the
complex and ctDNA.

1. Introduction complexes have been testified to exert good gain effects in interacting
with biological macromolecules [7–9]. In addition, as a framework that
Metal organic complexes are considered to hold great therapeutic can be functionalized and designed, mixed ligands complexes provide
and diagnostic potential [1]. A lot of research motivated the develop- abundant opportunities for developing new drugs [10–12]. N-hetero-
ment of new metal-based drugs to overcome the limitations of current cyclic carbons (NHCs) are a common class of auxiliary ligands to sta-
therapeutic products, particularly, in terms of toxicity, resistance, and bilize metal ions and fulfill the structural qualification for organome-
other pharmacological defects [2]. Various metal complexes have been tallic complexes [13,14].
used in clinical practice in terms of anti-inflammatory [3], antibacterial It has been proven that the use of fluoroquinolones would result in
[4], anti-oxidative [5], and anti-tumor [6] activities. Among them, zinc gene mutations and toxicities in the host cells [15]. Moreover, metal


Corresponding authors.
E-mail addresses: tangpeixiao@126.com (P. Tang), lihuilab@sina.com (H. Li).

https://doi.org/10.1016/j.inoche.2019.02.039
Received 23 January 2019; Received in revised form 22 February 2019; Accepted 24 February 2019
Available online 26 February 2019
1387-7003/ © 2019 Elsevier B.V. All rights reserved.
Y. Zhu, et al. Inorganic Chemistry Communications 103 (2019) 6–11

ions, given their positive electricity, are likely to bind negatively Table 1
charged nucleic acids [16]. Higher affinities towards DNA caused high Selected bond distance and angles for complex.
cytotoxicities [17]. As a result, the metal-fluoroquinolone complexes Bond distance (Å)
might potentially lead to DNA damage, even mammalian cell cyto- Zn(1)-O(1) 1.956(3) Zn(1)-O(3) 2.045(3)
toxicity. Hence, studies on the interaction between metal-based com- Zn(1)-N(1) 2.082(3) Zn(1)-N(2) 2.113(3)
plexes and mammal DNA can indicate their affinities, and provide the Zn(1)-O(4) 2.010(3)

basis for determining the impact of complexes on the structure and Bond angle (°)
function of DNA [18]. O(1)-Zn(1)-O(3) 90.24(12) O(1)-Zn(1)-O(4) 103.37(15)
O(1)-Zn(1)-N(1) 153.56(14) O(1)-Zn(1)-N(2) 90.31(13)
Fleroxacin (flrx) is a widely used fluoroquinolones with the target of
O(3)-Zn(1)-N(1) 91.76(12) O(3)-Zn(1)-N(2) 159.95(13)
topoisomerase II and IV [19]. However, it has been studied rarely in O(4)-Zn(1)-O(3) 92.98(13) O(4)-Zn(1)-N(1) 102.85(13)
metal complexes [20]. In our previous research, several Cu (II) com- O(4)-Zn(1)-N(2) 106.38(14) N(1)-Zn(1)-N(2) 79.09(13)
plexes with flrx and 1, 10-phenanthroline (phen) were synthesized and
characterized [21]. As an extension of our previous work, we synthe-
sized a Zn-based drug with flrx and phen as ligands and acquired a The angles of Zn(II) and H2O and oxygen atoms [O(4)-Zn(1)-O
single crystal. The structure was analyzed by X-ray crystallography and (1) = 103.37(15)°, O(4)-Zn(1)-O(3) = 92.98(13)°, O(4)-Zn(1)-N
characterized cooperating with other techniques. To explore the con- (1) = 102.85(13)°, and O(4)-Zn(1)-N(2) = 106.38(14)°] explain the O
formation damage in DNA induced by the binding of the Zn-fleroxacin (4) in the coordinative water is nearly perpendicular to the bottom
complex, the interaction between the complex and ctDNA was in- plane. The entire complex molecule involves a square pyramid.
vestigated. Molecular docking was performed to show the details of Therefore, the distances between the central Zn(II) and its coordinative
binding visually, which has been seldom applied in previous reports on atoms are approximate [Zn(1)-O(1) = 1.956(3) Å,Zn(1)-O
metal–quinolone compounds. (3) = 2.045(3) Å,Zn(1)-N(1) = 2.082(3) Å,Zn(1)-N
(2) = 2.113(3) Å,Zn(1)-O(4) = 2.010(3) Å].
Fig. A1 exhibits the hydrogen bonds among adjacent complex mo-
2. Results and discussion
lecules, and the specific data are recorded in Table A2. The main types
of hydrogen bonds present in this system are OeH⋯O and OeH⋯N.
A clear crystal structure is necessary for new drug research and
The oxygen atoms belong to coordinative water molecules, crystal
development. Single-crystal X-ray diffraction (SXRD) was used to
water molecules, flrx and nitrate ions, whereas nitrogen atoms were
characterize the structure of the complex. Crystal data and structure
derived from phen. According to the stacked plot, complex molecules
refinement are listed in Table A1, which suggests that the synthesized
showed anti-parallel structure. The distance between flrx and phen was
compound is in triclinic form with its space group behaving as p-1.
shorter than 3.4 Å, and the degree of the angle was less than 20°. Thus,
Fig. 1 exhibits a molecular model of the complex. We observed that the
π–π stacking exists between neighboring asymmetric molecules [24].
complex is a penta-coordinate compound. The ligating atoms consist of
The existence of hydrogen bond forces and aromatic ring accumulation
central Zn(II), two nitrogen atoms from phen, a pyridone oxygen, a
force together maintain the stability of the crystal structure.
carboxylate oxygen atom from the deprotonated flrx, and an oxygen
Crystallographic data of the complex was submitted to the
atom from water. The coordination of N1 and N2 of phen and Zn(II)
Cambridge Crystallographic Data Centre (CCDC), and it was given the
formed a stable five-membered ring, and then the coordination of O1
deposition numbers CCDC–1848167.
and O3 of flrx and Zn(II) constituted a stable six-membered ring. The
FT-IR spectroscopy illustrates the mechanism by showing changes in
existence of this structure is beneficial to the stability of complex mo-
the functional groups before and after synthesis. Information obtained
lecules. Similar structures and space group have been reported in pre-
from the spectra (Fig. 2 (a)) shows that stretching vibrations caused by
vious Zn(NO3)2-phen complexes [22,23]. But in this work, the aromatic
v(C]O)carboxylate and v(CeO)carboxylate of the carboxylate (eCOOH) in
ring of the flrx ligand is nearly in the same plane of the phen (Fig. A1),
free flrx produce peaks in 1720 and 1286 cm−1. After synthesis, the
which prefers to form aromatic ring accumulation force so as to make
peak in 1720 cm−1 disappears, whereas two new peaks appear in 1582
the complex crystal more stable.
and 1326 cm−1 as the asymmetric [νasym(CO2)] and symmetrical
This research demonstrates a penta-coordinate environment around
[νsym(CO2)] stretching vibrations of carbonxylate. This result means
Zn(II) with square pyramid geometry which is deduced by pivotal bond
that the carboxylate oxygen atom of the flrx ligand chelates to Zn(II).
distances and angles listed in Table 1. The angles of the Zn(II) with
The coordination mode of carboxylate ligand can be predicated as
oxygen atoms and nitrogen atoms [O(1)-Zn(1)-N(1) = 153.56(14)°, O
monodentate mode due to Δ[ν(CO2)asym-ν(CO2)sym] = 256 cm−1. The
(3)-Zn(1)-N(2) = 159.95(13)°] show that the central Zn(II) almost lo-
vibration absorption peak of keto groups of free flrx in 1625 cm−1
cates in the same plane with its coordination atoms from flrx and phen.
hypochromatic shifted to 1622 cm−1 after synthesis. Thus, we con-
sidered that the flrx ligands coordinate with Zn(II) by oxygen atoms of
3-carboxylate and 4-keto [25]. The conclusion was consistent with the
analytical results of the single crystal structure.
Thermogravimetric (TG) analysis was used to determine the com-
positions of complex single crystal. The results shown in Fig. 2(b)
suggest that the complex dehydrates in two steps. The initial mass loss
stage was from 40 °C to 60 °C with a weight loss of 4.84%. Then, the
complex lost 2.48% weight at 108 °C to 150 °C. The two crystal water
molecules (Clacd = 4.93%) were lost in the first step, whereas a co-
ordinated water molecule (Clacd = 2.47%) was lost in the second step.
That mean the complex molecule can remain stable below 108 °C. The
following thermal decomposition consists of two processes. The first
course began in 217 °C and finished in 405 °C with the loss of phen at
25.09% (Calcd = 24.70%) in weight. In the next course, in the range of
420 °C–630 °C, a weight loss of 57.24% (Calcd = 56.72%) was achieved
because of the releases of flrx and NO2. The remaining product weight
Fig. 1. Molecular structure of [Zn(flrx)(phen)(H2O)]·NO3·2H2O. of 10.71% (Calcd = 11.11%) corresponded to ZnO.

7
Y. Zhu, et al. Inorganic Chemistry Communications 103 (2019) 6–11

Fig. 2. (a) FT-IR spectra of flrx and [Zn(flrx)(phen)(H2O)]·NO3·2H2O. (b) TG-DTG curves of [Zn(flrx)(phen)(H2O)]·NO3·2H2O. (c) PXRD pattern of [Zn(flrx)(phen)
(H2O)]·NO3·2H2O.

Fig. 2(c) shows the PXRD result. The simulated PXRD pattern based Fluorescence spectra are used to provide information on the binding
on the single-crystal X-ray result and the measured PXRD contrasts to methods of complexes to ctDNA. Fig. 3 (a) reveals a decrease in fluor-
the seam coordinate system. The diagnostic peaks were at 6.05°, 9.15°, escence intensity of the complex when the ctDNA was titrated into a
17.81°, 21.82°, 22.74°, 25.47°, and 29.58°. The two patterns obviously fixed concentration of the complex. The quenching mechanisms can be
matched well, which indicates that the single-crystal X-ray information determined by the distinctions in behavior depending on temperature
was authentic. and calculated by the Stern–Volmer equation [26]:
The DAD signal shown in Fig. A2 was detected at 286 nm based on
F0 / F = 1 + K SV [Q] (1)
the absorption spectra (Fig. A3). Evidently, the complex peak
(tr = 16.775 min) appears at different time from flrx (tr = 2.244 min) The binding constants were also computed with the following fitting
and phen (tr = 12.101 min). It is reasonable to believe that the complex formula [21]:
still maintains a stable complexing form in aqueous solution. Certainly, lg [(F0 F )/ F ] = lg K a + n lg [Q] (2)
the above conclusion can be reflected through that the synthesis of the
complex also happened in solution. The stability of the complex in where F0 and F represent the fluorescence intensities when the ctDNA is
solution provides premise for following research. The solubility of the absent and present, respectively; [Q] represents the concentration of
complex in water is 2.6275 ± 0.0102 mM at 310 K. the complex; KSV is the Stern–Volmer quenching constant, obtained
through the linear regression plot of F0/F against [Q]; Ka is the binding

Fig. 3. (a) Fluorescence spectra of [Zn(flrx)(phen)(H2O)]·NO3·2H2O (25 μM) in the absence and presence of ctDNA at 298 K. DNA concentrations from a to h: 0, 12.5,
25.0, 37.5, 50.0, 62.5, 75.0, and 87.5 μM. (b) Stern–Volmer plots for complexes interacting with ctDNA at three different temperatures. (c) Plot of log[(F0–F)/F]
versus log[Q] for the complex–ctDNA system at three different temperatures. (d) Van't Hoff plot of the complex–ctDNA system.

8
Y. Zhu, et al. Inorganic Chemistry Communications 103 (2019) 6–11

Table 2
Binding constants, quenching constants, and thermodynamic parameters of three temperatures.
T(K) KSV × 103 (M−1) Ka × 103 (M−1) n ΔG (KJ·mol−1) ΔH (KJ·mol−1) ΔS (J·mol−1·K−1)

298 2.151 ± 0.165 0.681 ± 0.155 0.872 −16.372 96.767 379.659


304 1.829 ± 0.056 1.907 ± 0.419 1.010 −18.650
310 1.586 ± 0.127 2.898 ± 0.227 1.073 −20.928

constant; and n is the number of binding sites [27]. The linear regres- of static quenching involves the formation of a static ground-state
sion diagrams of three temperatures (298 K, 304 K, and 310 K) are complex, which must maintain the complex's decay time constant. This
displayed in Table 2. observation is consistent with the previous conclusion.
KSV obviously decreased with increased temperature for the CD is considered as a credible tool for studying the structural
complex–DNA system; therefore, the quenching mechanism can be changes. The characteristic peaks of DNA located in 245 and 275 nm
considered as static quenching because of their dependence on the were attributed to the helicity and base stacking effect. No change in
generation of a ground-state complex. Every value of n is approximately position and strength was noted regardless of the presence and absence
1, which suggests that the complex tends to bind to a single site in of the complex (Fig. A5 (a)). This result suggested that the structure of
ctDNA. To estimate the binding force of the complex–ctDNA system DNA was almost unchanged by the binding of the ligand. Hence, small
preliminarily, the Van't Hoff equation is run to obtain these thermo- molecules may well bind to DNA in the groove [33]. While comparing
dynamic parameters, including ΔH (enthalpy change), ΔS (entropy with intercalative binding, groove binding has slighter impact on the
change), and ΔG (Gibbs' free energy change): structure of DNA.
The above inference was also proven by iodide quenching studies.
ln K a = ( H / T + S )/R (3)
The iodide and DNA phosphate backbone both carry the negative
G= H T S (4) charge, which brought repulsion when iodide was in contact with the
DNA phosphate backbone. When small molecules intercalated into the
where Ka is the binding constant and R is the gas constant. The calcu- DNA base pairs, they were protected away from an ionic quencher. In
lations are exhibited in Table 2. The values of Ka are in the range of the groove mode, DNA only provided limited protection to small mo-
102–103 M−1, which are lower than that of many other reported Zn(II) lecule fluorescence, which led to an easier quenching by KI than in the
complexes-DNA systems ranging from 104 to 106 M−1 (Table A3) intercalation mode [34]. Fig. A5(b) shows clearly that the KSV of the
[28–30]. Intercalative binding generally produce a larger binding complex–KI system (15.57 M−1) with DNA was slightly greater than
constant than groove binding. The low binding affinity illustrates that that (14.95 M−1) in the absence of DNA. The value of KSV was obtained
the complex shows weak binding with DNA and the binding mode through the slope of the linear fit. This test also demonstrated that the
should be non-intercalative. So the binding of the complex resulted in binding mode of the complex was groove binding.
very limited conformation changes in DNA, and did not lead to sus- Afterward, the DNA melting temperature Tm was also measured to
taining bioaccumulative toxicity in mammalian cells. All the values of serve as evidence of groove binding. Tm represents the temperature at
ΔG were less than 0; thus, the binding of this system is spontaneous. which the H bonds between DNA base pairs break and allow half of the
Moreover, the magnitude of ΔH of all intercalators was negative be- double-stranded DNA to transform to single-stranded DNA. In Fig.
cause they altered the DNA conformations [30]. On the contrary, the A5(c), the Tm performs as the temperature corresponding to the UV
positive enthalpy indicated that the complex hardly altered the DNA absorption reached half of the maximum. The calculated Tm are given in
conformation. And ΔH and ΔS are both positive, which can be used to Table 4. We observed that Tm remained almost the same in the absence
demonstrate that hydrophobic force plays a key role in the binding of and presence of complex. As reported, the intercalation into the double
complexes to ctDNA [31,32]. As shown in Table A3, in other zinc helix would make the DNA helix become compact so as to increase Tm,
complex and DNA interaction systems, the dominant role is mainly while groove binding did minimal effect on transitions in the DNA
hydrogen bond, a stronger intermolecular force, which should be one of double helix conformation [35]. Therefore, the DNA melting tempera-
the primary reasons why the binding of this complex and DNA is ture test provides strong evidence for the groove binding mode of this
weaker than that of other systems. system.
To further verify the quenching mechanism, the time-resolved The docking exploration was carried by the YASARA software to
fluorescence lifetime was measured, which is the most reliable method provide information about binding mode, binding site and binding
to distinguish the mechanism between static and dynamic quenching at force type between complex and ctDNA. We further identified the
present. The following equation is used to obtain the average fluores- preponderant binding mode as groove binding as suggested in Fig. 4.
cence lifetime ({τ}). The binding model shown in Fig. 4 is derived by the Chimera pro-
= 1 1 + 2 2 (5) gram from the top-rated result in accordance with the YASARA scores.
Then, we obtained a 2D map by LigPlot. As revealed, the complex links
Fig. A4 shows the fluorescence lifetime curve of the complex. As to the DG10, DG12, DA17, DA18, and DT19 bases of DNA by hydro-
shown in Table 3, the average lifetime of the free complex was phobic interaction. Meanwhile, hydrophobic force also existed between
0.5152 ns. After DNA addition, the average lifetime of the system was the small molecule and adjacent backbone phosphodiester groups. In
almost unchanged with values of 0.5211 and 0.5289 ns. The mechanism

Table 3 Table 4
Fluorescence decay fitting parameters for the complex–ctDNA system with DNA melting temperatures in the presence and absence of
different concentrations of ctDNA. complex.
Sample τ1 (ns) α1 τ2 (ns) α2 τ (ns) χ2 System Tm

Free complex (0.5 μM) 0856 0.507 1.697 0.048 0.515 1.061 Free DNA 64.869
Complex:ctDNA (1:2) 0.849 0.498 1.640 0.060 0.521 1.095 ctDNA:Complex = 1: 0.5 64.874
Complex:ctDNA (1:3.5) 0.874 0.513 1.785 0.045 0.529 1.065 ctDNA:Complex = 1: 1 65.288

9
Y. Zhu, et al. Inorganic Chemistry Communications 103 (2019) 6–11

Fig. 4. Molecular modeling results of the energy-minimized structure of the complex–DNA system (left) and the 2D map of interaction calculated using LigPlot
(right).

conclusion, the hydrophobic force plays the leading role in the com- scientific Instruments Sharing Platform Ability Construction of Sichuan
plex–DNA system, which was identified with the result from fluores- Province (Grant No. 2016KJTS0037).
cence research.
Appendix A. Supplementary material
3. Conclusion
Supplementary data to this article can be found online at https://
In this study, a new metal-based compound of Zn(II) was synthe- doi.org/10.1016/j.inoche.2019.02.039.
sized with the quinolone antibacterial drug flrx as main ligand and
phen as auxiliary ligand. The crystal structure was revealed as space References
group p-1 of triclinic. The structure chart showed that the central Zn(II)
coordinated with two nitrogen atoms from phen, one pyridone oxygen, [1] V.A. Munoz, G.V. Ferrari, M.P. Montana, S. Miskoski, N.A. Garcia, Effect of Cu(2+)-
and one carboxylate oxygen atom from the deprotonated flrx. This complexation on the scavenging ability of chrysin towards photogenerated singlet
molecular oxygen (O2((1)Deltag)). Possible biological implications, J. Photochem.
structure stabilizes the compound. In the binding of the complex to Photobiol. B 162 (2016) 597–603, https://doi.org/10.1016/j.jphotobiol.2016.07.
ctDNA, the quenching mechanism was considered in a static manner, 027.
and the complex combined with the groove region of DNA through [2] N. Muhammad, Z. Guo, Metal-based anticancer chemotherapeutic agents, Curr.
Opin. Chem. Biol. 19 (2014) 144–153, https://doi.org/10.1016/j.cbpa.2014.02.
hydrophobic force primarily. This research indicated that the 003.
Zn–flrx–phen complex would cause incredibly tiny damage to mam- [3] W.H. El-Shwiniy, W.A. Zordok, Synthesis, spectral, DFT modeling, cytotoxicity and
malian cell DNA. Given its stability and low genetic toxicity, the microbial studies of novel Zr(IV), Ce(IV) and U(VI) piroxicam complexes,
Spectrochim. Acta A Mol. Biomol. Spectrosc. 199 (2018) 290–300, https://doi.org/
Zn–flrx–phen complex may be used as a metallodrug. 10.1016/j.saa.2018.03.074.
[4] J.R. Anacona, K. Mago, J. Camus, Antibacterial activity of transition metal com-
Conflict of interest plexes with a tridentate NNO amoxicillin derived Schiff base. Synthesis and char-
acterization, Appl. Organomet. Chem. 32 (7) (2018), https://doi.org/10.1002/aoc.
4374.
The authors declare no conflicts of interest. [5] V.A. Munoz, G.V. Ferrari, M.P. Montana, S. Miskoski, N.A. Garcia, Effect of cu(2+)-
complexation on the scavenging ability of chrysin towards photogenerated singlet
molecular oxygen (O2((1)Deltag)). Possible biological implications, J. Photochem.
Acknowledgements
Photobiol. B 162 (2016) 597–603, https://doi.org/10.1016/j.jphotobiol.2016.07.
027.
This work was supported by Sichuan Science and Technology [6] M. Bochmann, B. Bertrand, M.R.M. Williams, Gold(III) complexes for anti-tumour
Program (Grant No. 2018JY0188), the Fundamental Research Funds for applications: an overview, Chemistry. (2018), https://doi.org/10.1002/chem.
201800981.
the Central Universities (Grant No. 2018SCU12043), and the Large-

10
Y. Zhu, et al. Inorganic Chemistry Communications 103 (2019) 6–11

[7] S. Zehra, S. Tabassum, H.A. Al-Lohedan, F. Arjmand, A zwitterionic Zn(II) ben- complexes containing quinazoline and 1, 10-phenanthroline as mixed ligands, J.
zothiazole nanohybrid conjugate as hydrolytic DNA cleavage agent, Inorg. Chem. Lumin. 203 (2018) 234–246, https://doi.org/10.1016/j.jlumin.2018.06.058.
Commun. 93 (2018) 69–72, https://doi.org/10.1016/j.inoche.2018.05.008. [23] D.K. Saha, S. Padhye, C.E. Anson, A.K. Powell, Antimycobacterial activity of mixed-
[8] R.N. Jadeja, M. Chhatrola, V.K. Gupta, Zn(II) coordination compounds derived from ligand copper quinolone complexes, Transit. Met. Chem. 28 (5) (2003) 579–584,
4-acyl pyrazolones and 1,10 phenanthroline: syntheses, crystal structures, spectral https://doi.org/10.1023/a:1025050910811.
analysis and DNA binding studies, Polyhedron 63 (2013) 117–126, https://doi.org/ [24] I. Yousuf, M. Usman, M. Ahmad, S. Tabassum, F. Arjmand, Single X-ray crystal
10.1016/j.poly.2013.07.018. structure, DFT studies and topoisomerase I inhibition activity of a tailored ionic ag
[9] H. Wu, J. Yuan, Y. Bai, G. Pan, H. Wang, J. Shao, ... Y. Wang, Synthesis, crystal (i) nalidixic acid–piperazinium drug entity specific for pancreatic cancer cells, New
structure, DNA-binding properties, and antioxidant activity of a V-shaped ligand bis J. Chem. 42 (1) (2018) 506–519, https://doi.org/10.1039/c7nj03602g.
(N-methylbenzimidazol-2-ylmethyl)benzylamine and its zinc(II) complex, J. Coord. [25] E.K. Efthimiadou, M.E. Katsarou, A. Karaliota, G. Psomas, Copper(II) complexes
Chem. 65 (24) (2012) 4327–4341, https://doi.org/10.1080/00958972.2012. with sparfloxacin and nitrogen-donor heterocyclic ligands: structure–activity re-
741229. lationship, J. Inorg. Biochem. 102 (4) (2008) 910–920, https://doi.org/10.1016/j.
[10] P.C. Bruijnincx, P.J. Sadler, New trends for metal complexes with anticancer ac- jinorgbio.2007.12.011.
tivity, Curr. Opin. Chem. Biol. 12 (2) (2008) 197–206, https://doi.org/10.1016/j. [26] M. Gopalakrishnan, K. Senthilkumar, P.R. Rao, R. Siva, N. Palanisami, Synthesis,
cbpa.2007.11.013. crystal structure, DNA binding and molecular docking studies on new copper(II)
[11] T. Storr, K.H. Thompson, C. Orvig, Design of targeting ligands in medicinal in- salicylate [Cu(DTBSA)2(2,2′-bpy)](dmf), Inorg. Chem. Commun. 46 (46) (2014)
organic chemistry, Chem. Soc. Rev. 35 (6) (2006) 534–544, https://doi.org/10. 54–59, https://doi.org/10.1016/j.inoche.2014.03.043.
1039/b514859f. [27] D. Wu, Z. Chen, Study on the interaction between ginsenoside Rh2 and calf thymus
[12] C.M. Manzano, F.R.G. Bergamini, W.R. Lustri, A.L.T.G. Ruiz, E.C.S. de Oliveira, DNA by spectroscopic techniques, Luminescence 30 (8) (2015) 1212–1218, https://
M.A. Ribeiro, ... P.P. Corbi, Pt(II) and Pd(II) complexes with ibuprofen hydrazide: doi.org/10.1002/bio.2883.
characterization, theoretical calculations, antibacterial and antitumor assays and [28] N. Shahabadi, F. Shiri, M. Norouzibazaz, A. Falah, Disquisition on the interaction of
studies of interaction with CT-DNA, J. Mol. Struct. 1154 (2018) 469–479, https:// ibuprofen–Zn(II) complex with calf thymus DNA by spectroscopic techniques and
doi.org/10.1016/j.molstruc.2017.10.072. the use of Hoechst 33258 and methylene blue dyes as spectral probes, Nucleosides
[13] T. Zou, C.N. Lok, P.K. Wan, Z.F. Zhang, S.K. Fung, C.M. Che, Anticancer metal-N- Nucleotides & Nucleic Acids 37 (3) (2018) 125–146, https://doi.org/10.1080/
heterocyclic carbene complexes of gold, platinum and palladium, Curr. Opin. Chem. 15257770.2017.1400048.
Biol. 43 (2018) 30–36, https://doi.org/10.1016/j.cbpa.2017.10.014. [29] K. Sakthikumar, J.D. Raja, R.V. Solomon, M. Sankarganesh, Density functional
[14] R. Singh, R.N. Jadeja, M.C. Thounaojam, T. Patel, R.V. Devkar, D. Chakraborty, theory molecular modelling, DNA interactions, antioxidant, antimicrobial, antic-
Synthesis, DNA binding and antiproliferative activity of ternary copper complexes ancer and biothermodynamic studies of bioactive water soluble mixed ligand
of moxifloxacin and gatifloxacin against lung cancer cells, Inorg. Chem. Commun. complexes, J. Biomol. Struct. Dyn. (2018), https://doi.org/10.1080/07391102.
23 (27) (2012) 78–84. 2018.1492970.
[15] M.J. Suto, J.M. Domagala, G.E. Roland, G.B. Mailloux, M.A. Cohen, [30] M. Sedighipoor, A.H. Kianfar, M.R. Sabzalian, F. Abyar, Synthesis and character-
Fluoroquinolones: relationships between structural variations, mammalian cell cy- ization of new unsymmetrical Schiff base Zn (II) and Co (II) complexes and study of
totoxicity, and antimicrobial activity, J. Med. Chem. 35 (25) (1992), https://doi. their interactions with bovin serum albumin and DNA by spectroscopic techniques,
org/10.1021/jm00103a013. Spectrochimica Acta Part A Molecular & Biomolecular Spectroscopy 198 (2018)
[16] C.X. Zhang, S.J. Lippard, New metal complexes as potential therapeutics, Curr. 38–50, https://doi.org/10.1016/j.saa.2018.02.050.
Opin. Chem. Biol. 7 (4) (2003) 481–489, https://doi.org/10.1016/s1367-5931(03) [31] S. Kashanian, M.M. Khodaei, H. Roshanfekr, N. Shahabadi, G. Mansouri, DNA
00081-4. binding, DNA cleavage and cytotoxicity studies of a new water soluble copper(II)
[17] F. Ahmadi, N. Ebrahimi-Dishabi, K. Mansouri, F. Salimi, Molecular aspect on the complex: the effect of ligand shape on the mode of binding, Spectrochim. Acta A
interaction of zinc-ofloxacin complex with deoxyribonucleic acid, proposed model Mol. Biomol. Spectrosc. 86 (2012) 351–359, https://doi.org/10.1016/j.saa.2011.
for binding and cytotoxicity evaluation, Research in Pharmaceutical Sciences 9 (5) 10.048.
(2014) 367–383. [32] N. Shakibapour, F. Dehghani Sani, S. Beigoli, H. Sadeghian, J. Chamani, Multi-
[18] Q. Sun, Z. Suo, H. Pu, P. Tang, N. Gan, R. Gan, ... H. Li, Studies of the binding spectroscopic and molecular modeling studies to reveal the interaction between
properties of the food preservative thiabendazole to DNA by computer simulations propyl acridone and calf thymus DNA in the presence of histone H1: binary and
and NMR relaxation, RSC Adv. 8 (36) (2018) 20295–20303, https://doi.org/10. ternary approaches, J. Biomol. Struct. Dyn. (2018), https://doi.org/10.1080/
1039/c8ra03702g. 07391102.2018.1427629.
[19] M.T. Alghamdi, A.A. Alsibaai, M.S. Shahawi, M.S. Refat, Synthesis and spectro- [33] W. Huang, S. Kong, Z. Wang, C. Pan, H. Zhu, Ni(II) ternary complex based on an-
scopic studies of levofloxacin uni-dentate complexes of Ru(II), Pt(IV) and Ir(III): timicrobial drug Enoxacin: synthesis and biological properties, Chin. J. Chem. 32
third generation of quinolone antibiotic drug complexes, J. Mol. Liq. 224 (2016) (11) (2014) 1169–1175, https://doi.org/10.1002/cjoc.201400420.
571–579, https://doi.org/10.1016/j.molliq.2016.10.038. [34] Z. Aramesh-Boroujeni, M. Khorasani-Motlagh, M. Noroozifar, Multispectroscopic
[20] K.D. Mjos, J.F. Cawthray, E. Polishchuk, M.J. Abrams, C. Orvig, Gallium(iii) and DNA-binding studies of a terbium(III) complex containing 2,2-bipyridine ligand, J.
iron(iii) complexes of quinolone antimicrobials, Dalton Trans. 45 (33) (2016) Biomol. Struct. Dyn. 34 (2) (2016) 414–426, https://doi.org/10.1080/07391102.
13146–13160, https://doi.org/10.1039/c6dt01315e. 2015.1038585.
[21] Y. Xiao, Q. Wang, Y. Huang, X. Ma, X. Xiong, H. Li, Synthesis, structure, and bio- [35] M. Bordbar, F. Tavoosi, A. Yeganeh-Faal, M.H. Zebarjadian, Interaction study of
logical evaluation of a copper(ii) complex with fleroxacin and 1,10-phenanthroline, some macrocyclic inorganic schiff base complexes with calf thymus DNA using
Dalton Trans. 45 (27) (2016) 10928–10935, https://doi.org/10.1039/c6dt00915h. spectroscopic and voltammetric methods, J. Mol. Struct. 1152 (2018) 128–136,
[22] L.Q. Chai, Q. Hu, K.Y. Zhang, L. Zhou, J.J. Huang, Synthesis, structural char- https://doi.org/10.1016/j.molstruc.2017.09.088.
acterization, spectroscopic, and DFT studies of two penta-coordinated zinc(II)

11

Das könnte Ihnen auch gefallen