SlideShare a Scribd company logo
Polyethyleneimine anchored copper(II) complexes: Synthesis,
characterization, in vitro DNA binding studies and cytotoxicity studies
Jagadeesan Lakshmipraba a
, Sankaralingam Arunachalam a,⇑
, Anvarbatcha Riyasdeen b
,
Rajakumar Dhivya c
, Mohammad Abdulkader Akbarsha b
a
School of Chemistry, Bharathidasan University, Tiruchirappalli 620 024, Tamil Nadu, India
b
Mahatma Gandi-Doerenkamp Centre, Bharathidasan University, Tiruchirappalli 620 024, Tamil Nadu, India
c
Department of Biomedical Science, Bharathidasan University, Tiruchirappalli 620 024, Tamil Nadu, India
a r t i c l e i n f o
Article history:
Received 18 August 2014
Received in revised form 15 November 2014
Accepted 19 November 2014
Available online 27 November 2014
a b s t r a c t
The water soluble polyethyleneimine–copper(II) complexes, [Cu(phen)(L-tyr)BPEI]ClO4 (where
phen = 1,10-phenanthroline, L-tyr = L-tyrosine and BPEI = branched polyethyleneimine) with various
degree of copper(II) complex units in the polymer chain were synthesized and characterized by elemental
analysis and electronic, FT-IR, EPR spectroscopic techniques. The binding of these complexes with CT-
DNA was studied using UV–visible absorption titration, thermal denaturation, emission, circular dichro-
ism spectroscopy and cyclic voltammetric methods. The changes observed in the physicochemcial prop-
erties indicated that the binding between the polymer–copper complexes and DNA was mostly through
electrostatic mode of binding. Among these complexes, the polymer–copper(II) complex with the highest
degrees of copper(II) complex units (higher degrees of coordination) showed higher binding constant
than those with lower copper(II) complex units (lower degrees of coordination) complexes. The complex
with the highest number of metal centre bound strongly due to the cooperative binding effect. Therefore,
anticancer study was carried out using this complex. The cytotoxic activity for this complex on MCF-7
breast cancer cell line was determined adopting MTT assay, acridine orange/ethidium bromide (AO/EB)
staining and comet assay techniques, which revealed that the cells were committed to specific mode
of cell death either apoptosis or necrosis.
Ó 2014 Elsevier B.V. All rights reserved.
1. Introduction
During the past decade, there has been tremendous interest in
the synthesis and studies pertaining to the interaction of various
metal complexes with DNA and also screening of these complexes
for their anticancer activities so as to replace cisplatin [1–5].
Hence, much attention has been targeted on the design of metal-
based complexes, which can bind to DNA. Interaction of metal
complexes with DNA should be useful in the development of
molecular probes and new therapeutic agents.
Cancer research mainly targets the DNA molecule which is the
origin of uncontrolled cell division. To block this cell division, there
is a need for developing DNA targeted chemotherapy drugs. Sev-
eral small molecules/metal complexes have been interacted with
DNA to correlate their effects of binding/cleavage behavior on cyto-
toxicity [6,7]. In these aspects of drug designing, some of the issues
like solubility, transfection efficiency, targeted delivery, etc. may
enter as practical problems. To overcome these problems, drug
carriers like cationic polymers, surfactants, liposomes and dentri-
mers were employed for efficient drug delivery [8].
Copper is a physiologically important metal element that plays
an important role in the endogenous oxidative DNA damage asso-
ciated with aging and cancer [9]. Copper(II) complexes bearing
1,10-phenanthroline ligand have been widely used due to their
high nucleolytic efficiency [10] and numerous biological activities
such as antitumor, anti-candida and antimicrobial activities
[11,12]. It has been reported that a binuclear copper(II) complex
containing 1,10-phenanthroline and a trinuclear copper(II) com-
plex containing di-(2-picolyl)amine bind strongly with DNA and
cleave more effectively than their corresponding monomeric com-
plexes [13–15]. The reports in the literature that advocate design
studies on metal complexes that cooperative effect arising from
such non-covalent interactions would be a valuable principle in
the development of new metal based probes which recognize bio-
molecular targets with high specificity [16].
Recent literature indicate that mixed ligand copper(II) com-
plexes have been receiving considerable attention for various
http://dx.doi.org/10.1016/j.jphotobiol.2014.11.005
1011-1344/Ó 2014 Elsevier B.V. All rights reserved.
⇑ Corresponding author. Fax: +91 431 2407043.
E-mail address: arunasurf@yahoo.com (S. Arunachalam).
Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67
Contents lists available at ScienceDirect
Journal of Photochemistry and Photobiology B: Biology
journal homepage: www.elsevier.com/locate/jphotobiol
reasons. Copper(II) complexes having amino acid ligand are of
interest due to their biological relevance, good DNA binding ability,
antimicrobial and anticancer activities [17]. Also, there are reports
on drug polymer conjugates as potential candidates for the selec-
tive delivery of anticancer agents to tumor tissues. Particularly,
polyethyleneimine (PEI), possesses quite a number of advantages
as polymer chelating agent, such as good water solubility, high
content of functional groups, suitable molecular weights as well
as good physical and chemical stabilities [18]. Stable polyethylene-
imine–copper(II) complexes are reported where copper ions are
hard to elute from the polymer domain to the bulk solution and
binding constant values are 1–5 order magnitude greater in the
polymer–chelate systems than in the monomeric Cu-complex
systems [19].
Our research group has been involved in the synthesis and stud-
ies on mixed polymer–copper(II) complexes and their DNA binding
properties for the past several years [20,21]. We have reported syn-
thesis and nucleic acid binding of polyethyleneimine–copper(II)
complexes with L-valine and L-arginine as a co-ligand [22,23]. In
the present work we report the synthesis, DNA binding and antitu-
mor properties of polyethyleneimine–copper(II) complexes con-
taining L-tyrosine as co-ligand. L-tyrosine contains benzene ring
which can influence the binding of these complexes with DNA
through p–p interactions.
2. Experimental section
2.1. Materials
Calf thymus DNA and branched polyethyleneimine (BPEI) (Mw
ca. 25,000) were obtained from Sigma–Aldrich, Germany and were
used as obtained. Copper(II) chloride dihydrate, 1,10-phenanthro-
line (Merck, India) and tyrosine (Loba Chemie, India) and were used
as received. The precursor complex, [Cu(phen)(L-tyr)(H2O)]ClO4,
was prepared as reported earlier [24]. A solution of calf thymus
DNA in the buffer gave a UV absorbance ratio of $1.8–1.9: 1 at 260
and 280 nm, indicating that the DNA was sufficiently free of protein.
The concentration of CT DNA in base pairs was determined by UV
absorbance at 260 nm by taking the molar extinction coefficient
value 13,200 MÀ1
cmÀ1
for DNA at 260 nm [25,26]. All the experi-
ments involving the interaction of the polymer–copper(II) complex
with DNA were carried out using buffer containing 5 mM Tris–HCl/
50 mM NaCl at pH 7.0 in twice distilled water. MCF-7 cell line was
obtained from National Centre For Cell Science (NCCS), Pune.
2.2. [Cu(phen)(L-tyr)BPEI]ClO4
To a solution of branched polyethyleneimine (BPEI) (0.15 g,
3.4 mmol) dissolved in ethanol (15 ml), [Cu(phen)(L-tyr)
(H2O)]ClO4 (0.8 g, 1.4 mmol) in water was added slowly with
stirring. The mixture was heated between 50 and 60 °C for 15 h
in a water bath with stirring. The resulting dark blue solution
was dialyzed at 15 °C against distilled water for 4–5 days. The
solvent was then evaporated in a rotary evaporator under reduced
pressure at room temperature. The dark-bluish filmy substance
obtained was pulverized and dried. Yield = 0.23 g for x = 0.182.
(Anal. Calc.: C, 33.77, H, 4.40, N, 14.50, O, 5.81, Cu 18.35, Found:
H, 5.02, N, 14.25, O, 5.77, Cu 18.37% (x = 0.182 obtained from
carbon content). IR (KBr, cmÀ1
): m(NAH) 3446, m(CAC) 2924, m(COOA)
1099, m(C@C) 1471, m(C@N) 1390, 1099, m(CAH) 852, m(CAH) 730; UV
(kmax, nm,(e, MÀ1
cmÀ1
)): 227 (28,120) 272 (54,240), 294
(41,830), 645 (11,090). EPR (77 K, g|| 2.211 and g 2.017; RT
giso = 2.0753.
Polymer–copper(II) complex samples with various numbers of
copper(II) complex units bound to the polymer chain were synthe-
sized by changing the amount of reactants in the reaction solution,
the duration of the reaction time and the reaction temperature.
2.3. Physical measurements
Elemental analysis was determined at Sophisticated Analytical
Instrument Facility (SAIF), Lucknow, India. Absorption spectra
and thermal denaturation studies were recorded on a UV–Vis–
NIR Cary300 Spectrophotometer using cuvettes of 1 cm path
length in tris buffer solution, and emission spectra were recorded
on a JASCO FP 770 spectrofluorimeter. FT-IR spectra were recorded
on a FT-IR JASCO 460 PLUS spectrophotometer with samples pre-
pared as KBr pellets. EPR spectra were recorded on a JEOL-FA200
EPR spectrometer at room temperature and at LNT in methanol
solution. Absorption titration experiments of polymer–copper(II)
complexes in buffer (50 mM NaCl–5 mM Tris–HCl, pH 7.0) were
performed by using a fixed complex concentration to which incre-
ments of the DNA stock solutions were added. Polymer–copper(II)
complex-DNA solutions were incubated for 10 min before the
absorption spectra were recorded. Equal amount of DNA was
added to both the complex and reference solutions to eliminate
the absorbance of DNA itself.
For fluorescence quenching experiments DNA was pretreated
with ethidium bromide (EB) for 30 min. The polymer–copper(II)
complexes were then added to this mixture and their effect on
the emission intensity was measured. Samples were excited at
450 nm and emission was observed between 500 and 700 nm. Cir-
cular dichroism spectra were recorded at room temperature using
the same tris buffer. For the cyclic voltammetry experiments, the
electrode surfaces were freshly polished with alumina powder
and then sonicated in ethanol and distilled water for 1 min prior
to each experiment. Then the electrodes were rinsed throughly
with distilled water. Cyclic voltammetric experiments were per-
formed at 25.0 ± 0.2 °C in a single compartment cell with a three-
electrode configuration (glassy carbon working electrode,
platinum wire auxiliary electrode and saturated calomel reference
electrode). The solution was deoxygenated with nitrogen gas for
20 min prior to experiments.
2.4. Cell culture
The MCF-7 cancer cells were cultured in RPMI 1640 medium
(Sigma–Aldrich, St. Louis, MO, USA), supplemented with 10% fetal
bovine serum (Sigma, USA) and 10,000 IU of penicillin and
100 lg mlÀ1
of streptomycin as antibiotics (Himedia, Mumbai,
India), in 96 well culture plates, at 37 °C, in a humidified atmo-
sphere of 5% CO2, in a CO2 incubator (Forma, Thermo Scientific,
USA). All the experiments were performed using cells from passage
15 or less.
2.4.1. Cytotoxicity assay (MTT assay)
The polymer–copper(II) complex of the highest degree of coor-
dination was dissolved in DMSO, diluted in culture medium and
used to treat the model cell line over a complex concentration
range of 3–30 lg mlÀ1
for a period of 24 h and 48 h. DMSO at
0.5% concentration in the culture medium was used as negative
control. We have used cisplatin as positive control. A miniaturized
viability assay using 3-(4,5-di-methylthiazol-2-yl)-2,5-diphenyl-
2H-tetrazolium bromide (MTT) (Sigma, USA) (5 mg/ml in Phos-
phate-Buffered Saline (PBS)) was added to each well and the plates
were wrapped with aluminum foil and incubated at 37 °C for 4 h.
By this treatment a purple formazone product was formed due to
the reduction of MTT by the mitochondrial enzyme succinate
dehydrogenase of the cells [27]. The purple formazan product
was dissolved by addition of 100 ll of 100% DMSO to each well.
The absorbance was monitored at 570 nm (measurement) and
60 J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67
630 nm (reference) using a 96 well plate reader (Bio-Rad, Hercules,
CA, USA). Data were collected for four replicates each and used to
calculate the respective means. The percentage of inhibition was
calculated, from this data, using the formula:
The IC50 value was determined as the complex drug concentra-
tion that is required to reduce the absorbance to half that of the
control.
2.4.2. Acridine orange (AO) and ethidium bromide (EB) staining
Acridine orange/ethidium bromide staining was performed as
follows: the cell suspension of each sample containing 5 Â 105
cells, was treated with 25 ll of AO and EB solution (1 part of
100 lg mlÀ1
AO and 1 part of 100 lg mlÀ1
EB in PBS) and examined
in a fluorescent microscope (Carl Zeiss, Germany) using an UV filter
(450–490 nm). Three hundred cells per sample were counted in
tetraplicate for each dose point. The cells were scored as viable,
apoptotic or necrotic as judged by the staining, nuclear morphol-
ogy and membrane integrity [28], and the percentages of apoptotic
and necrotic cells were then calculated. Morphological changes
were also observed and photographed.
2.4.3. Comet assay
DNA damage was detected by adopting comet assay as reported
earlier [29]. Cells were suspended in low-melting-point agarose in
PBS and pipetted out to microscope slides pre-coated with a layer
of normal-melting-point agarose. Slides were chilled on ice for
10 min and then immersed in lysis solution (2.5 M NaCl, 100 mM
Na2EDTA, 10 mM Tris, 0.2 mM NaOH, pH 10.01 and Triton X-100)
and the solution was kept for 4 h at 4 °C in order to lyse the cells
and to permit DNA unfolding. Thereafter, the slides were exposed
to alkaline buffer (300 mM NaOH, 1 mM Na2EDTA, pH > 13) for
20 min to allow DNA unwinding. The slides were washed with buf-
fer (0.4 M Tris, pH 7.5) to neutralize excess alkali and to remove
detergents before staining with ethidium bromide (5 ll in
10 mg mlÀ1
). Photographs were obtained using the fluorescent
microscope. One hundred and fifty cells from each treatment group
were digitalized and analyzed using CASP software. The images
were used to estimate the DNA content of individual nuclei and
to evaluate the degree of DNA damage representing the fraction
of total DNA in the tail.
3. Results and discussion
3.1. Degree of coordination
The structure of the water soluble polymer–copper(II) complex
is shown in Fig. 1. In this figure ‘x’ represents the degree of coordi-
nation, which is the number of moles of copper(II) chelate per mole
of the repeating unit (amine group) of polymeric ligand. If the
entire repeating units (amine group) in the polymer are coordi-
nated to copper, then the value of x is 1. The degree of coordination
(x) was calculated from carbon content value [19,30,31]. The
degree of coordination thus obtained for the polymer–copper(II)
complex samples of the present work are 0.059, 0.149, 0.182.
The stability of the polymer–copper(II) complexes in solution
was verified occasionally by keeping the solution of the polymer–
copper(II) complexes in dialysis bags, small amount of the solution
from the dialysis bags was taken and the stability was confirmed
through absorption spectroscopy. Any free copper complex ion or
copper ion in the solution outside the dialysis bag was not observed
(which was verified using spectrophotometric method), indicating
that polymer–complexes were very stable during the handling of
our experiments. Also viscosity, is a significant experiment which
gives information about interaction in aqueous solution. In our
case, even after dialysis we do not have change in viscosity. This
proves that the complexes are very stable in solution.
3.2. Characterization of polymer–copper(II) complexes
The FT-IR spectra for the polymer–copper(II) complexes dis-
played stretching frequencies around 1471 cmÀ1
and 1390 cmÀ1
which can be attributed to the ring stretching frequencies viz.,
m(C@C) and m(C@N), respectively, of the coordinated 1,10-phenan-
throline and these are at slightly lower frequencies than that of
uncoordinated 1,10-phenanthroline. The m(CAH) out-of-plane
bending values, around 852 cmÀ1
and 730 cmÀ1
, for the phenan-
throline ligand were shifted to 838 cmÀ1
and 693 cmÀ1
, respec-
tively, in the complexes. These shifts can be explained by the fact
that each of the two nitrogen atoms of phenanthroline ligands
donates a pair of electrons to the central copper atom forming
coordinate bond [32]. The band obtained around 2924 cmÀ1
can
be assigned to CAC stretching vibration of aliphatic CH2 of BPEI
whereas the broad band observed around 3446 cmÀ1
can be
assigned to the NAH stretching of BPEI [33]. The uncoordinated
amino acid exhibited a stretch in the region 1750–1700 cmÀ1
cor-
responding to m(COOH). In the complexes, this band was shifted to
1630 cmÀ1
indicating the coordination of carboxylate group to the
copper(II) ion. The very strong band around 1099 cmÀ1
can be
assigned to the presence of perchlorate anion. The stretching fre-
quencies around 508 cmÀ1
and 491 cmÀ1
can be attributed, respec-
tively, to copper–nitrogen and copper–oxygen stretching.
The UV–visible absorption spectra of all the complexes were
recorded in the region 200–800 nm. All the complexes displayed
four bands in the regions 230–645 nm. In the UV region, the
absorption bands below 300 nm are attributed to intra-ligand tran-
sitions whereas in the visible region, the band around 645 nm is
assigned as d–d transition.
The solid state EPR spectra of the polymer–copper(II) complex
(x = 0.203) was recorded in X-band frequencies at room tempera-
ture as well as in frozen solution (77 K) in methanol as solvent
(Fig. 2). The room temperature EPR spectra of the polymer–cop-
per(II) complexes showed single isotropic feature at giso = 2.070–
2.075, and this broadening of isotropic peak is due to intermolecu-
lar spin exchange. This intermolecular type of spin exchange is
caused by the strong spin coupling which occurs during a coupling
Cu
N
H
N
H
(1-x)
x
N
N
.ClO4
NH2
OH
O O
Fig. 1. Schematic representation of [Cu(phen)(L-tyr)BPEI]ClO4.
½Mean absorbance of untreated cellsðcontrolÞ À Mean absorbance of treated cellsðtestÞŠ Â 100:
Mean absorbance of untreated cellsðcontrolÞ
J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67 61
of two paramagnetic species. At liquid nitrogen temperature we
observed three peaks with the third being broad. This is because
the copper complex units have been mounted on a polymer chain
resulting in some spin–spin coupling between the copper complex
units which leads to a small amount of broadening. In liquid nitro-
gen temperature the complexes showed the values of g|| = 2.211–
2.216 and g = 2.016–2.020. The existence of g|| > g > 2.00 sug-
gests that dx
2
À dy
2
is the ground state with the d9
(Cu2+
) configu-
ration and square pyramidal geometry.
3.3. Absorption studies
Electronic absorption spectroscopy is an effective method to
examine the binding mode of DNA with polymer–copper(II) metal
complexes. Thus, in order to provide evidence for the binding of
polymer–copper(II) complexes to DNA, the binding process was
monitored by absorption spectroscopy by following the changes
in absorption band intensity and its position. On addition of
DNA, the absorption spectra of polymer–copper(II) complex
showed hyperchromism and slight red shift (Fig. 3). The experi-
mental results derived from the UV–visible titration experiments
suggest that positively charged polymer-complexes can bind to
DNA, probably to the phosphate groups, by electrostatic interac-
tion resulting in the stabilization of DNA duplex. Nevertheless,
the metal complex units present in the polymer chain contain aro-
matic moieties so the binding of the complexes involving partial
intercalation of an aromatic ring between the base pairs of DNA
cannot be ruled out. From the above studies the intrinsic binding
constants (Kb) were determined from the increase of absorption
at 294 nm calculated by absorption spectral titration. In order to
compare quantitatively the binding affinity with nucleic acids
between polymer–copper(II) complexes having different degrees
of coordination, the intrinsic binding constants Kbs of the com-
plexes were determined using Eq. (4) by assuming a simple model,
in which the reaction between the nucleic acid site, P and the cop-
per complex unit of the polymer complex, D to form the nucleic
acid bound complex, PD as:
P þ D 

Kb
PD ð1Þ
Kb ¼ ½PDŠ=½PŠ½DŠ ð2Þ
where [PD], [P] and [D] represent the respective equilibrium con-
centrations of nucleic acid bound copper complex units, nucleic acid
sites in base pairs and the copper complex units of the polymer
complex.
A ¼ eD½DŠ þ ePD½PDŠ ð3Þ
CD=A À eDCD ¼ ð1=ePD À eDÞ þ 1=ðePD À eDÞKb1=½PŠ ð4Þ
where eD and ePD are the molar extinction coefficient of the free cop-
per complex units and apparent molar extinction coefficient of the
nucleic acid bound copper complex units respectively, CD total con-
centration of copper complex units and A is the experimental absor-
bance. An iterative procedure was employed as per the method
provided in Ref. [34] to arrive at the Kb values (first [D] set equal
to CD, then, once a first estimate of Kb and (ePD À eD) are obtained,
a new value of [D] was calculated and so on until convergence is
achieved). This procedure yields a binding constant value (Kb) for
each complex.
As seen from Table 1 the binding constants observed for poly-
mer–copper(II) complexes are higher those that of similar type of
simple metal complexes like [Cu(phen)(L-tyr) H2O]ClO4 (Kb = 3.75 -
 103
MÀ1
) as well as the polymer alone PEI (Kb = 1.2 MÀ1
) [35,36].
However, they are very much lower than the potential intercala-
tors like ethidium bromide (Kb, 7.0 Â 107
MÀ1
in 40 mM Tris/HCl,
pH 7.9) [37] and the partially intercalating complexes like
[Co(phen)2(dppz)]3+
(Kb = 9.09 Â 105
MÀ1
) and [Ru(imp)2(dppz)]2+
(Kb = 2.19 Â 107
MÀ1
) [38], which implies that these complexes
bind to DNA relatively less strongly than classical intercalators
and partial intercalators. Also, as seen from the Table, it was
observed that the binding constant changes with degree of coordi-
nation of copper(II) units in the polymer chain; greater the ratio of
copper(II) centres in the polymer chain, higher was the binding
constant because when one copper(II) complex unit binds with
DNA it will cooperatively act to increase the overall binding ability
of the other copper(II) complex units to DNA.
Fig. 2. EPR spectrum of [Cu(phen)(L-tyr)(BPEI)]ClO4 (x = 0.203) in methanol at
liquid nitrogen temperature.
200 300 400 500 600
0.0
0.5
1.0
0.00001 0.00002 0.00003
0.0000005
0.0000010
0.0000015
0.0000020
0.0000025
0.0000030
[DNA]/εa−εf
[DNA]
Absorbance
Wavelength, nm
Fig. 3. Absorption spectra of [Cu(phen)(L-tyr)BPEI]ClO4 (x = 0.182) in the absence of
DNA and in the presence of DNA, [complex] = 3 Â 10À5
M, [DNA] = 0–3.2 Â 10À5
M
(inset: plot of [DNA]/(ea À ef) vs. [DNA]).
Table 1
The intrinsic binding constant (Kb) of [Cu(phen)(L-tyr)BPEI]ClO4, with DNA and RNA
and thermal melting temperature in the presence of [Cu(phen)(L-tyr)BPEI]ClO4 with
different degree of coordination.
Complex Degree of
coordination (x)
Kb (MÀ1
)
±0.04
Ksv (MÀ1
)
±0.03
[Cu(phen)(L-tyr)BPEI]ClO4 0.059 2.10 Â 104
2.13 Â 104
0.149 2.03 Â 105
2.78 Â 104
0.182 7.80 Â 105
3.37 Â 104
62 J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67
3.4. Ethidium bromide displacement assay
All polymer–copper(II) complexes were non-emissive upon
excitation of the MLCT band, either in aqueous solution or in the
presence of DNA. The competitive binding experiments with a
well-established quenching assay based on the displacement of
the intercalating EB from ct-DNA was carried out in order to get
further information regarding the DNA binding properties of poly-
mer–metal complexes. The quenching of emission intensity of DNA
bound EB (Fig. 4) was analyzed through Stern–Volmer equation, I0/
I = 1 + Ksv[Q], where I0 and I are the fluorescence intensities in the
absence and presence of the complex, respectively, Ksv is the linear
Stern–Volmer constant and Q is the concentration of polymer–
copper(II) complex [39,40]. A plot of I0/I vs. [Q] was drawn and
Ksv was obtained from the ratio of slope to intercept (Table 1). As
seen from the Table, the Ksv value increases with increase in degree
of coordination of polymer–copper(II) complex. This is attributed
to the cooperative binding between copper(II) units on the same
polymer chain with DNA. This cooperative effect increases with
degree of coordination.
3.5. Effect of ionic strength
The change in fluorescence intensity of cationic copper(II)–poly-
mer complexes to DNA in the presence of NaCl can be used to verify
whether the binding mode is electrostatic or intercalative; a linear
relation between fluorescence intensity and concentration of NaCl
is highly indicative of an electrostatic mode of interaction whereas
a non-dependence of fluorescence intensity on ionic strength indi-
cates intercalation [41,42]. It is observed that as the concentration
of NaCl increases, the relative fluorescence intensity due to ethi-
dium bromide increases (Fig. 5). This is due to the competitive bind-
ing of Na+
ions to DNA which decreases the binding affinity of the
copper(II)–polymer complex to DNA. As the concentration of NaCl
increases, a linear increase of fluorescence is noticed, indicating
that the cationic copper(II)–polymer complex-DNA interactions
for the polymers studied are electrostatic. [43,44]
3.6. Circular dichroism spectral studies
CD spectral technique is useful method to monitor the
conformational variations of DNA during complex-DNA interac-
tions and achieve information on changing DNA conformation by
the binding of the metal complex to DNA. DNA has a major
longwave positive peak centred at 275 nm and the intensity of this
positive peak is similar in magnitude to that of the negative peak
centred at 245 nm (Fig. 6) corresponding to the p–p stacking of
the base pairs and right handed helicity of B-form DNA in buffer
solution [45]. Addition of polymer–copper(II) complex (x = 0.182)
to B-form DNA has been shown to induce a B to A transition, result-
ing in a CD spectrum with characteristics totally different from
those of B-form DNA; the long wave positive peak is larger with
a maximum at $270 nm and a very large shortwave peak results
below $230 nm [46]. Thus, the increased ellipticity observed at
275 nm when polymer–copper(II) complex binds to DNA can be
interpreted as unwinding of B form of DNA due to a decrease in
twist angle. This can be tentatively interpreted as B form of DNA
becoming more ‘A-like’ upon binding polymer–copper(II) complex.
3.7. Cyclic voltammetry studies
The binding of polymer–copper(II) complexes with DNA was
further confirmed by cyclic voltammetric studies. The cyclic
500 600 700
0
100
200
300
400
Intensity
Wavelength, nm
Fig. 4. Emission spectra of EB bound to DNA, [EB] = 2 Â 10À4
M in the absence of
complex and in the presence of complex (x = 0.182), [DNA] = 2 Â 10À3
M, [com-
plex] = 0–1 Â 10À4
M.
0.0 0.2 0.4 0.6 0.8
0.0
0.2
0.4
0.6
0.8
1.0
RelativeFluoresence
[NaCl] mM
1
2
3
Fig. 5. Titration of DNA[DNA] = 2 Â 10À4
M in the presence of ethidium bro-
mide[EB] = 2 Â 10À4
M in the presence of polymer–copper(II) complexes[com-
plex] = 1 Â 10À4
M as a function of NaCl concentration.
240 280 320
-80
-40
0
40
80
DNA
DNA+ complex
CD,mdeg
Wavelength, nm
Fig. 6. Circular dichroism spectra in the absence (black) and in the presence
[Cu(phen)(L-tyr)BPEI]ClO4 (x = 0.182), [complex] = 12 Â 10À5
M (red) with DNA,
[DNA] = 9 Â 10À5
M. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)
J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67 63
voltammogram of polymer–copper(II) complex (x = 0.182) in the
absence and presence of DNA is shown in Fig. 7. In the absence
of DNA, the cathodic peak potential (Epc) and the anodic peak
potential (Epa) of our complex are 339 mV and 574 mV, respec-
tively, with a large peak-to-peak separation, DEp, of 235 mV and
the ratio of cathodic to anodic peak current (ipc/ipa) is 1.05 indicat-
ing a quasi-reversible redox process [47]. The formal potential (E1/
2) which is taken as the average of Epc and Epa is 0.457 V in the
absence of DNA, whereas in the presence of DNA a negative shift
in E1/2 by 0.055 V along with increase in DEp of 30 mV has been
observed. The ipc/ipa value also increased with the increase of the
DNA concentration. Literature report [48] have pointed out that
the shift direction of electrochemical potential of metal complex,
after reacting with DNA, is related to its binding mode with DNA.
A positive shift of the peak potential and the negative shift indicat-
ing that electrostatic mode of interaction.
The dependence of cathodic current on scan rate was also inves-
tigated. In the case of our polymer–copper(II) complex the plot of
cathodic current vs. the square root of the scan rate (m1/2
), was lin-
ear for the complex alone, and the complex in the presence of DNA,
which indicates that the electrochemical process are diffusion
controlled process [48].
4. Cytotoxic assay
4.1. MTT assay
In vitro cytotoxicity of polymer–copper(II) complexes was eval-
uated by MTT assay on MCF-7 cells. The cytotoxic effects of the
polymer–copper(II) complex of the highest degree of coordination
was examined on cultured MCF-7 human breast cancer cells by
exposing cells for 24 h and 48 h to medium containing the complex
at 3–30 mg mlÀ1
concentration (Fig. 8). The polymer–copper(II)
complex inhibited the growth of the cancer cells significantly, in
a dose- and duration-dependent manner. The cytotoxic activity
was determined according to the dose values of the exposure of
the complex required to reduce survival to 50% (IC50), compared
to untreated cells. The IC50 values of the complexes are 20.4 ± 2.5
and 14.3 ± 1.7 lg mlÀ1
after 24 h and 48 h respectively. The poly-
mer–copper (II) complex showed highly effective cytotoxic activity
against MCF-7 cancer cells and the IC50 value of the complex was
lesser for 48 h treatment group than for 24 h treatment group.
Inspite of its high cytotoxic activity against MCF-7 cells, the cyto-
toxic effectiveness was relatively lower when compared to cis-
platin, the IC50 values of which were 13.71 ± 0.5 and
12.56 ± 0.8 lg mlÀ1
for 24 h and 48 h treatment periods, respec-
tively. However, cytotoxic potential apart, cisplatin has been estab-
lished to produce toxic side effects [49] which is not expected with
the polymer–copper(II) complex in present study [50]. The cyto-
toxic effect of the polymer–copper(II) complex may be interpret-
able as due to its amphiphilic nature [51] and, hence, would
penetrate the cell membrane easily, reduce the energy status in
tumors and also to alter hypoxia status in the cancer cell microen-
vironment, which are factors that would influence the antitumor
acidity. It is known that phenanthroline-containing metal com-
plexes have a wide range of biological activities such as antitumor,
antifungal, apoptosis [52–54], interaction with DNA thereby inhib-
iting replication, transcription, and other nuclear functions and
arresting cancer cell proliferation so as to arrest tumor growth.
4.2. Assessment of cell death based on morphological features
Apoptosis is a gene-controlled cell death process, which is char-
acterized by DNA fragmentation, chromatin condensation and
marginalization, membrane blebbing, cell shrinkage, and fragmen-
tation of cells into membrane-enclosed vesicles or apoptotic bodies
to be phagocytosed by macrophages [55]. To further confirm the
mode of cell death induced by the complex on cancer cells AO/EB
(acridine orange/ethidium bromide) staining (Apoptosis Assays)
was adopted, which would reveal the changes in the gross cytology
of the cell with special reference to cytoplasm and nucleus. After
treatment of MCF-7 cancer cells, polymer–copper(II) complex of
the highest degree of coordination, at the respective IC50 concen-
trations for 24 h and 48 h, the cells were observed for the gross
cytological changes. The treated cells revealed all the above cyto-
logical changes (Fig. 9). These cytological changes indicated that
the cells were committed to cell death, mostly, apoptosis and to
a certain extent necrosis.
4.3. Single-cell gel electrophoresis (Comet assay)
Among the different techniques used for measuring and analyz-
ing DNA strand breaks in mammalian cells, the single cell gel elec-
trophoresis assay (Comet assay) is considered as a rapid, simple,
visual and sensitive technique to asses DNA fragmentation typical
of toxic DNA damage and of an early stage of apoptosis [56]. As
Fig. 7. Cyclic voltammograms of [Cu(phen)(L-tyr)BPEI]ClO4 (x = 0.182), [com-
plex] = 1 Â 10À3
M (black) in the presence of DNA (red) [DNA] = 0–8.0 Â 10À4
M,
scan rate: 50 mV sÀ1
. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)
Fig. 8. Inhibition of in vitro cancer cells growth by [Cu(phen)(L-tyr) BPEI]ClO4
(x = 0.182).
64 J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67
Control 24h 48h
0
20
40
60
80
100%ofcells
Normal
Apoptosis
Necrosis
Fig. 9. Photomicrographs of control (the cells were viable as inferred from the green – fluorescence) and AO/EB stained MCF-7 cancer cells treated with the [Cu(phen)(L-
tyr)BPEI]ClO4 (x = 0.182) at 20.4 and 14.3 lg mlÀ1
concentration for 24 and 48 h. Scale bar: 35 lm. The graph shows data on percentage of cells that are normal afflicted with
apoptosis and necrosis in the control and 24 h and 48 h treatment groups. (For interpretation of the references to color in this figure legend, the reader is referred to the web
version of this article.)
Control 24 h 48 h
0
20
40
60
80
100
120
140
%ofcells
Dead
Highly Damaged
Damaged
Slightly Damaged
Intact
Fig. 10. Comet images of DNA double strand breaks at 12 and 24 h treatment of [Cu(phen)(L-tyr)BPEI]ClO4 (x = 0.182) at 20.4 and 14.3 lg mlÀ1
concentration. Cells were
grown in RPMI1640 medium containing FBS at 10% final concentration, and streptomycin (10 mg mlÀ1
) and penicillin (10,000 IU mlÀ1
) as antibiotics. The duration-
dependence of the DNA damage is revealed. Scale bar: 35 lm. DNA damage in MCF-7 cell populations as defined according to the percentage of DNA in the tail.
J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67 65
shown in Fig. 10, the images were used to estimate the DNA con-
tent of individual nuclei and to evaluate the degree of DNA damage
representing the fraction of total DNA in the tail. Cells were
assigned to five groups: 0–20% (intact), 20–40% (slightly damaged),
40–60% (damaged), 60–80% (highly damaged) and >80% (dead).
The results revealed that DNA damage was induced in MCF-7 can-
cer cells by the polymer–copper(II) complex, and the incidence was
greater at 48 h than at 24 h, as shown in Fig. 10.
5. Conclusions
Water soluble polyethyleneimine coordinated–copper(II) com-
plexes containing phenanthroline and L-tyrosine as co-ligands with
various degrees of coordination were synthesised. The complexes
were characterized adopting various spectroscopic techniques
and elemental analysis. The binding between the polymer–cop-
per(II) complexes and DNA was assessed in relation to the polymer
complex with different degrees of copper complex content in the
polymer chain. The electronic absorption spectral studies, emission
studies and ionic strength effect showed that these complexes bind
to DNA via electrostatic modes of binding. These studies indicates
that the binding affinity toward DNA increases with the increase
in the number of copper centres in the polymer. The changes in cir-
cular dichroism and cyclic voltammetry studies of the binding
between one of our complexes in the presence of DNA confirm
the above mentioned modes of binding. Thermal denaturation
studies of the binding between our complexes and DNA reveal that
the complex with higher degree of coordination binds with DNA
and stability enhanced. The polymer–copper(II) complex of the
highest degree of coordination showed good cytotoxic activity
against MCF-7 cancer cell with mostly through apoptosis although
a few cells succumbed to necrosis.
Acknowledgments
We are grateful to the UGC-SAP and DST-FIST programmes of
the Department of Chemistry, Bharathidasan University. Council
of Scientific and Industrial Research (CSIR), New Delhi is acknowl-
edged for financial support [Scheme. No. 09/475(0154)/2010-EMR-
I dated. 09/02/2011] for Senior Research Fellowship to JLP. One of
the authors, SA., thanks for sanction of research schemes, Grant No.
SR/S1/IC-13/2009 of DST, Grant No. 01(2461)/11/EMR-II of CSIR
and also Grant No. 41-223/2012(SR) of UGC. Grants from Doerenk-
amp-Zbinden Foundation, Switzerland, and King Saud University,
Riyadh, Kingdom of Saudi Arabia to MAA are gratefully
acknowledged.
References
[1] J.B. Chaires, Drug-DNA interactions, Curr. Opin. Chem. Biol. 8 (1998) 314–320.
[2] K.J. Du, J.Q. Wan, J.F. Kou, G.Y. Li, L.L. Wang, H. Chao, L.N. Ji, Synthesis, DNA-
binding and topoisomerase inhibitory activity of ruthenium(II) polypyridyl
complexes, Eur. J. Med. Chem. (2011) 1056–1065.
[3] S. Roy, K.D. Hagen, U. Maheswari, M. Lutz, A.L. Spek, J. Reedijk, G.P. van Wezel,
Phenanthroline derivatives with improved selectivity as DNA-targeting
anticancer or antimicrobial drugs, Chem. Med. Chem. 3 (2008) 1427–1434.
[4] P. Borst, S. Rottenberg, J. Jonkers, How do real tumors become resistant to
cisplatin?, Cell Cycle 15 (2008) 1353–1359
[5] X. Wang, Fresh platinum complexes with promising antitumor activity,
Anticancer Agents Med. Chem. 10 (2010) 396–411.
[6] W.H. Ang, P.J. Dyson, Classical and non-classical ruthenium-based anticancer
drugs: towards targeted chemotherapy, Eur. J. Inorg. Chem. 2006 (2006) 4003–
4018.
[7] X.-L. Zhao, Z.-S. Li, Z.-B. Zheng, A.-G. Zhang, K.-Z. Wang, PH luminescence
switch, DNA binding and photocleavage, and cytotoxicity of a dinuclear
ruthenium complex, Dalton Trans. 42 (2013) 5764–5777.
[8] R. Chadha, V.K. Kapoor, D. Thakur, R. Kaur, P. Arora, D.V.S. Jain, Drug carrier
systems for anticancer agents: a review, J. Sci. Ind. Res. 67 (2008) 185–197.
[9] B.N. Ames, M.K. Shigenaga, T.M. Hagen, Oxidants, antioxidants, and the
degenerative diseases of aging, Proc. Natl. Acad. Sci. U.S.A. 90 (1993) 7915–7922.
[10] D.S. Sigman, Chemical nucleases, Biochemistry 29 (1990) 9097–11105.
[11] G. Majella, S. Vivienne, M. Malachy, D. Michael, M. Vickie, Synthesis and anti-
Candida activity of copper(II) and manganese(II) carboxylate complexes: X-ray
crystal structures of [Cu(sal)(bipy)]ÁC2H5OHÁH2O and [Cu(norb)(phen)2]Á
6.5H2O (salH2 = salicylic acid; norbH2 = cis-5-norbornene-endo-2,3-
dicarboxylic acid; bipy = 2,2-bipyridine; phen = 1,10-phenanthroline),
Polyhedron 18 (1999) 2931–2939.
[12] J.D. Saha, U. Sandbhor, K. Shirisha, S. Padhye, D. Deobagkar, C.E. Ansond, A.K.
Powell, A novel mixed-ligand antimycobacterial dimeric copper complex of
ciprofloxacin and phenanthroline, Bioorg. Med. Chem. Lett. 14 (2004) 3027–
3032.
[13] S. Zhang, Y. Zhu, C. Tu, H. Wei, Z. Yang, L. Lin, J. Ding, J. Zhang, Z. Guo, A novel
cytotoxic ternary copper(II) complex of 1,10-phenanthroline and l-threonine
with DNA nuclease activity, J. Inorg. Biochem. 98 (2004) 2099–2106.
[14] Y. Ni, D. Lin, S. Kokot, Synchronous fluorescence, UV–visible
spectrophotometric, and voltammetric studies of the competitive interaction
of bis(1,10-phenanthroline)copper(II) complex and neutral red with DNA,
Anal. Biochem. 352 (2006) 231–242.
[15] T. Gupta, S. Dhar, M. Nethaji, A.R. Chakravarty,
Bis(dipyridophenazine)copper(II) complex as major groove directing
synthetic hydrolase, Dalton Trans. (2004) 1896–1900.
[16] A.D. Burrows, C.W. Chan, M.M. Chowdhry, J.E. McGrady, D.M.P. Mingos,
Multidimensional crystal engineering of bifunctional metal complexes
containing complementary triple hydrogen bonds, Chem. Soc. Rev. 24 (1995)
329–339.
[17] M. Chauhan, K. Banerjee, F. Arjmand, DNA binding studies of novel copper(II)
complexes containing L-tryptophan as chiral auxiliary. In vitro antitumor
activity of Cu–Sn2 complex in human neuroblastoma cells, Inorg. Chem. 46
(2007) 3072–3082.
[18] B.L. Rivas, K.E. Geckeler, Synthesis and metal complexation of
poly(ethylenimine) and derivatives, Adv. Polym. Sci. 102 (1992) 171–188.
[19] E. Tsuchida, H. Nishide, T. Ohkawa, Bulletin of the science and engineering
research laboratory, Waseda Univ. 69 (1975) 31–34.
[20] R.S. Kumar, K. Sasikala, S. Arunachalam, DNA interaction of some polymer–
copper(II) complexes containing 2,20
-bipyridyl ligand and their antimicrobial
activities, J. Inorg. Biochem. 102 (2008) 234–241.
[21] J. Lakshmipraba, S. Arunachalam, Studies on the interactions of polymer-
anchored copper(II) complexes with tRNA, Transition Met. Chem. 35 (2010)
477–482.
[22] J. Lakshmipraba, S. Arunachalam, D.A. Gandi, T. Thirunalasundari, Synthesis,
nucleic acid binding and cytotoxicity of polyethyleneimine–copper(II)
complexes containing 1,10-phenanthroline and L-valine, Eur. J. Med. Chem.
46 (2011) 3013–3021.
[23] J. Lakshmipraba, S. Arunachalam, A. Riyasdeen, R. Dhivya, S. Vignesh, M.A.
Akbarsha, R.A. James, DNA/RNA binding and anticancer/antimicrobial
activities of polymer–copper(II) complexes, Spectrochim. Acta, Part A 109
(2013) 23–31.
[24] T. Sugimori, H. Masuda, N. Ohata, K. Koiwai, A. Odani, O. Yamauchi, Structural
dependence of aromatic ring stacking and related weak interactions in ternary
amino acid–copper(II) complexes and its biological implication, Inorg. Chem.
36 (1997) 576–583.
[25] T. Lundback, T. Hard, Sequence-specific DNA-binding dominated by
dehydration, Proc. Natl. Acad. Sci. 93 (1996) 4754–4759.
[26] R. Marty, C.N. N’soukpoé-Kossi, D.M. Charbonneau, L. Kreplak, H.-A. Tajmir-
Riahi, Structural characterization of cationic lipid-tRNA complexes, Nucleic
Acids Res. (2009) 1–11.
[27] T. Mosmann, Rapid colorimetric assay for cellular growth and survival:
application to proliferation and cytotoxicity assays, J. Immunol. Methods 65
(1983) 55–63.
[28] D.L. Spector, R.D. Goldman, L.A. Leinwand, Cell: A Laboratory Manual, Culture
and Biochemical Analysis of Cells, vol. 1, Cold Spring Harbor Laboratory Press,
Cold Spring Harbor, New York, 1998, pp. 34.1–34.9.
[29] N.P. Singh, M.T. MacCoy, R.R. Tice, E.L. Schneider, A simple technique for
quantitation of low levels of DNA damage in individual cells, Exp. Cell 175
(1988) 184–191.
[30] Y. Kurimura, E. Tsuchida, M. Kaneko, Preparation and properties of some water
soluble Co(III)-poly-4-vinylpyridine complexes, J. Polym. Sci., Part A-1 9 (1971)
3511–3519.
[31] J.W. Kolis, D.E. Hamilton, N.K. Kildahl, Nickel(II) and copper(II) complexes of
linear pentadentate ligands derived from substituted o-hydroxyaceto- and o-
hydroxybenzophenones, Inorg. Chem. 18 (1979) 1826–1831.
[32] A.A. Schilt, R.C. Taylor, Infrared spectra of 1,10-phenanthroline metal
complexes in the rock salt region below 2000 cmÀ1
, J. Inorg. Nucl. Chem. 9
(1959) 211–221.
[33] K.E. Geckeler, G. Lange, H. Eberhardt, E. Bayer, Preparation and application of
water-soluble polymer–metal complexes, Pure Appl. Chem. 52 (1980) 1883–
1905.
[34] T. Biver, B. García, J.M. Leal, F. Secco, E. Turriani, Left-handed DNA:
intercalation of the cyanine thiazole orange and structural changes. A kinetic
and thermodynamic approach, Phys. Chem. Chem. Phys. 12 (2010) 13309–
13317.
[35] Q.-Q. Zhang, F. Zhang, W.-G. Wang, X.-L. Wang, Synthesis, crystal structure and
DNA binding studies of a binuclear copper(II) complex with phenanthroline, J.
Inorg. Biochem. 100 (2006) 1344–1352.
[36] Y.-I. Zhou, Y.-Z. Li, The interaction of poly(ethylenimine) with nucleic acids
and its use in determination of nucleic acids based on light scattering,
Spectrochim. Acta, Part A 60 (2004) 377–384.
66 J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67
[37] M.J. Waring, Complex formation between ethidium bromide and nucleic acids,
J. Mol. Biol. 13 (1965) 269–282.
[38] J.G. Liu, B.H. Ye, H. Li, L.N. Ji, R.H. Li, J.Y. Zhou, Synthesis, characterization and
DNA-binding properties of novel dipyridophenazine complex of ruthenium
(II): [Ru(IP)2(DPPZ)]2+
, J. Inorg. Biochem. 73 (1999) 117–122.
[39] S. Ramakrishnan, M. Palaniandavar, Interaction of rac-[Cu(diimine)3]2+
and
rac-[Zn(diimine)3]2+
complexes with CT DNA: effect of fluxional Cu(II)
geometry on DNA binding, ligand-promoted exciton coupling and prominent
DNA cleavage, Dalton Trans. (2008) 3866–3878.
[40] N. Shahabadi, S. Kashanian, M. Khosravi, M. Mahdavi, Multispectroscopic DNA
interaction studies of a water-soluble nickel(II) complex containing different
dinitrogen aromatic ligands, Transition Met. Chem. 35 (2010) 699–705.
[41] S. Kashanian, M.M. Khodaei, H. Roshanfekr, N. Shahabadi, G. Mansouri, DNA
binding, DNA cleavage and cytotoxicity studies of a new water soluble
copper(II) complex: the effect of ligand shape on the mode of binding,
Spectrochim. Acta – Part A 86 (2012) 351–359.
[42] S. Kashanian, M.M. Khodaei, H. Roshanfekr, H. Peyman, DNA interaction of
[Cu(dmp)(phen-dion)] (dmp = 4,7 and 2,9 dimethyl phenanthroline,
phen-dion = 1,10-phenanthroline-5,6-dion) complexes and DNA-based
electrochemical biosensor using chitosan-carbon nanotubes composite film,
Spectrochim. Acta – Part A 114 (2013) 642–649.
[43] M.T. Record Jr., T.M. Lohman, P. De Haseth, Ion effects on ligand– nucleic acid
interactions, J. Mol. Biol. 107 (1976) 145–158.
[44] M.X. Tang, F.C. Szoka, The influence of polymer structure on the interactions of
cationic polymers with DNA and morphology of the resulting complexes, Gene
Therapy 4 (1997) 823–832.
[45] K. van Holde, W.C. Johnson, P.S. Ho, Principles of Physical Biochemistry,
Prentice Hall, New York, 1998.
[46] S. Zhang, Y. Zhu, C. Tu, H. Wei, Z. Yang, L. Lin, J. Ding, J. Zhang, Z. Guo, A novel
cytotoxic ternary copper(II) complex of 1,10-phenanthroline and L-threonine
with DNA nuclease activity, J. Inorg. Biochem. 98 (2004) 2099–2106.
[47] T. Carter, M.A. Rodriguez, A.J. Bard, Voltammetric studies of the interaction of
metal chelates with DNA. 2. Tris-chelated complexes of cobalt(III) and iron(II)
with 1,10-phenanthroline and 2,2’-bipyridine, J. Am. Chem. Soc. 111 (1989)
8901–8911.
[48] M.T. Carter, A.J. Bard, Voltammetric studies of the interaction of tris(1,10-
phenanthroline)cobalt (III) with DNA, J. Am. Chem. Soc. 109 (1987) 7528–
7530.
[49] A.-M. Florea, D. Büsselberg, Cisplatin as an anti-tumor drug: cellular
mechanisms of activity, drug resistance and induced side effects, Cancers 3
(2011) 1351–1371.
[50] G.J. Brewer, Copper control as an antiangiogenic anticancer therapy: lessons
from treating Wilson’s disease, Exp. Biol. Med. (Maywood) 226 (2001) 665–
673.
[51] J.P. Behr, Gene transfer with synthetic cationic amphiphiles. Prospects for gene
therapy, Bioconjug. Chem. 5 (1994) 382–389.
[52] B. Coyle, P. Kinsella, M. McCann, M. Devereux, R. O’Connor, M. Clynes, K.
Kavanagh, Induction of apoptosis in yeast and mammalian cells by exposure to
1,10-phenanthroline metal complexes, Toxicol. In Vitro 18 (2004) 63–70.
[53] B. Coyle, M. McCann, K. Kavanagh, M. Devereux, M. Geraghty, Mode of anti-
fungal activity of 1,10-phenanthroline and its Cu(II), Mn(II) and Ag(I)
complexes, BioMetals 16 (2003) 321–329.
[54] M. McCann, M. Geraghty, M. Devereux, D. O’Shea, J. Mason, L. O’Sullivan,
Insights into the mode of action of the anti-Candida activity of
1,10-phenanthroline and its metal chelates, Metal-Based Drugs 7 (2000)
185–193.
[55] P. Moller, L.E. Knudsen, S. Loft, H. Wallin, The comet assay as a rapid screen test
in biomonitoring occupational exposure and effect of confounding factors,
Cancer Epidemiol. Biomarkers Prevent. 9 (2000) 1005–1015.
[56] M.A. Maire, C. Rast, Y. Landkocz, P. Vasseur, 2,4-Dichlorophenoxyacetic acid:
effects on Syrian hamster embryo (SHE) cell transformation, c-Myc expression,
DNA damage and apoptosis, Mutat. Res. 631 (2007) 124–136.
J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67 67

More Related Content

What's hot

Synthesis, spectroscopic, electrochemical, magnetic properties and super oxid...
Synthesis, spectroscopic, electrochemical, magnetic properties and super oxid...Synthesis, spectroscopic, electrochemical, magnetic properties and super oxid...
Synthesis, spectroscopic, electrochemical, magnetic properties and super oxid...
IOSR Journals
 
International Journal of Engineering and Science Invention (IJESI)
International Journal of Engineering and Science Invention (IJESI)International Journal of Engineering and Science Invention (IJESI)
International Journal of Engineering and Science Invention (IJESI)
inventionjournals
 
2046-1682-5-16
2046-1682-5-162046-1682-5-16
2046-1682-5-16
Rasha Bassam
 
26248 28893-1-pb
26248 28893-1-pb26248 28893-1-pb
26248 28893-1-pb
Taghreed Al-Noor
 
2006 Hamada Et Al Srmnc
2006 Hamada Et Al Srmnc2006 Hamada Et Al Srmnc
2006 Hamada Et Al Srmnc
yzhamada
 
Nanocatalyst
NanocatalystNanocatalyst
Nanocatalyst
Viji Vijitha
 
H0262048052
H0262048052H0262048052
H0262048052
inventionjournals
 
Njc15 publication 15
Njc15 publication 15Njc15 publication 15
Njc15 publication 15
dionisio31
 
Cucurbita maxima
Cucurbita maximaCucurbita maxima
Cucurbita maxima
Juan Manuel Garcia Ayala
 
The DNA cleavage and antimicrobial studies of Co(II), Ni(II), Cu(II) and Zn(I...
The DNA cleavage and antimicrobial studies of Co(II), Ni(II), Cu(II) and Zn(I...The DNA cleavage and antimicrobial studies of Co(II), Ni(II), Cu(II) and Zn(I...
The DNA cleavage and antimicrobial studies of Co(II), Ni(II), Cu(II) and Zn(I...
IOSR Journals
 
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
IOSRJAC
 
SYNTHESIS AND CHARACTERIZATION OF THERMAL ANALYSIS OF La (II) MACROCYCLIC COM...
SYNTHESIS AND CHARACTERIZATION OF THERMAL ANALYSIS OF La (II) MACROCYCLIC COM...SYNTHESIS AND CHARACTERIZATION OF THERMAL ANALYSIS OF La (II) MACROCYCLIC COM...
SYNTHESIS AND CHARACTERIZATION OF THERMAL ANALYSIS OF La (II) MACROCYCLIC COM...
inventionjournals
 
I0445159
I0445159I0445159
I0445159
IOSR Journals
 
Synthesis and characterization of water-soluble free-base, zinc and copper
Synthesis and characterization of water-soluble free-base, zinc and copperSynthesis and characterization of water-soluble free-base, zinc and copper
Synthesis and characterization of water-soluble free-base, zinc and copper
Angela Mammana
 
Synthesis and Characterization of Thermal Analysis of La (Ii) Macrocyclic Com...
Synthesis and Characterization of Thermal Analysis of La (Ii) Macrocyclic Com...Synthesis and Characterization of Thermal Analysis of La (Ii) Macrocyclic Com...
Synthesis and Characterization of Thermal Analysis of La (Ii) Macrocyclic Com...
inventionjournals
 
Nucleic Acids Research (1995) 23, 159-164
Nucleic Acids Research (1995) 23, 159-164Nucleic Acids Research (1995) 23, 159-164
Nucleic Acids Research (1995) 23, 159-164
Dinesh Barawkar
 
Celis 2015, KpCld
Celis 2015, KpCldCelis 2015, KpCld
Celis 2015, KpCld
Arianna Celis
 
PBL 490 Publication
PBL 490 PublicationPBL 490 Publication
PBL 490 Publication
Jen Klabis
 

What's hot (18)

Synthesis, spectroscopic, electrochemical, magnetic properties and super oxid...
Synthesis, spectroscopic, electrochemical, magnetic properties and super oxid...Synthesis, spectroscopic, electrochemical, magnetic properties and super oxid...
Synthesis, spectroscopic, electrochemical, magnetic properties and super oxid...
 
International Journal of Engineering and Science Invention (IJESI)
International Journal of Engineering and Science Invention (IJESI)International Journal of Engineering and Science Invention (IJESI)
International Journal of Engineering and Science Invention (IJESI)
 
2046-1682-5-16
2046-1682-5-162046-1682-5-16
2046-1682-5-16
 
26248 28893-1-pb
26248 28893-1-pb26248 28893-1-pb
26248 28893-1-pb
 
2006 Hamada Et Al Srmnc
2006 Hamada Et Al Srmnc2006 Hamada Et Al Srmnc
2006 Hamada Et Al Srmnc
 
Nanocatalyst
NanocatalystNanocatalyst
Nanocatalyst
 
H0262048052
H0262048052H0262048052
H0262048052
 
Njc15 publication 15
Njc15 publication 15Njc15 publication 15
Njc15 publication 15
 
Cucurbita maxima
Cucurbita maximaCucurbita maxima
Cucurbita maxima
 
The DNA cleavage and antimicrobial studies of Co(II), Ni(II), Cu(II) and Zn(I...
The DNA cleavage and antimicrobial studies of Co(II), Ni(II), Cu(II) and Zn(I...The DNA cleavage and antimicrobial studies of Co(II), Ni(II), Cu(II) and Zn(I...
The DNA cleavage and antimicrobial studies of Co(II), Ni(II), Cu(II) and Zn(I...
 
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
 
SYNTHESIS AND CHARACTERIZATION OF THERMAL ANALYSIS OF La (II) MACROCYCLIC COM...
SYNTHESIS AND CHARACTERIZATION OF THERMAL ANALYSIS OF La (II) MACROCYCLIC COM...SYNTHESIS AND CHARACTERIZATION OF THERMAL ANALYSIS OF La (II) MACROCYCLIC COM...
SYNTHESIS AND CHARACTERIZATION OF THERMAL ANALYSIS OF La (II) MACROCYCLIC COM...
 
I0445159
I0445159I0445159
I0445159
 
Synthesis and characterization of water-soluble free-base, zinc and copper
Synthesis and characterization of water-soluble free-base, zinc and copperSynthesis and characterization of water-soluble free-base, zinc and copper
Synthesis and characterization of water-soluble free-base, zinc and copper
 
Synthesis and Characterization of Thermal Analysis of La (Ii) Macrocyclic Com...
Synthesis and Characterization of Thermal Analysis of La (Ii) Macrocyclic Com...Synthesis and Characterization of Thermal Analysis of La (Ii) Macrocyclic Com...
Synthesis and Characterization of Thermal Analysis of La (Ii) Macrocyclic Com...
 
Nucleic Acids Research (1995) 23, 159-164
Nucleic Acids Research (1995) 23, 159-164Nucleic Acids Research (1995) 23, 159-164
Nucleic Acids Research (1995) 23, 159-164
 
Celis 2015, KpCld
Celis 2015, KpCldCelis 2015, KpCld
Celis 2015, KpCld
 
PBL 490 Publication
PBL 490 PublicationPBL 490 Publication
PBL 490 Publication
 

Viewers also liked

Magento One Page Checkout
Magento One Page CheckoutMagento One Page Checkout
Magento One Page Checkout
magestore co
 
Construction agency in london
Construction agency in londonConstruction agency in london
Construction agency in london
braina rdridge
 
Buccal pad fat for cyst Indian Jr SCT volume 2, issue 1, April 16
Buccal pad fat for cyst Indian Jr SCT volume 2, issue 1, April 16Buccal pad fat for cyst Indian Jr SCT volume 2, issue 1, April 16
Buccal pad fat for cyst Indian Jr SCT volume 2, issue 1, April 16
Avinash Gandi
 
Magento One Page Checkout extension
Magento One Page Checkout extensionMagento One Page Checkout extension
Magento One Page Checkout extension
magestore co
 
E portfoilio ashwani rehal 2013
E portfoilio ashwani rehal 2013E portfoilio ashwani rehal 2013
E portfoilio ashwani rehal 2013
ashwanirehal
 
Aitsl g.standards
Aitsl g.standardsAitsl g.standards
Aitsl g.standards
ashwanirehal
 

Viewers also liked (6)

Magento One Page Checkout
Magento One Page CheckoutMagento One Page Checkout
Magento One Page Checkout
 
Construction agency in london
Construction agency in londonConstruction agency in london
Construction agency in london
 
Buccal pad fat for cyst Indian Jr SCT volume 2, issue 1, April 16
Buccal pad fat for cyst Indian Jr SCT volume 2, issue 1, April 16Buccal pad fat for cyst Indian Jr SCT volume 2, issue 1, April 16
Buccal pad fat for cyst Indian Jr SCT volume 2, issue 1, April 16
 
Magento One Page Checkout extension
Magento One Page Checkout extensionMagento One Page Checkout extension
Magento One Page Checkout extension
 
E portfoilio ashwani rehal 2013
E portfoilio ashwani rehal 2013E portfoilio ashwani rehal 2013
E portfoilio ashwani rehal 2013
 
Aitsl g.standards
Aitsl g.standardsAitsl g.standards
Aitsl g.standards
 

Similar to 1-s2.0-S1011134414003510-main

D0212016025
D0212016025D0212016025
D0212016025
researchinventy
 
CV 22-09-15 - updated
CV  22-09-15 - updatedCV  22-09-15 - updated
CV 22-09-15 - updated
Syeda Gilani
 
science journal pharmaceutical & biological
science journal  pharmaceutical & biologicalscience journal  pharmaceutical & biological
science journal pharmaceutical & biological
nagarajukarnatik4hit
 
journals to publish paper
journals to publish paperjournals to publish paper
journals to publish paper
chaitanya451336
 
A STUDY TO EVALUATE THE IN VITRO ANTIMICROBIAL ACTIVITY AND ANTIANDROGENIC E...
A STUDY TO EVALUATE THE IN VITRO ANTIMICROBIAL ACTIVITY AND  ANTIANDROGENIC E...A STUDY TO EVALUATE THE IN VITRO ANTIMICROBIAL ACTIVITY AND  ANTIANDROGENIC E...
A STUDY TO EVALUATE THE IN VITRO ANTIMICROBIAL ACTIVITY AND ANTIANDROGENIC E...
Dr. Pradeep mitharwal
 
JML_2015_prakash
JML_2015_prakashJML_2015_prakash
JML_2015_prakash
omshamli
 
PNIPAM-b-PMAA Poster
PNIPAM-b-PMAA PosterPNIPAM-b-PMAA Poster
PNIPAM-b-PMAA Poster
Cayla Cook
 
Spectral studies of 5-({4-amino-2-[(Z)-(2-hydroxybenzylidene) amino] pyrimidi...
Spectral studies of 5-({4-amino-2-[(Z)-(2-hydroxybenzylidene) amino] pyrimidi...Spectral studies of 5-({4-amino-2-[(Z)-(2-hydroxybenzylidene) amino] pyrimidi...
Spectral studies of 5-({4-amino-2-[(Z)-(2-hydroxybenzylidene) amino] pyrimidi...
IOSR Journals
 
293-JMES-2247-Ellouz-Publishe Paper-July 2016
293-JMES-2247-Ellouz-Publishe Paper-July 2016293-JMES-2247-Ellouz-Publishe Paper-July 2016
293-JMES-2247-Ellouz-Publishe Paper-July 2016
Ibrahim Abdel-Rahman
 
Aijrfans14 270
Aijrfans14 270Aijrfans14 270
Aijrfans14 270
Iasir Journals
 
Matthew R Mills CV
Matthew R Mills CVMatthew R Mills CV
Matthew R Mills CV
millsware
 
Capstone3
Capstone3Capstone3
Capstone3
Sravanthi Reddy
 
SYNTHESIS, PHYSICO-CHEMICAL AND ANTIMICROBIAL PROPERTIES OF SOME METAL (II) -...
SYNTHESIS, PHYSICO-CHEMICAL AND ANTIMICROBIAL PROPERTIES OF SOME METAL (II) -...SYNTHESIS, PHYSICO-CHEMICAL AND ANTIMICROBIAL PROPERTIES OF SOME METAL (II) -...
SYNTHESIS, PHYSICO-CHEMICAL AND ANTIMICROBIAL PROPERTIES OF SOME METAL (II) -...
International Journal of Technical Research & Application
 
One pot synthesis of cu(ii) 2,2′ bipyridyl complexes of 5-hydroxy-hydurilic acid
One pot synthesis of cu(ii) 2,2′ bipyridyl complexes of 5-hydroxy-hydurilic acidOne pot synthesis of cu(ii) 2,2′ bipyridyl complexes of 5-hydroxy-hydurilic acid
One pot synthesis of cu(ii) 2,2′ bipyridyl complexes of 5-hydroxy-hydurilic acid
rkkoiri
 
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
madlovescience
 
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
tshankar20134
 
15.isca irjbs-2012-138
15.isca irjbs-2012-13815.isca irjbs-2012-138
15.isca irjbs-2012-138
Abhijith Anil
 
khalid & Ahmed Sabeeh research
khalid & Ahmed Sabeeh researchkhalid & Ahmed Sabeeh research
khalid & Ahmed Sabeeh research
Khalid Walled
 
56.Synthesis, Characterization and Antibacterial activity of iron oxide Nanop...
56.Synthesis, Characterization and Antibacterial activity of iron oxide Nanop...56.Synthesis, Characterization and Antibacterial activity of iron oxide Nanop...
56.Synthesis, Characterization and Antibacterial activity of iron oxide Nanop...
Annadurai B
 
Mudasir Presentation.pptx
Mudasir Presentation.pptxMudasir Presentation.pptx
Mudasir Presentation.pptx
MudasirHussain65
 

Similar to 1-s2.0-S1011134414003510-main (20)

D0212016025
D0212016025D0212016025
D0212016025
 
CV 22-09-15 - updated
CV  22-09-15 - updatedCV  22-09-15 - updated
CV 22-09-15 - updated
 
science journal pharmaceutical & biological
science journal  pharmaceutical & biologicalscience journal  pharmaceutical & biological
science journal pharmaceutical & biological
 
journals to publish paper
journals to publish paperjournals to publish paper
journals to publish paper
 
A STUDY TO EVALUATE THE IN VITRO ANTIMICROBIAL ACTIVITY AND ANTIANDROGENIC E...
A STUDY TO EVALUATE THE IN VITRO ANTIMICROBIAL ACTIVITY AND  ANTIANDROGENIC E...A STUDY TO EVALUATE THE IN VITRO ANTIMICROBIAL ACTIVITY AND  ANTIANDROGENIC E...
A STUDY TO EVALUATE THE IN VITRO ANTIMICROBIAL ACTIVITY AND ANTIANDROGENIC E...
 
JML_2015_prakash
JML_2015_prakashJML_2015_prakash
JML_2015_prakash
 
PNIPAM-b-PMAA Poster
PNIPAM-b-PMAA PosterPNIPAM-b-PMAA Poster
PNIPAM-b-PMAA Poster
 
Spectral studies of 5-({4-amino-2-[(Z)-(2-hydroxybenzylidene) amino] pyrimidi...
Spectral studies of 5-({4-amino-2-[(Z)-(2-hydroxybenzylidene) amino] pyrimidi...Spectral studies of 5-({4-amino-2-[(Z)-(2-hydroxybenzylidene) amino] pyrimidi...
Spectral studies of 5-({4-amino-2-[(Z)-(2-hydroxybenzylidene) amino] pyrimidi...
 
293-JMES-2247-Ellouz-Publishe Paper-July 2016
293-JMES-2247-Ellouz-Publishe Paper-July 2016293-JMES-2247-Ellouz-Publishe Paper-July 2016
293-JMES-2247-Ellouz-Publishe Paper-July 2016
 
Aijrfans14 270
Aijrfans14 270Aijrfans14 270
Aijrfans14 270
 
Matthew R Mills CV
Matthew R Mills CVMatthew R Mills CV
Matthew R Mills CV
 
Capstone3
Capstone3Capstone3
Capstone3
 
SYNTHESIS, PHYSICO-CHEMICAL AND ANTIMICROBIAL PROPERTIES OF SOME METAL (II) -...
SYNTHESIS, PHYSICO-CHEMICAL AND ANTIMICROBIAL PROPERTIES OF SOME METAL (II) -...SYNTHESIS, PHYSICO-CHEMICAL AND ANTIMICROBIAL PROPERTIES OF SOME METAL (II) -...
SYNTHESIS, PHYSICO-CHEMICAL AND ANTIMICROBIAL PROPERTIES OF SOME METAL (II) -...
 
One pot synthesis of cu(ii) 2,2′ bipyridyl complexes of 5-hydroxy-hydurilic acid
One pot synthesis of cu(ii) 2,2′ bipyridyl complexes of 5-hydroxy-hydurilic acidOne pot synthesis of cu(ii) 2,2′ bipyridyl complexes of 5-hydroxy-hydurilic acid
One pot synthesis of cu(ii) 2,2′ bipyridyl complexes of 5-hydroxy-hydurilic acid
 
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
 
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
One pot synthesis of chain-like palladium nanocubes and their enhanced electr...
 
15.isca irjbs-2012-138
15.isca irjbs-2012-13815.isca irjbs-2012-138
15.isca irjbs-2012-138
 
khalid & Ahmed Sabeeh research
khalid & Ahmed Sabeeh researchkhalid & Ahmed Sabeeh research
khalid & Ahmed Sabeeh research
 
56.Synthesis, Characterization and Antibacterial activity of iron oxide Nanop...
56.Synthesis, Characterization and Antibacterial activity of iron oxide Nanop...56.Synthesis, Characterization and Antibacterial activity of iron oxide Nanop...
56.Synthesis, Characterization and Antibacterial activity of iron oxide Nanop...
 
Mudasir Presentation.pptx
Mudasir Presentation.pptxMudasir Presentation.pptx
Mudasir Presentation.pptx
 

1-s2.0-S1011134414003510-main

  • 1. Polyethyleneimine anchored copper(II) complexes: Synthesis, characterization, in vitro DNA binding studies and cytotoxicity studies Jagadeesan Lakshmipraba a , Sankaralingam Arunachalam a,⇑ , Anvarbatcha Riyasdeen b , Rajakumar Dhivya c , Mohammad Abdulkader Akbarsha b a School of Chemistry, Bharathidasan University, Tiruchirappalli 620 024, Tamil Nadu, India b Mahatma Gandi-Doerenkamp Centre, Bharathidasan University, Tiruchirappalli 620 024, Tamil Nadu, India c Department of Biomedical Science, Bharathidasan University, Tiruchirappalli 620 024, Tamil Nadu, India a r t i c l e i n f o Article history: Received 18 August 2014 Received in revised form 15 November 2014 Accepted 19 November 2014 Available online 27 November 2014 a b s t r a c t The water soluble polyethyleneimine–copper(II) complexes, [Cu(phen)(L-tyr)BPEI]ClO4 (where phen = 1,10-phenanthroline, L-tyr = L-tyrosine and BPEI = branched polyethyleneimine) with various degree of copper(II) complex units in the polymer chain were synthesized and characterized by elemental analysis and electronic, FT-IR, EPR spectroscopic techniques. The binding of these complexes with CT- DNA was studied using UV–visible absorption titration, thermal denaturation, emission, circular dichro- ism spectroscopy and cyclic voltammetric methods. The changes observed in the physicochemcial prop- erties indicated that the binding between the polymer–copper complexes and DNA was mostly through electrostatic mode of binding. Among these complexes, the polymer–copper(II) complex with the highest degrees of copper(II) complex units (higher degrees of coordination) showed higher binding constant than those with lower copper(II) complex units (lower degrees of coordination) complexes. The complex with the highest number of metal centre bound strongly due to the cooperative binding effect. Therefore, anticancer study was carried out using this complex. The cytotoxic activity for this complex on MCF-7 breast cancer cell line was determined adopting MTT assay, acridine orange/ethidium bromide (AO/EB) staining and comet assay techniques, which revealed that the cells were committed to specific mode of cell death either apoptosis or necrosis. Ó 2014 Elsevier B.V. All rights reserved. 1. Introduction During the past decade, there has been tremendous interest in the synthesis and studies pertaining to the interaction of various metal complexes with DNA and also screening of these complexes for their anticancer activities so as to replace cisplatin [1–5]. Hence, much attention has been targeted on the design of metal- based complexes, which can bind to DNA. Interaction of metal complexes with DNA should be useful in the development of molecular probes and new therapeutic agents. Cancer research mainly targets the DNA molecule which is the origin of uncontrolled cell division. To block this cell division, there is a need for developing DNA targeted chemotherapy drugs. Sev- eral small molecules/metal complexes have been interacted with DNA to correlate their effects of binding/cleavage behavior on cyto- toxicity [6,7]. In these aspects of drug designing, some of the issues like solubility, transfection efficiency, targeted delivery, etc. may enter as practical problems. To overcome these problems, drug carriers like cationic polymers, surfactants, liposomes and dentri- mers were employed for efficient drug delivery [8]. Copper is a physiologically important metal element that plays an important role in the endogenous oxidative DNA damage asso- ciated with aging and cancer [9]. Copper(II) complexes bearing 1,10-phenanthroline ligand have been widely used due to their high nucleolytic efficiency [10] and numerous biological activities such as antitumor, anti-candida and antimicrobial activities [11,12]. It has been reported that a binuclear copper(II) complex containing 1,10-phenanthroline and a trinuclear copper(II) com- plex containing di-(2-picolyl)amine bind strongly with DNA and cleave more effectively than their corresponding monomeric com- plexes [13–15]. The reports in the literature that advocate design studies on metal complexes that cooperative effect arising from such non-covalent interactions would be a valuable principle in the development of new metal based probes which recognize bio- molecular targets with high specificity [16]. Recent literature indicate that mixed ligand copper(II) com- plexes have been receiving considerable attention for various http://dx.doi.org/10.1016/j.jphotobiol.2014.11.005 1011-1344/Ó 2014 Elsevier B.V. All rights reserved. ⇑ Corresponding author. Fax: +91 431 2407043. E-mail address: arunasurf@yahoo.com (S. Arunachalam). Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67 Contents lists available at ScienceDirect Journal of Photochemistry and Photobiology B: Biology journal homepage: www.elsevier.com/locate/jphotobiol
  • 2. reasons. Copper(II) complexes having amino acid ligand are of interest due to their biological relevance, good DNA binding ability, antimicrobial and anticancer activities [17]. Also, there are reports on drug polymer conjugates as potential candidates for the selec- tive delivery of anticancer agents to tumor tissues. Particularly, polyethyleneimine (PEI), possesses quite a number of advantages as polymer chelating agent, such as good water solubility, high content of functional groups, suitable molecular weights as well as good physical and chemical stabilities [18]. Stable polyethylene- imine–copper(II) complexes are reported where copper ions are hard to elute from the polymer domain to the bulk solution and binding constant values are 1–5 order magnitude greater in the polymer–chelate systems than in the monomeric Cu-complex systems [19]. Our research group has been involved in the synthesis and stud- ies on mixed polymer–copper(II) complexes and their DNA binding properties for the past several years [20,21]. We have reported syn- thesis and nucleic acid binding of polyethyleneimine–copper(II) complexes with L-valine and L-arginine as a co-ligand [22,23]. In the present work we report the synthesis, DNA binding and antitu- mor properties of polyethyleneimine–copper(II) complexes con- taining L-tyrosine as co-ligand. L-tyrosine contains benzene ring which can influence the binding of these complexes with DNA through p–p interactions. 2. Experimental section 2.1. Materials Calf thymus DNA and branched polyethyleneimine (BPEI) (Mw ca. 25,000) were obtained from Sigma–Aldrich, Germany and were used as obtained. Copper(II) chloride dihydrate, 1,10-phenanthro- line (Merck, India) and tyrosine (Loba Chemie, India) and were used as received. The precursor complex, [Cu(phen)(L-tyr)(H2O)]ClO4, was prepared as reported earlier [24]. A solution of calf thymus DNA in the buffer gave a UV absorbance ratio of $1.8–1.9: 1 at 260 and 280 nm, indicating that the DNA was sufficiently free of protein. The concentration of CT DNA in base pairs was determined by UV absorbance at 260 nm by taking the molar extinction coefficient value 13,200 MÀ1 cmÀ1 for DNA at 260 nm [25,26]. All the experi- ments involving the interaction of the polymer–copper(II) complex with DNA were carried out using buffer containing 5 mM Tris–HCl/ 50 mM NaCl at pH 7.0 in twice distilled water. MCF-7 cell line was obtained from National Centre For Cell Science (NCCS), Pune. 2.2. [Cu(phen)(L-tyr)BPEI]ClO4 To a solution of branched polyethyleneimine (BPEI) (0.15 g, 3.4 mmol) dissolved in ethanol (15 ml), [Cu(phen)(L-tyr) (H2O)]ClO4 (0.8 g, 1.4 mmol) in water was added slowly with stirring. The mixture was heated between 50 and 60 °C for 15 h in a water bath with stirring. The resulting dark blue solution was dialyzed at 15 °C against distilled water for 4–5 days. The solvent was then evaporated in a rotary evaporator under reduced pressure at room temperature. The dark-bluish filmy substance obtained was pulverized and dried. Yield = 0.23 g for x = 0.182. (Anal. Calc.: C, 33.77, H, 4.40, N, 14.50, O, 5.81, Cu 18.35, Found: H, 5.02, N, 14.25, O, 5.77, Cu 18.37% (x = 0.182 obtained from carbon content). IR (KBr, cmÀ1 ): m(NAH) 3446, m(CAC) 2924, m(COOA) 1099, m(C@C) 1471, m(C@N) 1390, 1099, m(CAH) 852, m(CAH) 730; UV (kmax, nm,(e, MÀ1 cmÀ1 )): 227 (28,120) 272 (54,240), 294 (41,830), 645 (11,090). EPR (77 K, g|| 2.211 and g 2.017; RT giso = 2.0753. Polymer–copper(II) complex samples with various numbers of copper(II) complex units bound to the polymer chain were synthe- sized by changing the amount of reactants in the reaction solution, the duration of the reaction time and the reaction temperature. 2.3. Physical measurements Elemental analysis was determined at Sophisticated Analytical Instrument Facility (SAIF), Lucknow, India. Absorption spectra and thermal denaturation studies were recorded on a UV–Vis– NIR Cary300 Spectrophotometer using cuvettes of 1 cm path length in tris buffer solution, and emission spectra were recorded on a JASCO FP 770 spectrofluorimeter. FT-IR spectra were recorded on a FT-IR JASCO 460 PLUS spectrophotometer with samples pre- pared as KBr pellets. EPR spectra were recorded on a JEOL-FA200 EPR spectrometer at room temperature and at LNT in methanol solution. Absorption titration experiments of polymer–copper(II) complexes in buffer (50 mM NaCl–5 mM Tris–HCl, pH 7.0) were performed by using a fixed complex concentration to which incre- ments of the DNA stock solutions were added. Polymer–copper(II) complex-DNA solutions were incubated for 10 min before the absorption spectra were recorded. Equal amount of DNA was added to both the complex and reference solutions to eliminate the absorbance of DNA itself. For fluorescence quenching experiments DNA was pretreated with ethidium bromide (EB) for 30 min. The polymer–copper(II) complexes were then added to this mixture and their effect on the emission intensity was measured. Samples were excited at 450 nm and emission was observed between 500 and 700 nm. Cir- cular dichroism spectra were recorded at room temperature using the same tris buffer. For the cyclic voltammetry experiments, the electrode surfaces were freshly polished with alumina powder and then sonicated in ethanol and distilled water for 1 min prior to each experiment. Then the electrodes were rinsed throughly with distilled water. Cyclic voltammetric experiments were per- formed at 25.0 ± 0.2 °C in a single compartment cell with a three- electrode configuration (glassy carbon working electrode, platinum wire auxiliary electrode and saturated calomel reference electrode). The solution was deoxygenated with nitrogen gas for 20 min prior to experiments. 2.4. Cell culture The MCF-7 cancer cells were cultured in RPMI 1640 medium (Sigma–Aldrich, St. Louis, MO, USA), supplemented with 10% fetal bovine serum (Sigma, USA) and 10,000 IU of penicillin and 100 lg mlÀ1 of streptomycin as antibiotics (Himedia, Mumbai, India), in 96 well culture plates, at 37 °C, in a humidified atmo- sphere of 5% CO2, in a CO2 incubator (Forma, Thermo Scientific, USA). All the experiments were performed using cells from passage 15 or less. 2.4.1. Cytotoxicity assay (MTT assay) The polymer–copper(II) complex of the highest degree of coor- dination was dissolved in DMSO, diluted in culture medium and used to treat the model cell line over a complex concentration range of 3–30 lg mlÀ1 for a period of 24 h and 48 h. DMSO at 0.5% concentration in the culture medium was used as negative control. We have used cisplatin as positive control. A miniaturized viability assay using 3-(4,5-di-methylthiazol-2-yl)-2,5-diphenyl- 2H-tetrazolium bromide (MTT) (Sigma, USA) (5 mg/ml in Phos- phate-Buffered Saline (PBS)) was added to each well and the plates were wrapped with aluminum foil and incubated at 37 °C for 4 h. By this treatment a purple formazone product was formed due to the reduction of MTT by the mitochondrial enzyme succinate dehydrogenase of the cells [27]. The purple formazan product was dissolved by addition of 100 ll of 100% DMSO to each well. The absorbance was monitored at 570 nm (measurement) and 60 J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67
  • 3. 630 nm (reference) using a 96 well plate reader (Bio-Rad, Hercules, CA, USA). Data were collected for four replicates each and used to calculate the respective means. The percentage of inhibition was calculated, from this data, using the formula: The IC50 value was determined as the complex drug concentra- tion that is required to reduce the absorbance to half that of the control. 2.4.2. Acridine orange (AO) and ethidium bromide (EB) staining Acridine orange/ethidium bromide staining was performed as follows: the cell suspension of each sample containing 5 Â 105 cells, was treated with 25 ll of AO and EB solution (1 part of 100 lg mlÀ1 AO and 1 part of 100 lg mlÀ1 EB in PBS) and examined in a fluorescent microscope (Carl Zeiss, Germany) using an UV filter (450–490 nm). Three hundred cells per sample were counted in tetraplicate for each dose point. The cells were scored as viable, apoptotic or necrotic as judged by the staining, nuclear morphol- ogy and membrane integrity [28], and the percentages of apoptotic and necrotic cells were then calculated. Morphological changes were also observed and photographed. 2.4.3. Comet assay DNA damage was detected by adopting comet assay as reported earlier [29]. Cells were suspended in low-melting-point agarose in PBS and pipetted out to microscope slides pre-coated with a layer of normal-melting-point agarose. Slides were chilled on ice for 10 min and then immersed in lysis solution (2.5 M NaCl, 100 mM Na2EDTA, 10 mM Tris, 0.2 mM NaOH, pH 10.01 and Triton X-100) and the solution was kept for 4 h at 4 °C in order to lyse the cells and to permit DNA unfolding. Thereafter, the slides were exposed to alkaline buffer (300 mM NaOH, 1 mM Na2EDTA, pH > 13) for 20 min to allow DNA unwinding. The slides were washed with buf- fer (0.4 M Tris, pH 7.5) to neutralize excess alkali and to remove detergents before staining with ethidium bromide (5 ll in 10 mg mlÀ1 ). Photographs were obtained using the fluorescent microscope. One hundred and fifty cells from each treatment group were digitalized and analyzed using CASP software. The images were used to estimate the DNA content of individual nuclei and to evaluate the degree of DNA damage representing the fraction of total DNA in the tail. 3. Results and discussion 3.1. Degree of coordination The structure of the water soluble polymer–copper(II) complex is shown in Fig. 1. In this figure ‘x’ represents the degree of coordi- nation, which is the number of moles of copper(II) chelate per mole of the repeating unit (amine group) of polymeric ligand. If the entire repeating units (amine group) in the polymer are coordi- nated to copper, then the value of x is 1. The degree of coordination (x) was calculated from carbon content value [19,30,31]. The degree of coordination thus obtained for the polymer–copper(II) complex samples of the present work are 0.059, 0.149, 0.182. The stability of the polymer–copper(II) complexes in solution was verified occasionally by keeping the solution of the polymer– copper(II) complexes in dialysis bags, small amount of the solution from the dialysis bags was taken and the stability was confirmed through absorption spectroscopy. Any free copper complex ion or copper ion in the solution outside the dialysis bag was not observed (which was verified using spectrophotometric method), indicating that polymer–complexes were very stable during the handling of our experiments. Also viscosity, is a significant experiment which gives information about interaction in aqueous solution. In our case, even after dialysis we do not have change in viscosity. This proves that the complexes are very stable in solution. 3.2. Characterization of polymer–copper(II) complexes The FT-IR spectra for the polymer–copper(II) complexes dis- played stretching frequencies around 1471 cmÀ1 and 1390 cmÀ1 which can be attributed to the ring stretching frequencies viz., m(C@C) and m(C@N), respectively, of the coordinated 1,10-phenan- throline and these are at slightly lower frequencies than that of uncoordinated 1,10-phenanthroline. The m(CAH) out-of-plane bending values, around 852 cmÀ1 and 730 cmÀ1 , for the phenan- throline ligand were shifted to 838 cmÀ1 and 693 cmÀ1 , respec- tively, in the complexes. These shifts can be explained by the fact that each of the two nitrogen atoms of phenanthroline ligands donates a pair of electrons to the central copper atom forming coordinate bond [32]. The band obtained around 2924 cmÀ1 can be assigned to CAC stretching vibration of aliphatic CH2 of BPEI whereas the broad band observed around 3446 cmÀ1 can be assigned to the NAH stretching of BPEI [33]. The uncoordinated amino acid exhibited a stretch in the region 1750–1700 cmÀ1 cor- responding to m(COOH). In the complexes, this band was shifted to 1630 cmÀ1 indicating the coordination of carboxylate group to the copper(II) ion. The very strong band around 1099 cmÀ1 can be assigned to the presence of perchlorate anion. The stretching fre- quencies around 508 cmÀ1 and 491 cmÀ1 can be attributed, respec- tively, to copper–nitrogen and copper–oxygen stretching. The UV–visible absorption spectra of all the complexes were recorded in the region 200–800 nm. All the complexes displayed four bands in the regions 230–645 nm. In the UV region, the absorption bands below 300 nm are attributed to intra-ligand tran- sitions whereas in the visible region, the band around 645 nm is assigned as d–d transition. The solid state EPR spectra of the polymer–copper(II) complex (x = 0.203) was recorded in X-band frequencies at room tempera- ture as well as in frozen solution (77 K) in methanol as solvent (Fig. 2). The room temperature EPR spectra of the polymer–cop- per(II) complexes showed single isotropic feature at giso = 2.070– 2.075, and this broadening of isotropic peak is due to intermolecu- lar spin exchange. This intermolecular type of spin exchange is caused by the strong spin coupling which occurs during a coupling Cu N H N H (1-x) x N N .ClO4 NH2 OH O O Fig. 1. Schematic representation of [Cu(phen)(L-tyr)BPEI]ClO4. ½Mean absorbance of untreated cellsðcontrolÞ À Mean absorbance of treated cellsðtestÞŠ Â 100: Mean absorbance of untreated cellsðcontrolÞ J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67 61
  • 4. of two paramagnetic species. At liquid nitrogen temperature we observed three peaks with the third being broad. This is because the copper complex units have been mounted on a polymer chain resulting in some spin–spin coupling between the copper complex units which leads to a small amount of broadening. In liquid nitro- gen temperature the complexes showed the values of g|| = 2.211– 2.216 and g = 2.016–2.020. The existence of g|| > g > 2.00 sug- gests that dx 2 À dy 2 is the ground state with the d9 (Cu2+ ) configu- ration and square pyramidal geometry. 3.3. Absorption studies Electronic absorption spectroscopy is an effective method to examine the binding mode of DNA with polymer–copper(II) metal complexes. Thus, in order to provide evidence for the binding of polymer–copper(II) complexes to DNA, the binding process was monitored by absorption spectroscopy by following the changes in absorption band intensity and its position. On addition of DNA, the absorption spectra of polymer–copper(II) complex showed hyperchromism and slight red shift (Fig. 3). The experi- mental results derived from the UV–visible titration experiments suggest that positively charged polymer-complexes can bind to DNA, probably to the phosphate groups, by electrostatic interac- tion resulting in the stabilization of DNA duplex. Nevertheless, the metal complex units present in the polymer chain contain aro- matic moieties so the binding of the complexes involving partial intercalation of an aromatic ring between the base pairs of DNA cannot be ruled out. From the above studies the intrinsic binding constants (Kb) were determined from the increase of absorption at 294 nm calculated by absorption spectral titration. In order to compare quantitatively the binding affinity with nucleic acids between polymer–copper(II) complexes having different degrees of coordination, the intrinsic binding constants Kbs of the com- plexes were determined using Eq. (4) by assuming a simple model, in which the reaction between the nucleic acid site, P and the cop- per complex unit of the polymer complex, D to form the nucleic acid bound complex, PD as: P þ D Kb PD ð1Þ Kb ¼ ½PDŠ=½PŠ½DŠ ð2Þ where [PD], [P] and [D] represent the respective equilibrium con- centrations of nucleic acid bound copper complex units, nucleic acid sites in base pairs and the copper complex units of the polymer complex. A ¼ eD½DŠ þ ePD½PDŠ ð3Þ CD=A À eDCD ¼ ð1=ePD À eDÞ þ 1=ðePD À eDÞKb1=½PŠ ð4Þ where eD and ePD are the molar extinction coefficient of the free cop- per complex units and apparent molar extinction coefficient of the nucleic acid bound copper complex units respectively, CD total con- centration of copper complex units and A is the experimental absor- bance. An iterative procedure was employed as per the method provided in Ref. [34] to arrive at the Kb values (first [D] set equal to CD, then, once a first estimate of Kb and (ePD À eD) are obtained, a new value of [D] was calculated and so on until convergence is achieved). This procedure yields a binding constant value (Kb) for each complex. As seen from Table 1 the binding constants observed for poly- mer–copper(II) complexes are higher those that of similar type of simple metal complexes like [Cu(phen)(L-tyr) H2O]ClO4 (Kb = 3.75 -  103 MÀ1 ) as well as the polymer alone PEI (Kb = 1.2 MÀ1 ) [35,36]. However, they are very much lower than the potential intercala- tors like ethidium bromide (Kb, 7.0  107 MÀ1 in 40 mM Tris/HCl, pH 7.9) [37] and the partially intercalating complexes like [Co(phen)2(dppz)]3+ (Kb = 9.09  105 MÀ1 ) and [Ru(imp)2(dppz)]2+ (Kb = 2.19  107 MÀ1 ) [38], which implies that these complexes bind to DNA relatively less strongly than classical intercalators and partial intercalators. Also, as seen from the Table, it was observed that the binding constant changes with degree of coordi- nation of copper(II) units in the polymer chain; greater the ratio of copper(II) centres in the polymer chain, higher was the binding constant because when one copper(II) complex unit binds with DNA it will cooperatively act to increase the overall binding ability of the other copper(II) complex units to DNA. Fig. 2. EPR spectrum of [Cu(phen)(L-tyr)(BPEI)]ClO4 (x = 0.203) in methanol at liquid nitrogen temperature. 200 300 400 500 600 0.0 0.5 1.0 0.00001 0.00002 0.00003 0.0000005 0.0000010 0.0000015 0.0000020 0.0000025 0.0000030 [DNA]/εa−εf [DNA] Absorbance Wavelength, nm Fig. 3. Absorption spectra of [Cu(phen)(L-tyr)BPEI]ClO4 (x = 0.182) in the absence of DNA and in the presence of DNA, [complex] = 3  10À5 M, [DNA] = 0–3.2  10À5 M (inset: plot of [DNA]/(ea À ef) vs. [DNA]). Table 1 The intrinsic binding constant (Kb) of [Cu(phen)(L-tyr)BPEI]ClO4, with DNA and RNA and thermal melting temperature in the presence of [Cu(phen)(L-tyr)BPEI]ClO4 with different degree of coordination. Complex Degree of coordination (x) Kb (MÀ1 ) ±0.04 Ksv (MÀ1 ) ±0.03 [Cu(phen)(L-tyr)BPEI]ClO4 0.059 2.10  104 2.13  104 0.149 2.03  105 2.78  104 0.182 7.80  105 3.37  104 62 J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67
  • 5. 3.4. Ethidium bromide displacement assay All polymer–copper(II) complexes were non-emissive upon excitation of the MLCT band, either in aqueous solution or in the presence of DNA. The competitive binding experiments with a well-established quenching assay based on the displacement of the intercalating EB from ct-DNA was carried out in order to get further information regarding the DNA binding properties of poly- mer–metal complexes. The quenching of emission intensity of DNA bound EB (Fig. 4) was analyzed through Stern–Volmer equation, I0/ I = 1 + Ksv[Q], where I0 and I are the fluorescence intensities in the absence and presence of the complex, respectively, Ksv is the linear Stern–Volmer constant and Q is the concentration of polymer– copper(II) complex [39,40]. A plot of I0/I vs. [Q] was drawn and Ksv was obtained from the ratio of slope to intercept (Table 1). As seen from the Table, the Ksv value increases with increase in degree of coordination of polymer–copper(II) complex. This is attributed to the cooperative binding between copper(II) units on the same polymer chain with DNA. This cooperative effect increases with degree of coordination. 3.5. Effect of ionic strength The change in fluorescence intensity of cationic copper(II)–poly- mer complexes to DNA in the presence of NaCl can be used to verify whether the binding mode is electrostatic or intercalative; a linear relation between fluorescence intensity and concentration of NaCl is highly indicative of an electrostatic mode of interaction whereas a non-dependence of fluorescence intensity on ionic strength indi- cates intercalation [41,42]. It is observed that as the concentration of NaCl increases, the relative fluorescence intensity due to ethi- dium bromide increases (Fig. 5). This is due to the competitive bind- ing of Na+ ions to DNA which decreases the binding affinity of the copper(II)–polymer complex to DNA. As the concentration of NaCl increases, a linear increase of fluorescence is noticed, indicating that the cationic copper(II)–polymer complex-DNA interactions for the polymers studied are electrostatic. [43,44] 3.6. Circular dichroism spectral studies CD spectral technique is useful method to monitor the conformational variations of DNA during complex-DNA interac- tions and achieve information on changing DNA conformation by the binding of the metal complex to DNA. DNA has a major longwave positive peak centred at 275 nm and the intensity of this positive peak is similar in magnitude to that of the negative peak centred at 245 nm (Fig. 6) corresponding to the p–p stacking of the base pairs and right handed helicity of B-form DNA in buffer solution [45]. Addition of polymer–copper(II) complex (x = 0.182) to B-form DNA has been shown to induce a B to A transition, result- ing in a CD spectrum with characteristics totally different from those of B-form DNA; the long wave positive peak is larger with a maximum at $270 nm and a very large shortwave peak results below $230 nm [46]. Thus, the increased ellipticity observed at 275 nm when polymer–copper(II) complex binds to DNA can be interpreted as unwinding of B form of DNA due to a decrease in twist angle. This can be tentatively interpreted as B form of DNA becoming more ‘A-like’ upon binding polymer–copper(II) complex. 3.7. Cyclic voltammetry studies The binding of polymer–copper(II) complexes with DNA was further confirmed by cyclic voltammetric studies. The cyclic 500 600 700 0 100 200 300 400 Intensity Wavelength, nm Fig. 4. Emission spectra of EB bound to DNA, [EB] = 2 Â 10À4 M in the absence of complex and in the presence of complex (x = 0.182), [DNA] = 2 Â 10À3 M, [com- plex] = 0–1 Â 10À4 M. 0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8 1.0 RelativeFluoresence [NaCl] mM 1 2 3 Fig. 5. Titration of DNA[DNA] = 2 Â 10À4 M in the presence of ethidium bro- mide[EB] = 2 Â 10À4 M in the presence of polymer–copper(II) complexes[com- plex] = 1 Â 10À4 M as a function of NaCl concentration. 240 280 320 -80 -40 0 40 80 DNA DNA+ complex CD,mdeg Wavelength, nm Fig. 6. Circular dichroism spectra in the absence (black) and in the presence [Cu(phen)(L-tyr)BPEI]ClO4 (x = 0.182), [complex] = 12 Â 10À5 M (red) with DNA, [DNA] = 9 Â 10À5 M. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.) J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67 63
  • 6. voltammogram of polymer–copper(II) complex (x = 0.182) in the absence and presence of DNA is shown in Fig. 7. In the absence of DNA, the cathodic peak potential (Epc) and the anodic peak potential (Epa) of our complex are 339 mV and 574 mV, respec- tively, with a large peak-to-peak separation, DEp, of 235 mV and the ratio of cathodic to anodic peak current (ipc/ipa) is 1.05 indicat- ing a quasi-reversible redox process [47]. The formal potential (E1/ 2) which is taken as the average of Epc and Epa is 0.457 V in the absence of DNA, whereas in the presence of DNA a negative shift in E1/2 by 0.055 V along with increase in DEp of 30 mV has been observed. The ipc/ipa value also increased with the increase of the DNA concentration. Literature report [48] have pointed out that the shift direction of electrochemical potential of metal complex, after reacting with DNA, is related to its binding mode with DNA. A positive shift of the peak potential and the negative shift indicat- ing that electrostatic mode of interaction. The dependence of cathodic current on scan rate was also inves- tigated. In the case of our polymer–copper(II) complex the plot of cathodic current vs. the square root of the scan rate (m1/2 ), was lin- ear for the complex alone, and the complex in the presence of DNA, which indicates that the electrochemical process are diffusion controlled process [48]. 4. Cytotoxic assay 4.1. MTT assay In vitro cytotoxicity of polymer–copper(II) complexes was eval- uated by MTT assay on MCF-7 cells. The cytotoxic effects of the polymer–copper(II) complex of the highest degree of coordination was examined on cultured MCF-7 human breast cancer cells by exposing cells for 24 h and 48 h to medium containing the complex at 3–30 mg mlÀ1 concentration (Fig. 8). The polymer–copper(II) complex inhibited the growth of the cancer cells significantly, in a dose- and duration-dependent manner. The cytotoxic activity was determined according to the dose values of the exposure of the complex required to reduce survival to 50% (IC50), compared to untreated cells. The IC50 values of the complexes are 20.4 ± 2.5 and 14.3 ± 1.7 lg mlÀ1 after 24 h and 48 h respectively. The poly- mer–copper (II) complex showed highly effective cytotoxic activity against MCF-7 cancer cells and the IC50 value of the complex was lesser for 48 h treatment group than for 24 h treatment group. Inspite of its high cytotoxic activity against MCF-7 cells, the cyto- toxic effectiveness was relatively lower when compared to cis- platin, the IC50 values of which were 13.71 ± 0.5 and 12.56 ± 0.8 lg mlÀ1 for 24 h and 48 h treatment periods, respec- tively. However, cytotoxic potential apart, cisplatin has been estab- lished to produce toxic side effects [49] which is not expected with the polymer–copper(II) complex in present study [50]. The cyto- toxic effect of the polymer–copper(II) complex may be interpret- able as due to its amphiphilic nature [51] and, hence, would penetrate the cell membrane easily, reduce the energy status in tumors and also to alter hypoxia status in the cancer cell microen- vironment, which are factors that would influence the antitumor acidity. It is known that phenanthroline-containing metal com- plexes have a wide range of biological activities such as antitumor, antifungal, apoptosis [52–54], interaction with DNA thereby inhib- iting replication, transcription, and other nuclear functions and arresting cancer cell proliferation so as to arrest tumor growth. 4.2. Assessment of cell death based on morphological features Apoptosis is a gene-controlled cell death process, which is char- acterized by DNA fragmentation, chromatin condensation and marginalization, membrane blebbing, cell shrinkage, and fragmen- tation of cells into membrane-enclosed vesicles or apoptotic bodies to be phagocytosed by macrophages [55]. To further confirm the mode of cell death induced by the complex on cancer cells AO/EB (acridine orange/ethidium bromide) staining (Apoptosis Assays) was adopted, which would reveal the changes in the gross cytology of the cell with special reference to cytoplasm and nucleus. After treatment of MCF-7 cancer cells, polymer–copper(II) complex of the highest degree of coordination, at the respective IC50 concen- trations for 24 h and 48 h, the cells were observed for the gross cytological changes. The treated cells revealed all the above cyto- logical changes (Fig. 9). These cytological changes indicated that the cells were committed to cell death, mostly, apoptosis and to a certain extent necrosis. 4.3. Single-cell gel electrophoresis (Comet assay) Among the different techniques used for measuring and analyz- ing DNA strand breaks in mammalian cells, the single cell gel elec- trophoresis assay (Comet assay) is considered as a rapid, simple, visual and sensitive technique to asses DNA fragmentation typical of toxic DNA damage and of an early stage of apoptosis [56]. As Fig. 7. Cyclic voltammograms of [Cu(phen)(L-tyr)BPEI]ClO4 (x = 0.182), [com- plex] = 1 Â 10À3 M (black) in the presence of DNA (red) [DNA] = 0–8.0 Â 10À4 M, scan rate: 50 mV sÀ1 . (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.) Fig. 8. Inhibition of in vitro cancer cells growth by [Cu(phen)(L-tyr) BPEI]ClO4 (x = 0.182). 64 J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67
  • 7. Control 24h 48h 0 20 40 60 80 100%ofcells Normal Apoptosis Necrosis Fig. 9. Photomicrographs of control (the cells were viable as inferred from the green – fluorescence) and AO/EB stained MCF-7 cancer cells treated with the [Cu(phen)(L- tyr)BPEI]ClO4 (x = 0.182) at 20.4 and 14.3 lg mlÀ1 concentration for 24 and 48 h. Scale bar: 35 lm. The graph shows data on percentage of cells that are normal afflicted with apoptosis and necrosis in the control and 24 h and 48 h treatment groups. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.) Control 24 h 48 h 0 20 40 60 80 100 120 140 %ofcells Dead Highly Damaged Damaged Slightly Damaged Intact Fig. 10. Comet images of DNA double strand breaks at 12 and 24 h treatment of [Cu(phen)(L-tyr)BPEI]ClO4 (x = 0.182) at 20.4 and 14.3 lg mlÀ1 concentration. Cells were grown in RPMI1640 medium containing FBS at 10% final concentration, and streptomycin (10 mg mlÀ1 ) and penicillin (10,000 IU mlÀ1 ) as antibiotics. The duration- dependence of the DNA damage is revealed. Scale bar: 35 lm. DNA damage in MCF-7 cell populations as defined according to the percentage of DNA in the tail. J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67 65
  • 8. shown in Fig. 10, the images were used to estimate the DNA con- tent of individual nuclei and to evaluate the degree of DNA damage representing the fraction of total DNA in the tail. Cells were assigned to five groups: 0–20% (intact), 20–40% (slightly damaged), 40–60% (damaged), 60–80% (highly damaged) and >80% (dead). The results revealed that DNA damage was induced in MCF-7 can- cer cells by the polymer–copper(II) complex, and the incidence was greater at 48 h than at 24 h, as shown in Fig. 10. 5. Conclusions Water soluble polyethyleneimine coordinated–copper(II) com- plexes containing phenanthroline and L-tyrosine as co-ligands with various degrees of coordination were synthesised. The complexes were characterized adopting various spectroscopic techniques and elemental analysis. The binding between the polymer–cop- per(II) complexes and DNA was assessed in relation to the polymer complex with different degrees of copper complex content in the polymer chain. The electronic absorption spectral studies, emission studies and ionic strength effect showed that these complexes bind to DNA via electrostatic modes of binding. These studies indicates that the binding affinity toward DNA increases with the increase in the number of copper centres in the polymer. The changes in cir- cular dichroism and cyclic voltammetry studies of the binding between one of our complexes in the presence of DNA confirm the above mentioned modes of binding. Thermal denaturation studies of the binding between our complexes and DNA reveal that the complex with higher degree of coordination binds with DNA and stability enhanced. The polymer–copper(II) complex of the highest degree of coordination showed good cytotoxic activity against MCF-7 cancer cell with mostly through apoptosis although a few cells succumbed to necrosis. Acknowledgments We are grateful to the UGC-SAP and DST-FIST programmes of the Department of Chemistry, Bharathidasan University. Council of Scientific and Industrial Research (CSIR), New Delhi is acknowl- edged for financial support [Scheme. No. 09/475(0154)/2010-EMR- I dated. 09/02/2011] for Senior Research Fellowship to JLP. One of the authors, SA., thanks for sanction of research schemes, Grant No. SR/S1/IC-13/2009 of DST, Grant No. 01(2461)/11/EMR-II of CSIR and also Grant No. 41-223/2012(SR) of UGC. Grants from Doerenk- amp-Zbinden Foundation, Switzerland, and King Saud University, Riyadh, Kingdom of Saudi Arabia to MAA are gratefully acknowledged. References [1] J.B. Chaires, Drug-DNA interactions, Curr. Opin. Chem. Biol. 8 (1998) 314–320. [2] K.J. Du, J.Q. Wan, J.F. Kou, G.Y. Li, L.L. Wang, H. Chao, L.N. Ji, Synthesis, DNA- binding and topoisomerase inhibitory activity of ruthenium(II) polypyridyl complexes, Eur. J. Med. Chem. (2011) 1056–1065. [3] S. Roy, K.D. Hagen, U. Maheswari, M. Lutz, A.L. Spek, J. Reedijk, G.P. van Wezel, Phenanthroline derivatives with improved selectivity as DNA-targeting anticancer or antimicrobial drugs, Chem. Med. Chem. 3 (2008) 1427–1434. [4] P. Borst, S. Rottenberg, J. Jonkers, How do real tumors become resistant to cisplatin?, Cell Cycle 15 (2008) 1353–1359 [5] X. Wang, Fresh platinum complexes with promising antitumor activity, Anticancer Agents Med. Chem. 10 (2010) 396–411. [6] W.H. Ang, P.J. Dyson, Classical and non-classical ruthenium-based anticancer drugs: towards targeted chemotherapy, Eur. J. Inorg. Chem. 2006 (2006) 4003– 4018. [7] X.-L. Zhao, Z.-S. Li, Z.-B. Zheng, A.-G. Zhang, K.-Z. Wang, PH luminescence switch, DNA binding and photocleavage, and cytotoxicity of a dinuclear ruthenium complex, Dalton Trans. 42 (2013) 5764–5777. [8] R. Chadha, V.K. Kapoor, D. Thakur, R. Kaur, P. Arora, D.V.S. Jain, Drug carrier systems for anticancer agents: a review, J. Sci. Ind. Res. 67 (2008) 185–197. [9] B.N. Ames, M.K. Shigenaga, T.M. Hagen, Oxidants, antioxidants, and the degenerative diseases of aging, Proc. Natl. Acad. Sci. U.S.A. 90 (1993) 7915–7922. [10] D.S. Sigman, Chemical nucleases, Biochemistry 29 (1990) 9097–11105. [11] G. Majella, S. Vivienne, M. Malachy, D. Michael, M. Vickie, Synthesis and anti- Candida activity of copper(II) and manganese(II) carboxylate complexes: X-ray crystal structures of [Cu(sal)(bipy)]ÁC2H5OHÁH2O and [Cu(norb)(phen)2]Á 6.5H2O (salH2 = salicylic acid; norbH2 = cis-5-norbornene-endo-2,3- dicarboxylic acid; bipy = 2,2-bipyridine; phen = 1,10-phenanthroline), Polyhedron 18 (1999) 2931–2939. [12] J.D. Saha, U. Sandbhor, K. Shirisha, S. Padhye, D. Deobagkar, C.E. Ansond, A.K. Powell, A novel mixed-ligand antimycobacterial dimeric copper complex of ciprofloxacin and phenanthroline, Bioorg. Med. Chem. Lett. 14 (2004) 3027– 3032. [13] S. Zhang, Y. Zhu, C. Tu, H. Wei, Z. Yang, L. Lin, J. Ding, J. Zhang, Z. Guo, A novel cytotoxic ternary copper(II) complex of 1,10-phenanthroline and l-threonine with DNA nuclease activity, J. Inorg. Biochem. 98 (2004) 2099–2106. [14] Y. Ni, D. Lin, S. Kokot, Synchronous fluorescence, UV–visible spectrophotometric, and voltammetric studies of the competitive interaction of bis(1,10-phenanthroline)copper(II) complex and neutral red with DNA, Anal. Biochem. 352 (2006) 231–242. [15] T. Gupta, S. Dhar, M. Nethaji, A.R. Chakravarty, Bis(dipyridophenazine)copper(II) complex as major groove directing synthetic hydrolase, Dalton Trans. (2004) 1896–1900. [16] A.D. Burrows, C.W. Chan, M.M. Chowdhry, J.E. McGrady, D.M.P. Mingos, Multidimensional crystal engineering of bifunctional metal complexes containing complementary triple hydrogen bonds, Chem. Soc. Rev. 24 (1995) 329–339. [17] M. Chauhan, K. Banerjee, F. Arjmand, DNA binding studies of novel copper(II) complexes containing L-tryptophan as chiral auxiliary. In vitro antitumor activity of Cu–Sn2 complex in human neuroblastoma cells, Inorg. Chem. 46 (2007) 3072–3082. [18] B.L. Rivas, K.E. Geckeler, Synthesis and metal complexation of poly(ethylenimine) and derivatives, Adv. Polym. Sci. 102 (1992) 171–188. [19] E. Tsuchida, H. Nishide, T. Ohkawa, Bulletin of the science and engineering research laboratory, Waseda Univ. 69 (1975) 31–34. [20] R.S. Kumar, K. Sasikala, S. Arunachalam, DNA interaction of some polymer– copper(II) complexes containing 2,20 -bipyridyl ligand and their antimicrobial activities, J. Inorg. Biochem. 102 (2008) 234–241. [21] J. Lakshmipraba, S. Arunachalam, Studies on the interactions of polymer- anchored copper(II) complexes with tRNA, Transition Met. Chem. 35 (2010) 477–482. [22] J. Lakshmipraba, S. Arunachalam, D.A. Gandi, T. Thirunalasundari, Synthesis, nucleic acid binding and cytotoxicity of polyethyleneimine–copper(II) complexes containing 1,10-phenanthroline and L-valine, Eur. J. Med. Chem. 46 (2011) 3013–3021. [23] J. Lakshmipraba, S. Arunachalam, A. Riyasdeen, R. Dhivya, S. Vignesh, M.A. Akbarsha, R.A. James, DNA/RNA binding and anticancer/antimicrobial activities of polymer–copper(II) complexes, Spectrochim. Acta, Part A 109 (2013) 23–31. [24] T. Sugimori, H. Masuda, N. Ohata, K. Koiwai, A. Odani, O. Yamauchi, Structural dependence of aromatic ring stacking and related weak interactions in ternary amino acid–copper(II) complexes and its biological implication, Inorg. Chem. 36 (1997) 576–583. [25] T. Lundback, T. Hard, Sequence-specific DNA-binding dominated by dehydration, Proc. Natl. Acad. Sci. 93 (1996) 4754–4759. [26] R. Marty, C.N. N’soukpoé-Kossi, D.M. Charbonneau, L. Kreplak, H.-A. Tajmir- Riahi, Structural characterization of cationic lipid-tRNA complexes, Nucleic Acids Res. (2009) 1–11. [27] T. Mosmann, Rapid colorimetric assay for cellular growth and survival: application to proliferation and cytotoxicity assays, J. Immunol. Methods 65 (1983) 55–63. [28] D.L. Spector, R.D. Goldman, L.A. Leinwand, Cell: A Laboratory Manual, Culture and Biochemical Analysis of Cells, vol. 1, Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York, 1998, pp. 34.1–34.9. [29] N.P. Singh, M.T. MacCoy, R.R. Tice, E.L. Schneider, A simple technique for quantitation of low levels of DNA damage in individual cells, Exp. Cell 175 (1988) 184–191. [30] Y. Kurimura, E. Tsuchida, M. Kaneko, Preparation and properties of some water soluble Co(III)-poly-4-vinylpyridine complexes, J. Polym. Sci., Part A-1 9 (1971) 3511–3519. [31] J.W. Kolis, D.E. Hamilton, N.K. Kildahl, Nickel(II) and copper(II) complexes of linear pentadentate ligands derived from substituted o-hydroxyaceto- and o- hydroxybenzophenones, Inorg. Chem. 18 (1979) 1826–1831. [32] A.A. Schilt, R.C. Taylor, Infrared spectra of 1,10-phenanthroline metal complexes in the rock salt region below 2000 cmÀ1 , J. Inorg. Nucl. Chem. 9 (1959) 211–221. [33] K.E. Geckeler, G. Lange, H. Eberhardt, E. Bayer, Preparation and application of water-soluble polymer–metal complexes, Pure Appl. Chem. 52 (1980) 1883– 1905. [34] T. Biver, B. García, J.M. Leal, F. Secco, E. Turriani, Left-handed DNA: intercalation of the cyanine thiazole orange and structural changes. A kinetic and thermodynamic approach, Phys. Chem. Chem. Phys. 12 (2010) 13309– 13317. [35] Q.-Q. Zhang, F. Zhang, W.-G. Wang, X.-L. Wang, Synthesis, crystal structure and DNA binding studies of a binuclear copper(II) complex with phenanthroline, J. Inorg. Biochem. 100 (2006) 1344–1352. [36] Y.-I. Zhou, Y.-Z. Li, The interaction of poly(ethylenimine) with nucleic acids and its use in determination of nucleic acids based on light scattering, Spectrochim. Acta, Part A 60 (2004) 377–384. 66 J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67
  • 9. [37] M.J. Waring, Complex formation between ethidium bromide and nucleic acids, J. Mol. Biol. 13 (1965) 269–282. [38] J.G. Liu, B.H. Ye, H. Li, L.N. Ji, R.H. Li, J.Y. Zhou, Synthesis, characterization and DNA-binding properties of novel dipyridophenazine complex of ruthenium (II): [Ru(IP)2(DPPZ)]2+ , J. Inorg. Biochem. 73 (1999) 117–122. [39] S. Ramakrishnan, M. Palaniandavar, Interaction of rac-[Cu(diimine)3]2+ and rac-[Zn(diimine)3]2+ complexes with CT DNA: effect of fluxional Cu(II) geometry on DNA binding, ligand-promoted exciton coupling and prominent DNA cleavage, Dalton Trans. (2008) 3866–3878. [40] N. Shahabadi, S. Kashanian, M. Khosravi, M. Mahdavi, Multispectroscopic DNA interaction studies of a water-soluble nickel(II) complex containing different dinitrogen aromatic ligands, Transition Met. Chem. 35 (2010) 699–705. [41] S. Kashanian, M.M. Khodaei, H. Roshanfekr, N. Shahabadi, G. Mansouri, DNA binding, DNA cleavage and cytotoxicity studies of a new water soluble copper(II) complex: the effect of ligand shape on the mode of binding, Spectrochim. Acta – Part A 86 (2012) 351–359. [42] S. Kashanian, M.M. Khodaei, H. Roshanfekr, H. Peyman, DNA interaction of [Cu(dmp)(phen-dion)] (dmp = 4,7 and 2,9 dimethyl phenanthroline, phen-dion = 1,10-phenanthroline-5,6-dion) complexes and DNA-based electrochemical biosensor using chitosan-carbon nanotubes composite film, Spectrochim. Acta – Part A 114 (2013) 642–649. [43] M.T. Record Jr., T.M. Lohman, P. De Haseth, Ion effects on ligand– nucleic acid interactions, J. Mol. Biol. 107 (1976) 145–158. [44] M.X. Tang, F.C. Szoka, The influence of polymer structure on the interactions of cationic polymers with DNA and morphology of the resulting complexes, Gene Therapy 4 (1997) 823–832. [45] K. van Holde, W.C. Johnson, P.S. Ho, Principles of Physical Biochemistry, Prentice Hall, New York, 1998. [46] S. Zhang, Y. Zhu, C. Tu, H. Wei, Z. Yang, L. Lin, J. Ding, J. Zhang, Z. Guo, A novel cytotoxic ternary copper(II) complex of 1,10-phenanthroline and L-threonine with DNA nuclease activity, J. Inorg. Biochem. 98 (2004) 2099–2106. [47] T. Carter, M.A. Rodriguez, A.J. Bard, Voltammetric studies of the interaction of metal chelates with DNA. 2. Tris-chelated complexes of cobalt(III) and iron(II) with 1,10-phenanthroline and 2,2’-bipyridine, J. Am. Chem. Soc. 111 (1989) 8901–8911. [48] M.T. Carter, A.J. Bard, Voltammetric studies of the interaction of tris(1,10- phenanthroline)cobalt (III) with DNA, J. Am. Chem. Soc. 109 (1987) 7528– 7530. [49] A.-M. Florea, D. Büsselberg, Cisplatin as an anti-tumor drug: cellular mechanisms of activity, drug resistance and induced side effects, Cancers 3 (2011) 1351–1371. [50] G.J. Brewer, Copper control as an antiangiogenic anticancer therapy: lessons from treating Wilson’s disease, Exp. Biol. Med. (Maywood) 226 (2001) 665– 673. [51] J.P. Behr, Gene transfer with synthetic cationic amphiphiles. Prospects for gene therapy, Bioconjug. Chem. 5 (1994) 382–389. [52] B. Coyle, P. Kinsella, M. McCann, M. Devereux, R. O’Connor, M. Clynes, K. Kavanagh, Induction of apoptosis in yeast and mammalian cells by exposure to 1,10-phenanthroline metal complexes, Toxicol. In Vitro 18 (2004) 63–70. [53] B. Coyle, M. McCann, K. Kavanagh, M. Devereux, M. Geraghty, Mode of anti- fungal activity of 1,10-phenanthroline and its Cu(II), Mn(II) and Ag(I) complexes, BioMetals 16 (2003) 321–329. [54] M. McCann, M. Geraghty, M. Devereux, D. O’Shea, J. Mason, L. O’Sullivan, Insights into the mode of action of the anti-Candida activity of 1,10-phenanthroline and its metal chelates, Metal-Based Drugs 7 (2000) 185–193. [55] P. Moller, L.E. Knudsen, S. Loft, H. Wallin, The comet assay as a rapid screen test in biomonitoring occupational exposure and effect of confounding factors, Cancer Epidemiol. Biomarkers Prevent. 9 (2000) 1005–1015. [56] M.A. Maire, C. Rast, Y. Landkocz, P. Vasseur, 2,4-Dichlorophenoxyacetic acid: effects on Syrian hamster embryo (SHE) cell transformation, c-Myc expression, DNA damage and apoptosis, Mutat. Res. 631 (2007) 124–136. J. Lakshmipraba et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 59–67 67