You are on page 1of 12

Inorganica Chimica Acta 467 (2017) 264–275

Contents lists available at ScienceDirect

Inorganica Chimica Acta


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

Research paper

Synthesis, physicochemical properties, thermal analysis and biological


application of phosphorescent cationic iridium(III) complexes
V. Thamilarasan a, V. Sethuraman a, P. Karunakaran b, M. Sethupathi a, P. Manisankar a, C. Selvaraju b,
N. Sengottuvelan a,⇑
a
Department of Industrial Chemistry, Alagappa University, Karaikudi 630 003, India
b
National Centre for Ultrafast Processes, University of Madras, Taramani Campus, Chennai 600 113, India

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

Article history: Cyclometalating iridium(III) complexes with formula, [Ir(cpy)2(biim)Cl] 1, [Ir(cpy)2(bpy)Cl] 2 and
Received 22 June 2017 [Ir(cpy)2(dppe)Cl] 3 (where, cpy = 9-(pyridine-2-yl)-9H-carbazole, biim = biimidazole, bpy = 2,20 -bipyri-
Received in revised form 26 July 2017 dine, dppe = 1,2-bis(diphenylphosphino)ethane) were synthesized and characterized by elemental anal-
Accepted 30 July 2017
ysis, 1H, 13C NMR, and ESI-Mass spectral studies. The complex 1–3 show their emissions in green emitting
Available online 1 August 2017
region (kPL,max = 536–580 nm), and the emission maxima can be tuned by changing the ancillary ligand.
The phosphorescent line shape indicates that the emissions originate predominantly from 3MLCT states
Keywords:
with little admixture of ligand-based 3(p-p⁄) excited states. These complexes exhibited biexponential
Iridium(III) complexes
Photophysical
phosphorescence decays with relatively short lifetimes of 0.40–12.35 ms in solution state at room temper-
Electrochemical properties ature and the relaxation dynamics of 1–3 are O2 dependent. The electrochemical and thermal properties
DNA were systematically evaluated. Binding of these complexes (1–3) with calf thymus DNA was investigated
Protein interaction by UV–Vis, fluorescence, circular dichroic spectroscopy, viscosity measurements and computer aided
Cytotoxic activities molecular docking studies. The binding constant of complexes (1–3) with ct-DNA obtained by UV–Vis
absorption studies were 2.28  104, 1.98  104 and 2.98  104 M1, respectively. The result indicated that
the complexes (1–3) were able to bind to DNA with different binding affinity in the order: 3 > 1 > 2.
Complexes (1–3) also exhibit a good binding propensity to bovine serum albumin (BSA). Gel elec-
trophoresis assay demonstrated that complexes (1–3) cannot promote the cleavage ability of the
pBR322 plasmid DNA or plasmid pBluescript II KS+ DNA or pUC19 DNA in the presence of the reducing
agent 3-mercaptopropionic acid. The cytotoxicity studies of complexes 1–3 were tested in vitro on in
human cervical cancer cell line (HeLa) and they found to be active. We propose that the phosphorescent
iridium(III) complexes (1–3) are possible candidates for the green phosphors in organic light emitting
diodes (OLEDs) applications as well as better anticancer agents.
Ó 2017 Elsevier B.V. All rights reserved.

1. Introduction ties than their free analogs due to changes in the rigidity and
hydrophobicity of the surrounding environment [14]. Thus, lumi-
The photoluminescence of organometallic iridium(III) com- nescent labels provide a facile method for monitoring bioconjuga-
plexes has attracted much interest since it can be utilized for a tion reactions and quantifying the binding affinity between
variety of applications such as oxygen sensors, biological probes different substrates [15,16]. This has been achieved by screening
and phosphorescent dopants in optoelectronic devices [1–12]. In a large variety of cyclometalating and ancillary ligands, often bear-
particular, cyclometalating complexes of iridium(III) have received ing functional groups with electron withdrawing or releasing prop-
great attention because of the high tunability of their emission in erties. The strong spin-orbital coupling effect of the heavy metal,
terms of color and efficiency. Luminescent iridium(III) complexes these complexes are capable of harvesting both singlet and triplet
can be used to label biomaterials due their ability to either excitons in devices to achieve internal quantum efficiency of 100 %,
covalently [13] or noncovalently [14] bind biological substrates. when acting as the emissive material [17]. Cyclometalating car-
Bound complexes typically express different luminescent proper- bazole based iridium complexes exhibiting green emission, having
unique hole transport ability, charge carrier mobility and good
⇑ Corresponding author at: DDE, Chemistry, Alagappa University, Karaikudi 630 thermal stability which play a very important role in the electroac-
003, Tamil Nadu, India. tive and photoactive materials [18–23].
E-mail address: nsvelan1975@yahoomail.com (N. Sengottuvelan).

http://dx.doi.org/10.1016/j.ica.2017.07.061
0020-1693/Ó 2017 Elsevier B.V. All rights reserved.
V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275 265

Herein we report the synthesis of three carbazole based cationic


heteroleptic cyclometalated iridium(III) complexes: [Ir(cpy)2(biim)
Cl] 1, [Ir(cpy)2(bpy)Cl] 2 and [Ir(cpy)2(dppe)Cl] 3. The complexes
were characterized by various physicochemical techniques. Further
we describe the color tuning of iridium complexes by varying the
ancillary ligand units in complexes which may alter its photophys-
ical properties which could offer new biological sensing possibili-
ties, hence the complexes has been found to significantly perturb
its biomolecular binding, DNA cleavage and cytotoxic activities. In
the present biological study, the DNA-binding behaviors of three
Ir(III) complexes were explored by absorption, emission, circular
dichroism spectroscopy and viscosity measurements and their
affinity for bovine serum albumin (BSA) proteins involved in the
transport of metal ions and metal complexes with drugs through
the blood stream, performed by fluorescence spectroscopy. Their
abilities to induce cleavage of pBR322 DNA or plasmid pBluescript
II KS+ DNA or pUC19 DNA and in vitro cytotoxicity on human cervi-
cal cancer cell line (HeLa) were explored.

Fig. 1a. UV–Vis spectra of complexes (1–3).


2. Results and discussion

The Ir(III) l-chloro bridged dimer was synthesized by the reac- 2.1. UV–Vis spectroscopy
tion of iridium trichloride hydrate with ligand cpy. Then diiridium
complexes was converted to mononuclear iridium(III) complexes The UV–Visible absorption spectra (Fig. 1a) of iridium(III) com-
such as [(cpy)2Ir(biim)Cl] 1, [(cpy)2Ir(bpy)Cl] 2, and [(cpy)2Ir plexes 1–3, show absorption band around 220–260 nm is assigned
(dppe)Cl] 3 with bidentate ligands biimidazole, 2,20 -bipyridine to a typical ligand-centered states with mostly spin-allowed 1p-p⁄
and 1,2-bis(diphenylphosphino)ethane, respectively in 41–52% transition of the cpy ligand because the corresponding transition
yield (Scheme 1). 1H NMR, 13C NMR, FT-IR, mass and elemental was also observed in the free ligand cpy based UV spectral profile
analyses of complexes (1–3) are consistent with the proposed (Table 1). On the other hand, the band in the region 300–360 nm,
structure of complexes (1–3). which is not observed in the spectrum of the free ligand, is

Scheme 1.
266 V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275

Table 1
Photophysical parameters of iridium(III) complexes (1–3).

Complexes Absorption kabs/nm (e, 103 M1 cm1) kmaxphos (nm) s1 s2


1 253(1077), 290(754), 345(300), 466(87), 485(11) 425,580 631 ns 972 ns
899 ns* 546 ns*
2 227(708), 310(602), 349(299), 386(330), 451(195) 463,565 403 ns 1.08 ls
716 ns* 3.35 ls*
3 231(665), 309(252), 407(98), 487(28) 456, 536 1.51 ls 3.02 ls
5.08 ls* 12.35 ls*
*
Degassed.

assignable to the singlet metal-to-ligand charge-transfer (1MLCT) the expected blue-shift of the spectral characteristics. To obtain a
transition. The moderately weaker absorptions above 380 nm can better picture of the excited state quenching processes in these
be assigned to a spin-forbidden triplet metal-to-ligand charge- materials, we have studied the photoluminescence decay tran-
transfer (3MLCT) transition [24]. sients using time-correlated single photon counting (TCSPC) has
been studied. It is worth noting that the nitrogen-degassed CH2Cl2
of complexes exhibit fluorescence decays were multiexponential
2.2. Photoluminescence spectra and lifetime measurement
and the average lifetimes are reported in the Table 1. As listed in
Table 1, the relaxation dynamics of 1–3 are O2 dependent. For
The photoluminescence spectra of the complexes (1–3) in
example, the decay time of complex 1 was measured to be
dichloromethane at room temperature are shown in Fig. 1b and
546 ns (s2) in the degassed CH2Cl2, while it was decreased to
the corresponding photophysical data of the iridium complexes
972 ns upon aeration. The result is consistent with the O2-quench-
are summarized in Table 1. Complexes 1–3 exhibits duel intense
ing triplet state according to the theory of electron-exchange type
PL emission at [425,580], [463,565] and [456,536] nm, respectively
(Dexter type) of energy transfer expressed as [27],
[25,26]. The pronounced dual emission band for complexes shows
that there is an excitation of two separated LC transitions involving T þ 3 O2 !ko2 S0 þ 1 O2
two ligand subunits. As depicted in Fig. 1b, complexes 1–3 revealed
two distinct bands, which varied according to the anionic ancillary
ligands. The shorter-wavelength band, showing characteristics of a 2.3. Correlation of absorption, emission and lifetime measurements
short lifetime, is classified as fluorescence, while the longer wave-
length band can be assigned to the phosphorescence on the basis of The MLCT absorption bands above 380 nm, exhibit good spec-
its longer lifetime and drastic quenching effect under oxygen. tral overlap with the PL emission band [28]. This good spectral
It is believed that both of these emissions are dominated by p- overlap indicated that an efficient Forster or Dexter energy transfer
p⁄ transition in combination with some amount of ILCT character, from the host (poly(vinylcarbazole) (PVK), to the iridium com-
the latter, in part, incorporates the transfer of electron densities plexes guests would be expected in the electroluminescent device.
from the ancillary to main ligand system. In the present work, sig- The absorption bands above 420 nm are correlated to the photolu-
nificant emission color tuning was realized (Fig. 1b); the emission minescence (PL) of the complexes. It is notable that the absorption
wavelength could be tuned from 536 to 580 nm. The higher wave- edges are blue-shifted gradually in the order 3 < 2 < 1. The PL peaks
length emission maxima of 3 is 44 nm blue-shifted compared to 1, of the complexes are also blue-shifted gradually in the order
reflecting that the p-accepting character of the phosphine group 3 < 2 < 1 that are in good consistent with the absorption spectra.
increased the associated MLCT energy levels, resulting in a blue- Further the luminescent lifetimes fall between 1 and 12 ls, consis-
shift emission. The spectroscopic results reveal an increased possi- tent with emission from a triplet excited state. The positions of the
bility of color-tuning through changing the ancillary ligand with maximums in the excitation spectra for these complexes are very
similar to those in their absorption spectra. The results, in combi-
nation with the rather small radiative decay rate of <106 s1, leads
to unambiguously assign the dominant emission in complexes (1–
3) to be of phosphorescence character. The structured emission
spectra and long phosphorescence lifetime (Fig. 2) (described
below) suggest that the lowest triplet state of complexes (1–3) also
has ligand-centered 3p-p⁄ character (3p?p⁄ (N^C and N^N)) with
mixing of 3MLCT (dp (Ir) ? p.

2.4. Electrochemical properties

The electrochemical behavior of complexes were investigated


by cyclic voltammetry (Fig. 3) in CH2Cl2 solutions with ferroce-
nium/ferrocene (Fc+/Fc) redox couple as an internal reference and
the results are given in Table 2. The complexes 1–3 exhibit
quasi-reversible reduction processes, with reduction potentials of
the complexes stay in a range of 0.96 to 1.02. The complexes
(1–3) display a quasi-reversible one-electron oxidation couple in
the range 1.10–1.14 V assigned to a IrIII/IrIV oxidation. These values
are similar to the reported bis-cyclometalated iridium complexes
[29–31]. The redox peaks potential are almost in the same region.
Fig. 1b. PL spectra of complexes (1–3): ( ) aerated solution, ( ) argon The reduction peak can be assigned to the reduction of the pyridine
degassed solutions. heterocyclic portion of the ligands. As revealed previously by
V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275 267

Eox is the onset oxidation potential and the energy band gap
(DEgap) of the iridium(III) complexes deduced from the absorption
edge of the absorption spectra from DEgap + 1240/kedge, where kedge
is the onset value of the absorption spectrum. The LUMO energy
levels were estimated from the HOMO values and optical band
gaps (DEgap) by ELUMO = EHOMO + DEgap. Table 2 summarizes the
electrochemical data and the energy levels of the Ir(III) complexes.

2.5. Thermal properties

The thermal properties of the iridium complexes 1–3 were


examined by thermogravimetric analysis (TGA) under a nitrogen
atmosphere at a heating rate of 10 min (Fig. S1). The iridium(III)
complexes exhibited 5 % weight loss at 298 °C for 1, 293 °C for 2
and 276 °C for 3, which is beneficial to the long-term stability
and suitability for PhOLEDs device applications.

Fig. 2. Lifetime measurements of complexes (1–3): ( ) aerated solution,


( ) argon degassed solutions.
2.6. DNA binding studies

2.6.1. Absorption spectral studies


Transition metal complexes can bind with DNA via both cova-
lent (via replacement of a labile ligand of the complex by a nitro-
gen base of DNA such as guanine N7) and/or noncovalent
(intercalation, electrostatic or groove binding) interactions
[36,37]. Absorption titration can be used to observe the interaction
of molecules with DNA. The absorption spectra of complexes (1–3)
in the absence and presence of CT-DNA (at a constant concentra-
tion of complex 100 mM) in 2% DMSO at 25 °C by varying the con-
centration of DNA (buffer = 7.2 pH) are shown in Fig. 4. When the
amount of DNA is increased the intensity of the charge transfer
band is also changed. In the UV spectrum of 1, the band at
346 nm exhibited hyperchromism of 9.33 % and blue shifts of
1 nm suggesting the tight binding to CT-DNA. Additionally, the
hyperchromism of band are accompanied by a blue-shift of 1 nm
which is an indicative of stabilization. The behavior of complexes
2 and 3 is quite similar to 1 (hyperchromism and blue-shift). Com-
plexes 2 and 3 exhibited the hyperchromism of 8.98 and 10.21% at
344 and 335 nm, respectively with the blue shifts. The hyper-
chromism of complexes 1–3 indicate the DNA-stabilization and
the strong binding to CT DNA, possibly due to electrostatic interac-
tion or to the partial uncoiling of DNA helix structure, exposing
more bases of DNA [38]. Moreover, hyperchromic effect reflects
Fig. 3. Cyclic voltammograms of complexes (1–3).
the corresponding changes in the secondary structure of DNA in
its conformation after the complex-DNA interaction. The results
electrochemistry and theoretical calculations [32–35] of cyclomet- derived from the UV titration experiments suggest that all the
alated iridium(III) complexes, the reduction occurs primarily on complexes can bind to CT-DNA, although the exact mode of bind-
the more electron accepting heterocyclic portion of the cyclomet- ing cannot be merely proposed by UV spectroscopic titration stud-
alated cpy ligands (LUMO contribution) whereas the oxidation pro- ies. The existence of hypochromism could be considered as
cess is to largely involve in the Ir-cpy center (HOMO contribution). evidence for the binding of the complexes involving intercalation
Consistent with this conclusion, these complexes have similar between the base pairs of CT-DNA cannot be ruled out [39]. These
LUMO energy levels due to the same pyridyl based frame, which results indicate that these complexes interact with DNA through
makes the reduction potential of these complexes stay in a narrow groove binding.
range. The HOMO and LUMO level of complexes were calculated to The kmax, hyperchromism, blue shift and binding constant val-
be 5.63 to 5.71 eV and 4.04 to 4.06 eV respectively. ues of all the complexes with CT-DNA (Table 3) indicate that there
The oxidation potential of the Ir(III) complexes was translated was a finite interaction between these complexes and CT-DNA. In
into the HOMO using the equation, EHOMO = (Eox + 4.8) eV, where order to compare the affinity of the complexes toward DNA

Table 2
Electrochemical properties of the iridium(III) complexes (1–3).

Complexes Eox (V) Ered (V) Eox


onset (V) HOMO (eV) LUMO (eV) Egap (eV)
1 1.11 0.96 0.83 5.63 4.04 1.59
2 1.14 1.02 0.91 5.71 4.06 1.65
3 1.10 1.02 0.89 5.69 -4.05 1.64
268 V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275

Fig. 4. Absorption spectra of complexes (1–3) in 5 mM Tris-HCl/50 mM NaCl buffer upon addition of DNA. Arrow shows the absorbance changing upon increase of DNA
concentration. The inner plot of [DNA/(ea  ef) vs [DNA] for the titration of DNA with complexes.

EB at 598 nm by various complex concentrations are shown in


Table 3
The binding constant values of complexes (1–3). Fig. S2. Upon the addition of complexes, a significant decrease in
the fluorescence intensity of the EB-DNA system occurred, indicat-
Complexes kmax Hyperchromism H (%) Binding constant Kb (M1)
ing the binding of complexes 1–3 to DNA. These results indicate
(nm)
that complexes (1–3) could partially displace EB from the DNA-
1 346 9.33 2.28  104 EB system, as often observed in intercalative complex-DNA mode.
2 344 8.98 1.98  104
3 335 10.21 2.98  104
Addition of varying concentrations of complexes (1–3) to a
solution of EB-bound to CT-DNA causes lowering of the initial
emission intensity (51%, 1; 33%, 2; and 73%, 3) at diverse ‘r’ values
(Fig. S3), showing that Ir(III) complexes with EB to bind with DNA.
quantitatively, the binding constant, Kb has been determined from
The spectral data indicate that complexes (1–3) efficiently displace
the spectroscopic titration date using the equation [40].
EB and bind to CT-DNA through intercalation [46–48]. Binding
½DNA=ðea  ef Þ ¼ f½DNA=ðeb  ef Þg þ 1=kb ðeb  ef Þ strength of complexes (1–3) with CT-DNA has been estimated
using the Stern-Volmer equation:
where ea is the extinction coefficient observed for the charge trans-
fer absorption at a given DNA concentration, ef the extinction coef- I0
¼ 1 þ Ksv ½Q 
ficient of the free complex in solution, eb the extinction coefficient I
of the complex when fully bound to DNA, Kb the equilibrium bind- where, I0 and I are emission intensity in absence and presence of the
ing constant, and [DNA] gives Kb as the ratio of the slope to the complexes. Ksv is a linear Stern-Volmer quenching constant, [Q] is
intercept. The complex 3 serves as better DNA binding agent with the ratio of the total concentration of complex to that of DNA. The
efficient activity compared with that of the other complexes, which quenching plot illustrates that the quenching of EB bound to DNA
suggests that the interaction of the complexes with DNA is strong by the iridium(III) complexes is in good agreement with the linear
and through groove binding. In the complex 3, the nature of the Stern-Volmer equation, which also indicated that the complexes
metal phosphorus bond has a strong influence on the electron den- binds to DNA.
sity on the metal center. It is possible to distinguish between pure In the plot of I0/I vs [Complex]/[DNA], Ksv is given by the ratio of
r-donor ligands and r-donor/p-acceptor ligands within a family the slope to intercept (Fig. S4) and resulting Ksv values for be
of complexes by a correlation of two appropriate properties of a complexes 1–3 are 2.17  103, 4.83  103, 3.63  103 and
family of complexes when these properties are sensitive to the elec- 5.71  103 M1, respectively which varies in the order: 3 > 1 > 2.
tron density of the metal [41]. As a result, the binding interaction of The results clearly suggested that 3 have greater tendency to
the iridium(III) complexes with DNA follows the trend from high to replace EB relatively to other complexes. The quenching constant
low: 3 > 1 > 2. value of Ir(III) complexes (1–3) may suggest that the complexes
(1–3) have intercalative mode of binding that involves a stacking
2.6.2. Competitive binding experiments between the complex and the base pairs of DNA. The data suggest
To further confirm the interaction between the test complex that the interaction of 3 with DNA is the strongest, followed by 1,
and CT-DNA, emission experiments were carried out. Fluorescence and then 2, which is consistent with the above absorption spectral
quenching measurements can be used to monitor metal binding results.
[42]. EB emits intense fluorescent light in the presence of CT
DNA due to its strong intercalation between the adjacent DNA base 2.6.3. Circular dichroic spectral studies
pairs. It was previously reported that the enhanced fluorescence Circular dichroic spectral studies (Fig. 5) are useful in diagnos-
could be quenched by the addition of another molecule [43,44]. ing changes in the structure of DNA during interaction of metal
The study involves the addition of the complexes to DNA pre- complexes with DNA [49]. The CD spectrum of CT-DNA consists
treated with EB and the measurement of the intensity of emission. of a positive band at 275 nm due to base stacking and a negative
The results from fluorescence titration spectra have been con- band at 244 nm due to helicity, and is characteristic of DNA in
firmed to be effective for characterizing the binding mode of metal right-handed B-form. The groove binding interaction of small
complexes to DNA [45]. The binding of complexes to DNA were molecules with DNA showed a little or no perturbations on the
given by competitive binding experiments using complexes 1–3 base stacking and helicity bands, while intercalation enhances
as quenchers. The fluorescence measurement for complexes 1–3 the intensities of both bands, stabilizing the right-handed B form
showed no emission band with or without CT-DNA at ambient conformation of CT-DNA. With the addition of complexes 1–3 to
temperature. The fluorescence quenching spectra of DNA-bound the solution of CT-DNA there is increase in the intensity of the
V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275 269

tion, with one ligand intercalating into the base pairs and the other
ligand being left outside the helix [54]. The intercalative interac-
tion with DNA is related to the molecular structure, as in this com-
plex there is a little distorted plane that may lead to the weak
intercalative mode. This result also parallels the pronounced emis-
sion enhancement of the complexes, and is comparable with the
proven classical intercalator EB. On the basis of the viscosity
results, the complexes bind with DNA through the intercalation
mode. Almost all the Ir(III) complexes enhance the viscosity and
the ability of the complexes to increase the viscosity of DNA fol-
lows the order EB > 3 > 1 > 2, iridium(III) complexes proposes to
be bound to DNA by intercalation.

2.6.5. Molecular docking studies


To elucidate the mode of interaction and binding affinity dock-
ing studies have been performed on B-DNA (PDB ID: 1BNA) in pres-
ence of complexes (1–3). The study revealed that complexes
interact with DNA via an electrostatic mode involving outside edge
stacking interaction with oxygen atom of the phosphate backbone.
Fig. 5. CD spectra of CT-DNA, and the interaction with 1, 2 and 3. All the spectra The planarity of the complexes is compatible for strong p-p stack-
were recorded in 5 mM Tris–HCl/ 50 mM NaCl buffer, pH 7.2 and 25 °C. ing interactions and a better match of the complexes inside the
DNA strands. From the ensuing docked structures (Fig. 6) it is clear
that complex 1–3 fits well into the minor groove of the targeted
positive band (Fig. 5) with a blue shift of 2, 2 and 2 nm, respec- DNA and GC rich region stabilized by van der Waals interaction
tively, and decrease in the intensity of the negative band with a and hydrophobic contacts [55]. Relative binding energy of the
no significant shift observed, compared to the DNA. The increase docked structures of complex (1–3) was found to be 292.56,
in positive and a decrease in negative ellipticity indicate strong 284.32 and 295.52 eV, respectively. These are consistent with
conformational changes [50]. These changes are revealed the non- the spectral studies and indicated that 3 have greater binding affin-
intercalative mode of binding of these complexes and offer support ity relative to other complexes. Overall results suggested that there
to their groove binding nature [51,52]. is good correlation between spectroscopic results and molecular
modeling.
2.6.4. Viscosity measurements
To investigate the binding nature of the complexes to DNA, vis- 2.7. BSA binding studies
cosity measurements on the solutions of DNA incubated with the
complexes (1–3) have been carried out. In the viscosity measure- The study on the interaction of drugs and their compounds with
ments, the rate of flow of the buffer (5 mM HCl/50 mM NaCl buffer, blood plasma proteins especially with serum albumin, which is the
pH 7.2), DNA (50 mM) and DNA with the complexes at various con- most abundant protein in plasma and is involved in the transport
centrations (2–12 mM) were measured. Data are presented as (g/ of metal ions and metal complexes with drugs through the blood
g0)1/3 vs. R {R = [complex]/[DNA]} where g is the viscosity of stream, is of increasing interest [56]. Binding of these proteins
DNA in the presence of complex and g0 is the viscosity of DNA may lead to loss or enhancement of the biological properties of
alone (Fig. S5). The viscosity of DNA increases greatly with increas- the original drug, or provide paths for drug transportation [56].
ing concentration of complex, which is similar to that of the proven Bovine serum albumin (BSA) is the most extensively studied serum
intercalator EB [53]. This observation suggests that the principal albumins. BSA (containing two tryptophans, Trp-134 and Trp-212)
mode of DNA binding by complexes involve base-pair intercala- which, when excited at 295 nm exhibit a strong fluorescence

Fig. 6. Molecular model of complexes (1–3) with DNA dodecamer duplex of sequence d(CGCGAATTCGCG)2(PDB ID: 1BNA) and modeling studies were carried out using Hex
6.0 software.
270 V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275

emission at 351 nm and 343 nm, respectively, due to the trypto- the existence of a static quenching mechanism [61,62]. Using the
phan residues [57]. Scatchard equation [59]:
The interaction of complexes (1–3) with serum albumins has  
been studied from tryptophan emission-quenching experiments DI=I0 DI
¼ nK  K
(Fig. S6). The changes in the emission spectra of tryptophan in ½Q I0
BSA are primarily due to change in protein conformation, subunit
where n is the number of binding sites per albumin and K is the
association, substrate binding or denaturation. An examination of
association binding constant, K (M1), may be calculated from the
the spectrum showed a noteworthy decrease in fluorescence inten-
slope in plots (DI0/I)/[Q] vs (DI/I0) (Fig. S9) and n is given by the
sity of complexes (1–3) at 354 nm (24 % for 1, 20% for 2 and 25%
ratio of the y intercept to the slope. It is obvious (Table 4) that
for 3) (Fig. S7) and some interaction between the complexes and
the coordination of iridium(III) complexes results in K value for
BSA protein, due to possible changes in proteins secondary struc-
BSA with complex 3 exhibiting the highest K value among the other
ture of BSA indicating the binding of the complexes to BSA [58].
complexes. The Stern-Volmer plot applied for the interaction with
The Stern-Volmer and Scatchard graphs may be used in order to
BSA in Fig. S8 shows that the curves have fine linear relationships
study the interaction of a quencher with serum albumins. Accord-
(r = 0.9920–0.9968). The calculated values of Ksv and kq are given
ing to Stern-Volmer quenching equation [59,60].
in Table 4 and indicate their good BSA binding propensity with
I0 complex 3 exhibiting the highest BSA quenching ability. From the
¼ 1 þ kq s0 ½Q  ¼ 1 þ K sv ½Q 
I Scatchard graph (Fig. S8), the association binding constant to BSA
where, I0 = the initial tryptophan fluorescence intensity of BSA, of each complex has been calculated (Table 4). The ‘n’ values of
I = the tryptophan fluorescence intensity of BSA after the addition complexes (1–3) are given Table 4. Thus the, complex 3 exhibit
of the quencher, kq = the quenching rate constants of BSA, Ksv = the higher binding affinity for BSA than other complexes which is sim-
Stern-Volmer quenching constant, s0 = the average lifetime of BSA ilar to that of DNA binding studies.
without the quencher, [Q] = the concentration of the quencher
respectively, 2.8. DNA cleavage studies
K sv ¼ K q s 0 ð1Þ
The ability of complexes to cleave supercoiled DNA was deter-
and, taking as fluorescence lifetime (s0) of tryptophan in BSA at mined by agarose gel electrophoresis. When circular plasmid
around 108 s, the Stern-Volmer quenching constant (Ksv, M1) DNA in the presence of an inorganic molecule is subjected to elec-
can be obtained by the slope of the diagram I0/I vs [Q] (Fig. S8), trophoresis, relatively fast migration will be observed for the intact
and subsequently the approximate quenching constant (kq, M1 - supercoiled form (Form I). If scission occurs on one strand, the
s1) may be calculated. The calculated values of Ksv and kq for the supercoil will relax to generate a slower moving open circular form
interaction of the complexes with BSA are given in Table 4 and indi- (Form II). If both the strands are cleaved, a linear form (Form III)
cate a good BSA binding propensity of the complexes in the order: that migrates in between Form I and Form II will be generated.
3 > 1 > 2. The kq values are higher than diverse kinds of quenchers The DNA cleavage ability of complexes (1–3) were investigated

Table 4
The BSA binding constant and parameters (Ksv, kq, K, n) derived for complexes (1–3).

Complexes Ksv (M1) kq (M1 s-1) K (M1) n r


4 12
1 4.5  10 4.5  10 0.0546 1.0134 0.9922
2 3.8  104 3.8  1012 0.0352 0.4074 0.9921
3 6.8  104 6.8  1012 0.0885 1.0171 0.9968

Fig. 7. (a–e) (a) pBR322 DNA. lane 1, DNA control; lane 2, DNA + 1 (100 mM); lane 3, DNA + 2 (100 mM); lane 4, DNA + 3 (100 mM). (b). pBR322 DNA. lane 1, DNA control; lane
2, DNA + MPA (10 mM) + 1 (100 mM); lane 3, DNA + MPA (10 mM) + 2 (100 mM); lane 4, DNA + MPA (10 mM) + 3 (100 mM). (c). plasmid pBluescript II KS+ DNA. lane 1, DNA
control; lane 2, DNA + 1 (100 mM); lane 3, DNA + 2 (100 mM); lane 4, DNA + 3 (100 mM). (d). plasmid pBluescript II KS+ DNA. lane 1, DNA control; lane 2, DNA + MPA (10 mM) + 1
(100 mM); lane 3, DNA + MPA (10 mM) + 2 (100 mM); lane 4, DNA + MPA (10 mM) + 3 (100 mM). (e). pUC19 DNA. lane 1, DNA control; lane 2, DNA + 1 (100 mM); lane 3, DNA + 2
(100 mM); lane 4, DNA + 3 (100 mM); lane 5, DNA + MPA (10 mM) + 1 (100 mM); lane 6, DNA + MPA (10 mM) + 2 (100 mM); lane 7, DNA + MPA (10 mM) + 3 (100 mM).
V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275 271

Fig. 8. Cytotoxic effect of complexes (1–3) against human cervical cancer cell line (HeLa) at different concentration (A = 0.25 mM, B = 2.5 mM, C = 25 mM, D = 50 mM and
E = 100 mM).

using different DNA such as, pBR322 DNA (Fig. 7a and b), plasmid changes in the secondary structure of DNA, and their stronger
pBluescript II KS+ (Fig. 7c and d) and pUC19 DNA (Fig. 7e). No DNA binding affinity and organometallic biphosphine-based
DNA cleavage was observed for the control DNA and DNA with cyclometalated complexes that are more stable and better p
the metal complexes (1–3) or with activators 3-mercaptopropionic accepting ability [63]. Further the surfaces that are more relevant
acid (MPA). for biological processes and thus obtain new knowledge about
the processes with which cells ingest particles [64], and particles
can control the release behavior of bioactive molecules as well as
2.9. Cytotoxic activities the surface activity, therefore providing a promising strategy to
enhance the efficiency of particulate drug-delivery systems [65].
Cytotoxic potential of newly synthesized Ir(III) complexes 1–3
was investigated on human cervical cancer cell line (HeLa). The
positive results obtained from DNA binding, protein binding and 3. Conclusions
DNA cleavage studies of the iridium(III) complexes (1–3) encour-
aged us to test its cytotoxicity against a panel of human cervical In conclusion, we have synthesized three novel cyclometalated
cancer cell line (HeLa) by the MTT assay. Cytotoxic activities of iridium(III) complexes bearing cpy [(9-(pyridine-2-yl)-9H-car-
the complexes 1–3 were analysed in range (0.25–100 mM) of con- bazole] ligands and characterized. Significant mixing of the singlet
centration by means MTT assays towards HeLa cells at 48 h. On and triplet excited states is clearly observed in both the absorption
the basic of optical density, the cytotoxicity of complexes increases and emission spectra of the complexes can be fine tuned. The alter-
with increasing concentration of complexes (1–3) from moderate ation of emission wavelength correlates with the variation of
to high. The activities of the complexes were determined by MTT energy gap evaluated from the results of cyclic voltammetry. The
test in vitro and the results were expressed in terms of IC50 values. photoluminescence spectra of the complexes (1–3) in dichloro-
The relations of inhibition rates and complexes concentrations methane at room temperature shows the emission peaks in the
against human cervical cancer cells (HeLa cell) were shown in region 536–580 nm and lifetimes (complexes 1–3) are in the range
Fig. 8. As shown in Fig. 8, complex 3 has stronger inhibition ratios of microseconds (sp = 0.40–12.35 ls). The iridium complexes
than complexes 1 and 2 against tested human cervical cancer cells shows radioactive decay rate in the order of 106 s1, leads to
at low concentrations. The inhibition effects were further unambiguously phosphorescence character. The complexes show
enhanced by increasing the concentration of complexes. At the proper HOMO and LUMO energy gap, longer lifetimes and good
concentration of 100 lM, inhibition rates of the complexes (1–3) thermal stability. So, it can be used for the fabrication of good
against human cervical cancer cells reached 62.14, 27.06 and phosphorescent OLEDs materials.
93.22 %, respectively. The IC50 values of complexes (1–3) were The binding properties of the complexes with CT-DNA, investi-
>100 for 2, 9.3 for 1 and 6.8 for 3. Also, the cytotoxicity of com- gated by UV–visible, fluorescence emission and CD spectroscopy,
plexes follows order 3 > 1 > 2. The cytotoxicity of complex 3 is suggests an efficient binding ability by the groove binding mode
higher than other complexes due to their ability to effective in the order 3 > 1 > 2. Gel electrophoresis assay demonstrated the
272 V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275

ability of the complexes 1–3 cannot promote the cleavage ability in trode and an Ag/AgCl reference electrode in nitrogen atmosphere.
the presence of the reducing agent 3-mercaptopropionic acid. The The concentration of complexes was 103 M in dichloromethane
in vitro cytotoxicity of these complexes (1–3) on human cervical and Tetramethylammonium hexafluorophosphate (101 M) was
cancer cell line (HeLa) indicates that complexes have the potential used as the supporting electrode. The CD spectra were recorded
to act as effective anticancer drugs. The complex 3 serves as better on Jasco J-810 spectropolarimeter. Molar conductivity of the com-
DNA/BSA binding agents and with efficient cytotoxic activity com- plexes was measured on a Elico model SX 80 conductivity bridge
pared with that of the other complexes, which suggests that the using freshly prepared solution of the complexes in DCM. Thermo-
interaction of the complexes with DNA is strong. In complex 3, gravimetric (TG) analyses, measured from room temperature to
the nature of the metal phosphorus bond has a strong influence 1000 °C at a heating rate for 10 °C min-1, were obtained by using
on the electron density on the metal center. It is believed that a Perkin-Elmer Diamond TG/derivative thermogravimetric (DTG)
the possibility to distinguish between pure r-donor ligands and instrument.
r-donor/p-acceptor ligands within a family of complexes by a cor-
relation of two appropriate properties of a family of complexes 4.3. Molecular docking studies
when these properties are sensitive to the electron density on
the metal. Hence, from these findings, complex 3 was suggested The rigid molecular docking studies were performed by using
to demonstrate better activity and further evaluation of in vivo HEX 8.0 software [66]. The coordinates of complexes (1–3) were
anticancer activities of complex 3 is in progress. drawn using Chemsketch (http://www.acdlabs.com) and saved in
mol files. The complex structure of the B-DNA dodecamer d
(CGCGAATTCGCG)2 (PDB ID: 1BDNA) was downloaded from the
4. Material and methods
protein data bank (http://www.resb.org./pdb). All calculations
were carried out on an Intel pentium 4.2.4 GHz based machine
4.1. Materials
running MS Windows XP SP2 as operating system. Visualization
of the docked pose has done by CHIMERA (www.cgl.ucsf.edu/
Iridium(III) chloride, 2-ethoxyethanol were purchased from Alfa
chimera).
Aesar, India. 2-bromopyridine, Agarose, carbazole, 1,2-bis
(diphenylphosphino)ethane, ethidium bromide (EB) and bovine
4.4. Synthesis of complexes
serum albumin (BSA) from Sigma-Aldrich (India). Potassium car-
bonate and 2,20 -bipyridine were purchased from Merk (India).
4.4.1. [(cpy)2Ir(biim)Cl] (1)
CT-DNA, pBR322 DNA, plasmid pBluescript II KS+DNA and pUC19
The ligand [67] (cpy) (0.40 g, 1.6 mmol), IrCl3H2O (0.2 g,
DNA were purchased from Genei, India. The human cervical cancer
0.67 mmol) in a mixed solvent of 2-ethoxyethanol (10 mL) and
cell line (HeLa cells) (3-[4,5-dimethylthiazol-2-yl]2,5-diphenylte-
water (3 mL) was stirred under nitrogen at 120 °C for 24 h. Cooled
trazolium bromide) was obtained from National Centre for Cell
to room temperature, the precipitate was collected by filtration
Science (NCCS), Pune.
and washed with water, ethanol and hexane successively and then
The CT-DNA stock solution was prepared by diluting DNA to
dried in vacuum to give a cyclometalated Ir(III) m-chloro-bridged
Tris-HCl/NaCl buffer (5 mM Tris-HCl, 50 mM NaCl, pH = 7.2), and
dimer. The dimer (0.2 g, 0.17 mmol), biim (0.0771 g, 0.58 mmol)
kept at 4 °C for no longer than a week. The UV absorbance at 260
and Na2CO3 (0.18 g, 1.75 mmol) were dissolved in 2-ethoxyethanol
and 280 nm of CT-DNA solution in Tris buffer give a ratio of 1.8–
(10 mL) and the mixture was then stirred under argon at 100 °C for
1.9, indicating that the DNA was sufficiently free of protein. The
16 h. After cooling to room temperature, the precipitate was fil-
concentration of CT-DNA was determined from its absorption
tered off and washed with water, ethanol and hexane. The crude
intensity at 260 nm with a molar extinction coefficient of
product was chromatographed on silica gel using CH2Cl2 as eluent
6600 M1 cm1. Buffer solution was prepared using 2% DMSO.
to afford the desired Ir(III) complex [(cpy)2Ir(biim)Cl] as green
solid. Yield: 35%. (0.18 g). M.p. 280 °C (dec). Anal. Calc. (%) for C40-
4.2. Physical measurements H28IrN8Cl: C, 56.62, H, 3.33, N, 13.21. Found (%): C, 56.36, H, 3.47, N,
12.98. FT-IR (KBr, m, cm1 selected peaks: 3100 s, 3037 s, 1608 s,
The mass spectra of the complexes were determined at Shi- 1175 s, 762 s (br, broad; s, sharp). UV–Vis in DCM [kmax/nm(emax/-
madzu QP 2000 mass spectrometer, SAIF, Lucknow, India. Elemen- mol1 cm1)]: 253(1077), 290(754), 345(300), 466(87), 485(11).
1
tal analysis of carbon, hydrogen and nitrogen were performed on a H NMR (300 MHz, CDCl3) d (ppm): 6.15(d, J = 7.2 Hz, 1H), 6.35
Carlorerba-1106 microanalyzer. NMR spectra were measured on a (t, J = 6.5 Hz, 1H), 6.42 (m. 1H), 6.92 (m, 2H), 7.22 (m, 4H), 7.42
MERCURY-300BB spectrometer and chemical shift were referenced (m, 5H), 7.61 (m, 5H), 7.85 (m, 5H), 7.92 (m, 2H), 8.15 (m, 2H).
13
to CDCl3 as an internal standard. The electronic spectra were C NMR (75.46 MHz; CDCl3) d (ppm): 76.60, 77.22, 113.31,
recorded on a Shimadzu UV-3101PC spectrophotometer. The fluo- 115.38, 118.78, 119.25, 121.18, 123.34, 124.01, 126.07, 125.49,
rescence spectral measurements were carried out using Fluoro- 126.40, 129.88, 136.08, 138.49, 144.24, 151.62, 156.18 and
max-4 spectrophotometer (Horiba Jobin Yvon), and time-resolved 173.56. ESI-MS (CH3OH) m/z (%): 848 [M+], 813(75) [M+Cl], 680
photoluminescence decays were measured using nanosecond laser (10) [M+Cl, biim]. Conductivity (KM/S cm2 mol1) in DMF: 55.
flash photolysis (Applied photo physics) setup. The third harmon- The complexes [(cpy)2Ir(bpy)Cl] (2) and [(cpy)2Ir(dppe)Cl] (3)
ics (355 nm) of a Q-switched Nd:YAG laser (Quanta-Ray, Spectra were synthesized by following the above procedure using 2,2-
Physics) pulse with 8 ns was used to excite the iridium complexes. bipyridine (0.09 g, 0.58 mmol) and 1,2-bis (diphenylphosphino)
The emitted light was dispersed through a Czerny-Turner ethane (0.23 g, 0.58 mmol), respectively, instead of using
monochromator and detected using a Hamamatsu R928 photomul- biimidazole.
tiplier tube. The luminescence decay were captured with an Agi-
lent infiniium digital storage oscilloscope and the data were 4.4.2. [(cpy)2Ir(bpy)Cl] (2)
transferred to the computer for further analysis. Red colour solid. Yield: 49%. (0.14 g). m.p. 188 °C (dec). Anal.
Cyclic voltammetry (CV) was performed with a CHI604D elec- Calc. (%) for C44H30IrN6Cl: C, 60.70, H, 3.48, N, 9.65. Found (%): C,
trochemical analyzer for studying the electrochemical behavior 60.44, H, 3.62, N, 9.42. FT-IR (KBr, m, cm1 selected peaks: 3035 s,
of complexes using a three-electrode cell in which a glassy carbon 1601 s, 1173 s, 753 s (br, broad; s, sharp). UV–Vis in DCM [kmax/
electrode was the working electrode, a platinum wire counter elec- nm (emax/mol1 cm1)]: 227(708), 310(602), 349(299), 386(330),
V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275 273

451(195). 1H NMR (300 MHz, CDCl3) d (ppm): 7.223–7.324 (m, 4.6. BSA binding studies
9Hs), 7.748–7.841 (m, 8H), 8.308–8.406 (m, 7H), 8.687–8.606 (m,
6H). 13C NMR (75.46 MHz; CDCl3) d (ppm): 76.62, 77.25, 114.17, Protein binding studies have been performed by tryptophan flu-
115.59, 119.54, 121.38, 122.79, 123.99, 124.35, 126.77, 136.17, orescence quenching experiments using BSA (3 mM) in buffer (con-
138.59, 139.20, 140.62, 142.89, 149.10, 151.51, 153.81, 157.22. taining 15 mM trisodium citrate and 150 mM NaCl at pH 7.0).
ESI-MS (CH3OH) m/z (%): 870 [M+], 835 (100) [M+Cl]. Conductiv- Fluorescence emission spectra were recorded from 300 to
ity (KM/S cm2 mol1) in DMF: 45. 500 nm at an excitation wavelength of 295 nm.

4.7. Cytotoxic assay


4.4.3. [(cpy)2Ir(dppe)Cl] (3)
Green colour solid. Yield: 48%. (0.19 g). M.p. 178 °C (dec). Anal.
The cell line was cultured in the Dulbecco’s Modified Eagles
Calc. (%) for C60H46IrN4P2Cl: C, 64.76, H, 4.17, N, 5.03. Found (%): C,
Medium (DMEM) supplemented with 10% fetal bovine serum
64.50, H, 4.31, N, 4.80. FT-IR (KBr, m, cm1 selected peaks: 3046s,
(FBS), 200 mM L-glutamine, 100 m/mL penicillin, and 10 mg/mL
1623s, 1120s, 1476s, 1178s, 753s, 502s (br, broad; s, sharp). UV–
streptomycin in a humidified atmosphere consisting of 5 % CO2
Vis in DCM [kmax/nm (emax/mol1 cm1)]: 231(665), 309(252),
at 37 °C.
407(98), 487(28). 1H NMR (300 MHz, CDCl3) d (ppm): 1.60 (d,
4H), 5.91 (m, 1H), 6.23 (m, 1H), 6.45 (m, 2H), 6.85 (m, 4H), 7.15
4.8. Evaluation of cytotoxicity
(m, 1H), 7.25–7.21 (m, 6H), 7.49–7.40 (m, 15H), 7.58–7.51 (m,
8H), 7.69–7.62 (m, 4H). 13C NMR (75.46 MHz; CDCl3) d (ppm):
The cytotoxic effect of complex against human cervical cancer
31.80, 76.57, 77.20, 110.98, 114.33, 116.34, 119.51, 120.41,
(HeLa) cell line was evaluated by MTT [3-(4,5-dimethylthiazol-2-
123.94, 124.95, 125.23, 126.08, 128.76, 129.55, 130.68, 131.44,
yl)-2,5-diphenyl tetrazolium bromide] assay. Briefly, HeLa cells
132.27, 134.46, 138.28, 139.88, 152.40. ESI-MS (CH3OH) m/z (%);
were seeded (5  104 cells/well) in a 96-well plate and kept in
1112 (4) [M+], 1077 (100) [M+Cl]. Conductivity (KM/S cm2 mol1)
CO2 for attachment and growth for 24 h [70–72]. Then, the cells
in DMF: 30.
were treated with various concentrations of complex dissolved in
DMSO (0.25–100 mM) and incubated for 24 h. After incubation,
4.5. DNA binding studies the culture medium was removed and 15 lL of the MTT solution
(5 mg/mL in PBS) was added to each well. Following 4 h incubation
4.5.1. Absorption spectral studies in dark, MTT was discarded and DMSO (100 lL/well) was added to
Absorption spectra titrations were performed at room tempera- solubilize the purple formazan product. The experiment was car-
ture in Tris HCl/NaCl buffer (5 mM/50 mM buffer, pH 7.2) to inves- ried out in triplicates and the medium without complex served
tigate the binding affinity between CT-DNA and complex. A fixed as control. The absorbance was measured colorimetrically at
concentration of the complex (10 mM) was titrated with increasing 570 nm using an ELISA microplate reader. The percentage of cell
amounts of DNA concentration. The intrinsic binding constant for viability was calculated using the following formula and expressed
the interaction of complex with DNA was obtained from absorp- as IC50 = (OD value of treated cells)/(OD value of untreated cells
tion spectral data. (control)  100
The IC50 value is calculated using linear regression from excel.

4.5.2. Competitive binding experiments 4.9. DNA cleavage of complexes (1–3)


The relative binding of complex to CT-DNA were determined
with an EB-bound CT-DNA solution in Tris-HCl/NaCl buffer Cleavage of supercoiled pBR322 DNA, plasmid pBluescript II KS+
(5 mM/50 mM, pH = 7.2). The fluorescence on quenching effect DNA and pUC19 DNA by the complexes was studied by agarose gel
on addition of complex to the EB-DNA complex have been analyzed electrophoresis. In a typical experiment, the reaction mixtures
by recording the fluorescence emission spectra with excitation at (25 mL total volume) containing the DNA, dissolved in 50 mM
510 nm and emission at 596–600 nm. Tris–HCl buffer, 50 mM NaCl, were treated with the metal com-
plexes in the 100 mM concentration. The reaction mixtures were
incubated at 37 °C for 4 h, and then the addition of 5 mL loading
4.5.3. Circular dichroic spectral studies
buffer (0.25% bromophenol blue, 0.25% xylene cyanol, 30% glyc-
The CD spectra of CT-DNA in the presence or absence of com-
erol), samples were loaded on 0.8 % agarose gel containing EB
plexes were collected after 12 h incubation of 100 mM CT-DNA with
(1 mg/mL) in TAE (Tris-acetate-EDTA) buffer. The gel was run at
50 mM individual complex in Tris-HCl/NaCl buffer (5 mM/50 mM,
50 V for 1.30 h and photograph has been taken under UV light.
pH 7.2) at 37 °C. For each spectrum, three scans were collected
The proportion of DNA in each fraction was quantitatively esti-
and the background was subtracted from all of the reagents by
mated from the intensity of each band with the BioRad Gel Doc
using a corresponding solution without CT-DNA as a reference
XR system using the Labwork software. Anaerobic run was per-
solution. All CD experiments were performed on a Jasco J-810 spec-
formed in a stopcock-equipped cuvette after extensive purging of
tropolarimeter at room temperature from 320 to 220 nm.
the reaction mixture with N2 prior to the addition of N2-purged
complexes. To enhance the DNA cleaving ability by the complexes,
4.5.4. Viscosity measurements activators MPA was used. All the experiments were carried out in
The viscosity of solution relied on the time flowed through the triplicate under the same conditions.
capillary viscometer. The viscosity measurements were conducted
by keeping the viscometer immersed in a water-bath maintained Acknowledgements
at 25 °C ± 0.1 °C. Flow times were measured and the relative vis-
cosity was obtained by the equation: g = (t  t0)/t0, where t and We are thankful to the University Grants Commission, New
t0 represent the observed flow time of CT-DNA containing solu- Delhi (project No. 40-46/2011 (SR) and DST (DST-SR/FT/CS-
tions and buffer solution alone, respectively. The data was pre- 049/2009 for the financial support. The authors thank the SAIF
sented as (g/g0)1/3 versus the ratio of the concentration of Central Drug Research Institute (CDRI), Lucknow for analyzing
complex to that of CT-DNA ([Complex]/[CT-DNA]) [68,69]. the mass spectra.
274 V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275

Appendix A. Supplementary data hydrogen production from an ionic iridium(III) complex, Chem. Mater. 17
(2005) 5712–5719.
[25] J. Jiang, J. Liang, S. Zhao, Y. Wang, J. Qi, Y. Huang, T. Xiong, D. Zhu, Syntheses,
Supplementary data associated with this article can be found, in crystal structures and luminescent properties of drum and ladder-like
the online version, at http://dx.doi.org/10.1016/j.ica.2017.07.061. organooxotin clusters with carbazole ligand, J. Cluster Sci. 28 (2017) 971–982.
[26] B.L. Li, L. Wu, Y.M. He, Q.H. Fan, The synthesis and properties of Iridium(III)-
cored dendrimers with carbazole peripherally functionalized b-diketonato
References dendrons, Dalton Trans. (2007) 2048–2057.
[27] S. Reineke, K. Walzer, K. Leo, Triplet-exciton quenching in organic
phosphorescent light-emitting diodes with Ir-based emitters, Phys. Rev. B 75
[1] E. Holder, B.M.W. Langeveld, U.S. Schubert, New trends in the use of transition
(2007) 125328.
metal-ligand complexes for applications in electroluminescent devices, Adv.
[28] P. Leriche, P. Frere, A. Cravino, O. Aleveque, J. Roncali, Triplet-exciton
Mater. 17 (2005) 1109–1121.
quenching in organic phosphorescent light-emitting diodes with Ir-based
[2] H. Yersin, Triplet emitters for OLED applications. Mechanisms of exciton
emitters, J. Org. Chem. 72 (2007) 8332–8336.
trapping and control of emission properties, Top. Curr. Chem. 241 (2004) 1–26.
[29] C.H. Yang, C. Tai, I.W. Sun, Synthesis of a high-efficiency red phosphorescent
[3] H. Sun, S. Liu, W. Lin, K.Y. Zhang, W. Lv, X. Huang, F. Huo, H. Yang, G. Jenkins, Q.
emitter for organic light-emitting diodes, J. Mater. Chem. 14 (2004) 947–
Zhao, W. Huang, Smart responsive phosphorescent materials for data
950.
recording and security protection, Nat Commun. 5 (2014) 3601.
[30] F.I. Wu, H.J. Su, C.F. Shu, L. Luo, W.G. Diau, C.H. Cheng, J.P. Duan, G.H. Lee,
[4] W. Xu, X. Zhao, W. Lv, H. Yang, S. Liu, H. Liang, Z. Tu, H. Xu, W. Qiao, Q. Zhao, W.
Tuning the emission and morphology of cyclometalated iridium complexes
Huang, Rational design of phosphorescent chemodosimeter for reaction-based
and their applications to organic light-emitting diodes, J. Mater. Chem. 15
one- and two-photon and time-resolved luminescent imaging of biothiols in
(2005) 1035–1042.
living cells, Adv. Healthc. Mater. 3 (2014) 658–669.
[31] K.A. King, P.J. Spellane, R.J. Watts, Excited-state properties of a triply ortho-
[5] W. Zhang, C. Liang, Z. He, H. Pang, Y. Wang, S. Zhao, Stable orange and white
metalated iridium(III) complex, J. Am. Chem. Soc. 107 (1985) 1431–1432.
electro phosphorescence based on spirobifluorenyltrifluoromethylpyridine
[32] J. Kalinowski, G. Giro, M. Cocchi, V. Fattori, P.D. Macro, Unusual disparity in
iridium complexes, Synth. Metals 210 (2015) 214–222.
electroluminescence and photoluminescence spectra of vacuum-evaporated
[6] S. Tobita, T. Yoshihara, Intracellular and in vivo oxygen sensing using
films of 1, 1-bis ((di-4-tolylamino) phenyl) cyclohexane, Appl. Phys. Lett. 76
phosphorescent iridium(III) complexes, Cur. Opin. Chem. Biol. 33 (2016) 39–45.
(2000) 2352–2354.
[7] W. Su, X. Wang, X. Lei, Q. Xiao, S. Huang, P. Li, Synthesis, characterization,
[33] G. Calogero, G. Giuffrida, S. Serroni, V. Ricevuto, S. Campagna, Absorption
cytotoxic activity of half-sandwich rhodium(III), and iridium(III) complexes
spectra, luminescence properties, and electrochemical behavior of
with curcuminoids, J. Organomet. Chem. 833 (2017) 54–60.
cyclometalated iridium(III) and rhodium(III) complexes with a bis
[8] S. Liu, P. Wang, Q. Zhao, H. Yang, J. Wong, H. Sun, X. Dong, W. Lin, W. Huang,
(pyridyl)triazole ligand, Inorg. Chem. 34 (1995) 541–545.
Single polymer-based ternary electronic memory material and device, Adv.
[34] G. Ge, G. Zhang, H. Guo, Y.C. Huai, D. Zou, Yellow organic light-emitting diodes
Mater. 24 (2012) 2901–2905.
based on phosphorescent iridium(III) pyrazine complexes: Fine tuning of
[9] S. Liu, Z. Lin, Q. Zhao, Y. Ma, H. Shi, M. Yi, Q. Ling, Q. Fan, C. Zhu, E. Kang, W.
emission color, Inorg. Chim. Acta 362 (2009) 2231–2236.
Huang, Flash-memory effect for polyfluorenes with on-chain iridium(III)
[35] C. Dragonetti, L. Falciola, P. Mussini, S. Righetto, D. Roberto, R. Ugo, A. Valore, F.
complexes, Adv. Funct. Mater. 21 (2011) 979–985.
De Angelis, S. Fantacci, A. Sgamellotti, M. Ramon, M. Muccini, The role of
[10] W. Xu, S. Liu, H. Sun, X. Zhao, Q. Zhao, S. Sun, S. Cheng, T. Ma, L. Zhou, W. Huang,
substituents on functionalized 1,10-phenanthroline in controlling the
FRET-based probe for fluoride based on a phosphorescent iridium(III) complex
emission properties of cationic iridium(III) complexes of interest for
containing triarylboron groups, J. Mater. Chem. 21 (2011) 7572–7581.
electroluminescent devices, Inorg. Chem. 46 (2007) 8533–8547.
[11] H. Shi, S. Liu, H. Sun, W. Xu, Z. An, J. Chen, S. Sun, X. Lu, Q. Zhao, W. Huang,
[36] F. Dimiza, F. Perdih, V. Tangoulis, I. Turel, D.P. Kessissoglou, G. Psomas,
Simple conjugated polymers with on-chain phosphorescent iridium(III)
Interaction of copper(II) with the non-steroidal anti-inflammatory drugs
complexes: toward ratiometric chemodosimeters for detecting trace
naproxen and diclofenac: synthesis, structure, DNA- and albumin-binding, J.
amounts of mercury(II), Chem. Eur. J. 16 (2010) 12158–12167.
Inorg. Biochem. 105 (2011) 476–489.
[12] W. Xu, S. Liu, X. Zhao, S. Sun, S. Cheng, T. Ma, H. Sun, Q. Zhao, W. Huang,
[37] F. Dimiza, A.N. Papadopoulos, V. Tangoulis, V. Psycharis, C.P. Raptopoulou, D.P.
Cationic iridium(III) complex containing both triarylboron and carbazole
Kessissoglou, G. Psomas, Biological evaluation of non-steroidal anti-
moieties as a ratiometric fluoride probe that utilizes a switchable triplet-
inflammatory drugs-cobalt(II) complexes, Dalton Trans. 39 (2010) 4517–4528.
singlet emission, Chem. Eur. J. 16 (2010) 7125–7133.
[38] N. Shahabadi, S. Kashanian, M. Khosravi, M. Mahdavi, Multispectroscopic DNA
[13] K.K.-W. Lo, C.-K. Chung, N. Zhu, Synthesis, photophysical and electrochemical
interaction studies of a water-soluble nickel(II) complex containing different
properties, and biological labeling studies of cyclometalated iridium(III) bis
dinitrogen aromatic ligands, Trans. Metal Chem. 35 (2010) 699–705.
(pyridylbenzaldehyde) complexes: novel luminescent cross-linkers for
[39] F. Dimiza, S. Fountoulaki, A.N. Papadopoulos, C.A. Kontogiorgis, V. Tangoulis, C.
biomolecules, Chem. Eur. J. 9 (2003) 475–483.
P. Raptopoulou, V. Psycharis, A. Terzis, D.P. Kessissoglou, G. Psomas, Non-
[14] K.K.-W. Lo, C.-K. Chung, N. Zhu, Nucleic acid intercalators and avidin probes
steroidal antiinflammatory drug–copper(II) complexes: structure and
derived from luminescent cyclometalated iridium (III)–dipyridoquinoxaline
biological perspectives, Dalton Trans. 40 (2011) 8555–8568.
and–dipyridophenazine complexes, Chem. Eur. J. 12 (2006) 1500–1512.
[40] P. Sathyadevi, P. Krishnamoorthy, R.R. Butorac, A.H. Cowley, N.S.P. Bhuvaneshc,
[15] K.K.-W. Lo, W.-K. Hui, C.-K. Chung, K.H.-K. Tsang, D.C.-M. Ng, N. Zhu, K.-K.
N.A. Dharmaraj, Effect of substitution and planarity of the ligand on DNA/BSA
Cheung, Biological labelling reagents and probes derived from luminescent
interaction, free radical scavenging and cytotoxicity of diamagnetic Ni(II)
transition metal polypyridine complexes, Coord. Chem. Rev. 249 (2005) 1434–
complexes: a systematic investigation, Dalton Trans. 40 (2011) 9690–9702.
1450.
[41] M.D. Matiur Rahman, H.Y. Liu, K. Eriks, A. Prock, W.P. Giering, Quantitative
[16] K.K.-W. Lo, C.-K. Chung, T.K.-M. Lee, L.-H. Lui, K.H.-K. Tsang, N. Zhu, New
analysis of ligand effects. Part 3. Separation of phosphorus(III) ligands into
luminescent cyclometalated iridium(III) diimine complexes as biological
pure sigma-donors and sigma-donor/pi-acceptors. Comparison of basicity and
labeling reagents, Inorg. Chem. 42 (2003) 6886–6897.
sigma-donicity, Organometallics 8 (1989) 1–7.
[17] M.A. Baldo, D.F. O’Brien, M.E. Thompson, S.R. Forrest, Excitonic singlet-triplet
[42] A. Tarushi, G. Psomas, C.P. Raptopoulou, D.P. Kessissoglou, Zinc complexes of
ratio in a semiconducting organic thin film, Phys. Rev. B 60 (1999) 14422–14428.
the antibacterial drug oxolinic acid: structure and DNA-binding properties, J.
[18] X. Zhang, Z. Chen, C. Yang, Z. Li, K. Zhang, H. Yao, J. Qin, J. Chen, Y. Cao, Highly
Inorg. Biochem. 103 (2009) 898–905.
efficient polymer light-emitting diodes using color-tunable carbazole-based
[43] B.S. Yang, J.Y. Feng, Y.Q. Li, F. Gao, Y.Q. Zhao, J.L. Wang, Spectral studies on
iridium complexes, Chem. Phys. Lett. 422 (2006) 386–390.
aluminum ion binding to the ligands with phenolic group(s): implications for
[19] T.-H. Kwon, H.S. Cho, M.K. Kim, J.-W. Kim, J.-J. Kim, K.H. Lee, S.J. Park, I.-S. Shin,
the differences between N- and C-terminal binding sites of human serum
H. Kim, D.M. Shin, Y.K. Chung, J.-I. Hong, Color tuning of cyclometalated iridium
apotransferrin, J. Inorg. Biochem. 96 (2003) 416–424.
complexes through modification of phenylpyrazole derivatives and ancillary
[44] B.C. Baguley, M. Le Bret, Quenching of DNA-ethidium fluorescence by
ligand based on ab initio calculations, Organometallics 24 (2005) 1578–1585.
amsacrine and other antitumor agents: a possible electron-transfer effect,
[20] J. Sun, W. Wu, H. Guo, J. Zhao, Visible-light harvesting with cyclometalated
Biochemistry 23 (1984) 937–943.
iridium(III) complexes having long-lived 3IL excited states and their
[45] D.S. Raja, N.S.P. Bhuvanesh, K. Natarajan, Biological evaluation of a novel water
application in triplet–triplet-annihilation based upconversion, Eur. J. Inorg.
soluble sulphur bridged binuclear copper(II) thiosemicarbazone complex, Eur.
Chem. 2011 (2011) 3165–3173.
J. Med. Chem. 46 (2011) 4584–4594.
[21] Y. You, S.Y. Park, Inter-ligand energy transfer and related emission change in
[46] H.M. Torshizi, M. Saeidifar, F. Khosravi, A. Divsalar, A.A. Saboury, F. Hassani,
the cyclometalated heteroleptic iridium complex: facile and efficient colour
DNA binding and antitumor activity of a-diimineplatinum(II) and Palladium
tuning over the whole visible range by the ancillary ligand structure, J. Am.
(II) dithiocarbamate complexes, Bioinorg. Chem. Appl. (2011) 1–11.
Chem. Soc. 127 (2005) 12438–12439.
[47] Z.C. Liu, B.D. Wang, Z.Y. Yang, Y. Li, D.D. Qin, T.R. Li, Synthesis, crystal structure,
[22] L. Ma, H. Guo, Q. Li, S. Guo, J. Zhao, Visible light-harvesting cyclometalated Ir
DNA interaction and antioxidant activities of two novel water-soluble Cu2+
(III) complexes as triplet photosensitizers for triplet–triplet annihilation based
complexes derivated from 2-oxo-quinoline-3-carbaldehyde Schiff-bases, Eur.
upconversion, Dalton Trans. 41 (2012) 10680–10689.
J. Med. Chem. 44 (2009) 4477–4484.
[23] Y. You, K.S. Kim, T.K. Ahn, D. Kim, S.Y. Park, Direct spectroscopic observation of
[48] G.M. Howe, K.C. Wu, W.R. Bauer, S.J. Lippard, Binding of platinum and
interligand energy transfer in cyclometalated heteroleptic iridium(III)
palladium metallointercalation reagents and antitumor drugs to closed and
complexes: a strategy for phosphorescence color tuning and white light
open DNAs, Biochemistry 19 (1976) 4339–4347.
generation, J. Phys. Chem. C 111 (2007) 4052–4060.
[49] M. Baldini, M. Belicchi-Ferrari, F. Bisceglie, P.P. Dall’Aglio, G. Pelosi, S. Pinelli,
[24] M.S. Lowry, J.I. Goldsmith, J.D. Slinker, R. Rohl, R.A. Pascal, G.G. Malliaras, S.
P. Tarasconi, Copper(II) complexes with substituted thiosemicarbazones of
Bernhard, Single-layer electroluminescent devices and photoinduced
V. Thamilarasan et al. / Inorganica Chimica Acta 467 (2017) 264–275 275

a-ketoglutaric acid: synthesis, X-ray structures, DNA binding studies and [61] M.T. Carter, A.J. Bard, Voltammetric studies of the interaction of tris(1,10-
nuclease and biological activity, Inorg. Chem. 43 (2004) 7170–7179. phenanthroline)cobalt(III) with DNA, J. Am. Chem. Soc. 109 (1987) 7528–
[50] P. Lincoln, E. Tuite, B. Norden, Short-circuiting the molecular wire: cooperative 7530.
binding of D-[Ru(phen)2dppz]2+ and D-[Rh(phi)2bipy]3+ to DNA, J. Am. Chem. [62] D.H. Johnston, K.C. Glasgow, H.H. Thorp, Electrochemical measurement of the
Soc. 119 (1997) 1454–1455. solvent accessibility of nucleobases using electron transfer between DNA and
[51] V. Uma, M. Kanthimathi, T. Weyhermuller, B. Unni Nair, Oxidative DNA metal complexes, J. Am. Chem. Soc. 117 (1995) 8933–8938.
cleavage mediated by a new copper (II) terpyridine complex: crystal structure [63] N. Malecevic, T. Srdic, S. Radulovic, D. Sladic, V. Radulovic, I. Brceski, K.
and DNA binding studies, J. Inorg. Biochem. 99 (2005) 2299–2307. Andelkoivc, Synthesis and characterization of a novel Pd(II) complex with the
[52] P. Uma Maheswari, M. Palaniandavar, DNA binding and cleavage properties of condensation product of 2-(diphenylphosphino)benzaldehyde and ethyl
certain tetrammine ruthenium(II) complexes of modified 1,10- hydrazinoacetate. Cytotoxic activity of the synthesized complex and related
phenanthrolines – effect of hydrogen-bonding on DNA-binding affinity, J. Pd(II) and Pt(II) complexes, J. Inorg. Biochem. 100 (2006) 1811–1818.
Inorg. Biochem. 98 (2004) 219–230. [64] A.M. Javier, O. Kreft, A.P. Alberola, C. Kirchner, B. Zebli, A.S. Susha, E. Horn, S.
[53] Q.L. Zhang, J.G. Liu, H. Chao, G.Q. Xue, L.N. Ji, NA-binding and photocleavage Kempter, A.G. Skirtach, A.L. Rogach, J. Radler, G.B. Sukhorukov, M. Benoit, W.J.
studies of cobalt(III) polypyridyl complexes: [Co(phen)2IP]3+ and [Co Parak, Combined atomic force microscopy and optical microscopy
(phen)2PIP]3+, J. Inorg. Biochem. 83 (2001) 49–55. measurements as a method to investigate particle uptake by cells, Small 2
[54] P. Xi, Z. Xu, F. Chen, Z. Zeng, X. Zhang, Study on synthesis, structure, and DNA- (2006) 394–400.
binding of Ni, Zn complexes with 2-phenylquinoline-4-carboylhydrazide, J. [65] R. Luo, B. Neu, S.S. Venkatraman, Surface functionalization of nanoparticles to
Inorg. Biochem. 103 (2009) 210–218. control cell interactions and drug release, Small 8 (2012) 2585–2594.
[55] S. Tabassum, M. Zaki, M. Afzal, F. Arjmand, New modulated design and [66] D. Mustard, D.W. Ritchie, Docking essential dynamics eigen structures,
synthesis of quercetin–CuII/ZnII–SnIV 2 scaffold as anticancer agents: in vitro Proteins Struct. Funct. Bioinf. 60 (2005) 269–274.
DNA binding profile, DNA cleavage pathway and Topo-I activity, Dalton Trans. [67] V. Thamilarasan, A. Jayamani, P. Manisankar, Young-Inn Kim, N.
42 (2013) 10029–10041. Sengottuvelan, Green-emitting phosphorescent iridium(III) complex:
[56] C. Tan, J. Liu, H. Li, W. Zheng, S. Shi, L. Chen, L. Ji, Differences in structure, structural, photophysical and electrochemical properties, Inorg. Chim. Acta
physiological stability, electrochemistry, cytotoxicity, DNA and protein 408 (2013) 240–245.
binding properties between two Ru(III) complexes, J. Inorg. Biochem. 102 [68] Y. Li, G. Zhang, M. Tao, Binding properties of herbicide chlorpropham to DNA:
(2008) 347–358. spectroscopic, chemometrics and modeling investigations, J. Photochem.
[57] Y. Wang, H. Zhang, G. Zhang, W. Tao, S. Tang, Interaction of the flavonoid Photobiol. B 138 (2014) 109–117.
hesperidin with bovine serum albumin: a fluorescence quenching study, J. [69] X. Li, X.J. Li, Y.T. Li, Z.Y. Wu, C.W. Yan, Syntheses and structures of tetracopper
Luminescence 126 (2007) 211–218. (II) complexes with an N-benzoate-N0 -[3-(2-hydroxylethylammino)propyl]
[58] D.S. Raja, N.S.P. Bhuvanesh, K. Natarajan, A novel water soluble ligand bridged oxamide ligand: reactivity towards DNA, cytotoxic and antimicrobial
cobalt(II) coordination polymer of 2-oxo-1,2-dihydroquinoline-3- activities, New J. Chem. 36 (2012) 2472–2483.
carbaldehyde (isonicotinic) hydrazone: evaluation of the DNA binding, [70] S.J. Mehdi, A. Ahmad, M. Irshad, N. Manzoor, M.M.A. Rizvi, Cytotoxic effect
protein interaction, radical scavenging and anticancer activity, Dalton Trans. of carvacrol on human cervical cancer cells, Biol. Med. 3 (2) (2011) 307–
41 (2012) 4365–4377. 312.
[59] V. Thamilarasan, N. Sengottuvelan, A. Sudha, P. Srinivasan, A. Siva, Synthesis, [71] Mustofa. Masrianif, Sunarti. Jumina, Pycnarrhena cauliflora ethanolic extract
molecular structure, theoretical calculation, DNA/protein interaction and induces apoptosis and cell cycle arrest in hela human cervical cancer cells, Int.
cytotoxic activity of manganese(III) complex with 8-hydroxyquinoline, J. J. Res. Pharm. Biomed. Sci. 4 (2013) 1060–1068.
Photochem. Photobiol. B 142 (2015) 220–231. [72] C. Zhang, X. Jia, K. Wang, J. Bao, P. Li, M. Chen, J.B. Wan, H. Su, Z. Mei, C. He,
[60] A. Tarushi, E. Polatoglou, J. Kljun, I. Turel, G. Psomas, D.P. Kessissoglou, Polyphyllin VII induces an autophagic cell death by activation of the JNK
Interaction of Zn(II) with quinolone drugs: Structure and biological evaluation, pathway and inhibition of PI3K/AKT/mTOR pathway in HepG2 cells, PLoS One
Dalton Trans. 40 (2011) 9461–9473. (2016) 1–5.

You might also like