Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Dalton

Transactions
PAPER

Nickel(II), copper(II) and zinc(II) metallo-


Cite this: Dalton Trans., 2014, 43,
intercalators: structural details of the DNA-binding
6108 by a combined experimental and computational
investigation†
Antonino Lauria,a Riccardo Bonsignore,a Alessio Terenzi,a Angelo Spinello,a
Francesco Giannici,b Alessandro Longo,c,d Anna Maria Almericoa and
Giampaolo Barone*a,e

We present a thorough characterization of the interaction of novel nickel(II) (1), copper(II) (2) and zinc(II) (3)
Schiff base complexes with native calf thymus DNA (ct-DNA), in buffered aqueous solution at pH 7.5.
UV-vis absorption, circular dichroism (CD) and viscometry titrations provided clear evidence of the inter-
calative mechanism of the three square-planar metal complexes, allowing us to determine the intrinsic
DNA-binding constants (Kb), equal to 1.3 × 107, 2.9 × 106, and 6.2 × 105 M−1 for 1, 2 and 3, respectively.
Preferential affinity, of one order of magnitude, toward AT compared to GC base pair sequences was
detected by UV-vis absorption titrations of 1 with [ poly(dG-dC)]2 and [ poly(dA-dT)]2. Structural details of
the intercalation site of the three metal complexes within [dodeca(dA-dT)]2 were obtained by molecular
dynamics (MD) simulations followed by density functional theory/molecular mechanics (DFT/MM) calcu-
lations. The calculations revealed that three major intermolecular interactions contribute to the strong
affinity between DNA and the three metal complexes: (1) the electrostatic attraction between the two
positively charged triethylammoniummethyl groups of the metal complexes and the negatively charged
phosphate groups of the DNA backbone; (2) the intercalation of the naphthalene moiety within the four
nitrogen bases of the intercalation site; (3) the metal coordination by exocyclic donor atoms of the bases,
specifically the carbonyl oxygen and amine nitrogen atoms. Remarkably, the Gibbs formation free energy
calculated for the intercalation complexes of 1, 2 and 3 with [dodeca(dA-dT)]2 in the implicit water solu-
tion is in agreement with the experimental Gibbs free energy values obtained from the DNA-binding con-
stants as ΔG° = −RT ln(Kb). In particular, the DNA-binding affinity trend, 1 > 2 > 3, is reproduced. Finally,
Received 30th October 2013, the first shell coordination distances calculated for the intercalation complex 3/[dodeca(dA-dT)]2 are in
Accepted 20th January 2014
excellent agreement with the experimental distances extracted from the extended X-ray absorption fine
DOI: 10.1039/c3dt53066c structure (EXAFS) spectrum of the corresponding 3/ct-DNA solutions. The latter results provided the first
www.rsc.org/dalton evidence of metal ion coordination by native DNA in aqueous solution.

a
Dipartimento di Scienze e Tecnologie Biologiche, Chimiche e Farmaceutiche,
Viale delle Scienze, Edificio 17, I-90128 Palermo, Italy.
E-mail: giampaolo.barone@unipa.it; Fax: +39 091 596825
b
Introduction
Dipartimento di Fisica e Chimica, Università di Palermo, Viale delle Scienze,
Edificio 17, I-90128 Palermo, Italy The interaction of metal complexes with DNA is associated
c
Istituto per lo Studio su Materiali Nanostrutturati del Consiglio Nazionale delle
with interesting chemical and biological properties of the
Ricerche (ISMN-CNR), Via Ugo La Malfa n. 153, I-90146 Palermo, Italy
d
Dubble CRG at ESRF, Netherlands Organization for Scientific Research (NWO),
resulting supramolecular system. In fact, the cytotoxic activity
c/o ESRF, BP 220, F-38043 Grenoble Cedex, France of metallodrugs has often been correlated to their DNA-
e
Istituto EuroMediterraneo di Scienza e Tecnologia, Via Emerico Amari 123, binding properties.1–3 For this reason, we have been recently
90139 Palermo, Italy involved in the synthesis of novel metal complexes and organic
† Electronic supplementary information (ESI) available: Additional figures
compounds as potential DNA binders.4–10
(Fig. S1–S5) and tables (Tables S1–S3), reporting NMR spectra and DFT energies,
in vacuo and in solution, thermal corrections, Cartesian coordinates. See DOI: The interaction modes of a small molecule with the DNA
10.1039/c3dt53066c macromolecule have been broadly categorized into the following

6108 | Dalton Trans., 2014, 43, 6108–6119 This journal is © The Royal Society of Chemistry 2014
Dalton Transactions Paper

three types:4,11,12 (1) covalent binding DNA-alkylators,13,14 or Experimental section


with Lewis acid metal ions, such as platinum in many clinically
used anticancer drugs;15–17 (2) major or minor grooves- Materials and methods
binding;18,19 and (3) intercalation.20,21 The latter two categories Solvents and reagents (reagent grade) were all commercially
are classified as non-covalent interactions. Grooves binding available and used without further purification. 1H NMR and
13
involves the contributions of electrostatic interaction of polar or C NMR spectra were recorded in DMSO-d6 solution, with TMS
cationic molecules with the negatively charged phosphate as an internal reference, on a Bruker Avance 200 MHz NMR
groups, and/or the formation of hydrogen bonds with the spectrometer at 200 and 50 MHz, respectively. IR spectra were
oxygen and/or nitrogen atoms of the DNA bases or of the sugar recorded on a Perkin-Elmer Spectrum 100 FT-IR spectrometer.
fragment, as both donor and acceptor atoms. Intercalation Lyophilized calf thymus DNA (Fluka, BioChemika) was
involves specifically the π–π interaction of the planar aromatic resuspended in 1.0 mM tris-hydroxymethyl-aminomethane
section of the small molecules with the stacked aromatic planes (Tris-HCl) at pH = 7.5 and dialyzed as described in the litera-
of the nitrogen bases. ture.47 DNA concentration, expressed in monomer units
Mixed modes of interactions have been also advised. For ([DNAphosphate]), was determined by UV spectrophotometry
example, many intercalators, such as anthracyclines, also using 7000 M−1 cm−1 as the molar absorption coefficient at
interact by electrostatic attraction or by hydrogen bonding with 258 nm.48 Lyophilized synthetic polynucleotides [ poly(dG-
the atoms exposed in the grooves.22–27 Moreover, it has been dC)]2 and [poly(dA-dT)]2 (Sigma) were dissolved in 1.0 mM
recently shown that transition metal complexes of planar aro- Tris-HCl, at pH = 7.5, and their molar concentration, in
matic ligands may combine coordination with intercalation in monomer units, was determined using 8400 M−1 cm−1 and
their DNA-binding.20,28 Such a feature confers them enhanced 6650 M−1 cm−1 as the molar absorption coefficient at 254 nm
anticancer properties. and 262 nm, respectively.49,50
Of course, fine structural details of the molecule–DNA UV-vis spectra were collected on a Varian UV-vis Cary 1E
binding mode in solution cannot be completely obtained by double beam spectrophotometer equipped with a Peltier temp-
techniques such as UV-vis absorption, CD and viscometry. erature controller. Circular dichroism spectra were recorded on
Nonetheless, the above-mentioned techniques are interesting a Jasco J-715 spectropolarimeter. 1 cm path-length quartz cuv-
because they can be routinely applied to solution samples ettes were used. Viscosity measurements were performed on
under physiological conditions. an Ubbelodhe viscometer maintained at 25.0 ± 0.1 °C. The
Computational chemistry may support the interpretation flow time was measured with a digital stopwatch. UV-visible
of experimental data with atomistic models. In this absorption titrations were carried out adding increasing
context, quantum mechanics/molecular mechanics (QM/MM) amounts of ct-DNA to a metal-complex solution with constant
calculations provide reliable structural information on a concentration. DNA melting experiments were carried out by
local region of biomolecular systems, such as the active site recording the absorbance of ct-DNA solutions at 258 nm while
of metalloproteins.29–31 We have recently shown that the temperature was elevated gradually from 20 to 95 °C at a
molecular dynamics (MD) simulations, followed by QM/MM rate of 0.5 °C min−1. CD, melting temperature and viscometry
calculations, with the QM part described by density titrations were carried out adding increasing amounts of metal
functional theory (DFT), provided a detailed structure of the complex stock solution to a DNA solution with constant con-
intercalation complexes of dodecanucleotide double-helices centration. All experiments were carried out in Tris-HCl 1 mM
with a copper(II) metallointercalator, and formation energy aqueous buffer at pH = 7.5.
data in good agreement with the experimental DNA-binding X-Ray absorption spectroscopy (XAS) experiments were
constant.32 carried out at the BM26A beamline of European Synchrotron
On the other hand, detailed local structural information Radiation Facility (ESRF), Grenoble. The spectra were recorded
can be experimentally obtained by X-ray crystallography or by in fluorescence mode with a 9-element Ge detector on Ni (8333
NMR spectroscopy only of drug–DNA supramolecular com- eV), Cu (8979 eV) and Zn (9659 eV) respectively, on buffered
plexes involving small synthetic DNA oligonucleotides, typi- aqueous solutions of [1] = 503 μM, [2] = 95 μM and [3] = 457 μM,
cally in the range 2–12-mer oligonucleotides.33–41 Luckily, the both with and without DNA at a molar concentration of 4 : 1
analysis of the X-ray absorption spectrum allows detecting DNA–metal complex. The sample solutions were quenched in
changes in the coordination structure of metal complexes, in liquid nitrogen and the spectra recorded at 80 K using a cryo-
particular the formation of new covalent bonds, when a metal stat, averaging 10 spectra per sample for about 6 hours. XAS
complex interacts with biomolecules in aqueous solution.42 spectra were averaged and normalized with Athena.51 EXAFS
For such reasons, in the present paper we have exploited also spectra were fitted with Viper.52 FEFF8.4 was used to calculate
this technique. Even though the extended X-ray absorption theoretical amplitudes and phases for the EXAFS analysis, and
fine structure (EXAFS) has been widely used to characterize the full multiple scattering simulations on the XANES.53
active site of metalloproteins,43–46 to the best of our knowl-
edge, this is the first time this spectroscopic technique is used Synthesis and characterization
to detect the bonds between native DNA and a metal complex Complexes 1–3 (Scheme 1) were synthesized by the reaction
in water solution. of 5-(triethylammoniummethyl)salicylaldehyde chloride, 2,3-

This journal is © The Royal Society of Chemistry 2014 Dalton Trans., 2014, 43, 6108–6119 | 6109
Paper Dalton Transactions

Elemental analysis for C38Cl2H52N4O12Cu (2, 2ClO4−,


2H2O). Found: C, 51.72%, H, 5.46%, N, 6.55%; calc.: C,
51.21%, H, 5.88%, N, 6.29%). IR: ν (CvN) 1625 cm−1; ν (N–Cu)
503 cm−1; ν (O–Cu) 469 cm−1. UV [nm (ε, M−1 cm−1)]: 247 (5.14
× 104), 316 (2.26 × 104), 406 (1.91 × 103).
3: 5-(triethylammoniummethyl)salicylaldehyde chloride
(135.4 mg, 0.50 mmol) was dissolved in H2O–EtOH with
two equivalents of NaOH (20.1 mg, 0.50 mmol). This solution
was added to an ethanol solution of 2,3-naphthalendiamine
Scheme 1 Reaction pathway leading to the nickel(II), copper(II) and zinc(II) (39.2 mg, 0.25 mmol) and Zn(ClO4)2·6H2O (94.0 mg,
complexes, 1–3, of N,N’-bis-5-(triethyl ammonium methyl)-salicy- 0.25 mmol). The resulting mixture was stirred for 4 h at room
lidene-2,3-naphthalendiiminato).
temperature. The brilliant yellow precipitate was washed with
cold ethanol and diethyl ether, and recrystallized from 50%
diaminonaphthalene and metal(II) perchlorate in a methanol–water solutions to afford the compound as a yellow
2 : 1.1 : 1 molar ratio. Salicylaldehyde in EtOH–H2O basic solu- solid (yield 169.0 mg, 79%).
1
tion was added dropwise to the diamine and the metal H NMR δ ( ppm): 1.18–1.44 (m, 18H, CH3); 3.13–3.15
perchlorate in ethanol. The 5-(triethylammoniummethyl)sali- (m, 12H, CH2); 4.32 (s, 4H, CH2); 6.70 (d, 2H, J = 12.6 Hz, Ar);
cylaldehyde chloride ligand was prepared from 5-chloromethyl 7.27 (d, 2H, J = 12.9 Hz, Ar); 7.45–7.60 (m, 4H, Ar); 7.89–8.02
salicylaldehyde and triethylamine in tetrahydrofuran, as (m, 2H, Ar); 8.22–8.34 (m, 2H, Ar); 9.07 (s, 2H, CH).
reported.54 The products were characterized by NMR, IR, 13
C NMR δ ( ppm): 7.5 (CH3); 51.1 (CH2); 59.5 (CH2); 110.8
elemental analysis, and UV-vis spectroscopy. (Ar); 114.1 (Ar); 119.5 (Ar); 123.8 (Ar); 126.4 (Ar); 127.7 (Ar);
1: 5-(triethylammoniummethyl)salicylaldehyde chloride 132.1 (Ar); 137.4 (Ar); 139.2 (Ar); 140.9 (Ar); 163.4 (Ar); 173.2
(135.5 mg, 0.50 mmol) was dissolved in H2O–EtOH with two (CH).
equivalents of NaOH (20.0 mg, 0.50 mmol). This solution was Elemental analysis for C38Cl2H48N4O10Zn (3, 2ClO4−):
added to an ethanol solution of 2,3 naphthalendiamine Found: C, 53.01%, H, 5.76%, N, 5.90%; calc.: C, 53.25%, H,
(39.5 mg, 0.25 mmol) and Ni(ClO4)2·6H2O (92.3 mg, 5.64%, N, 6.54%. IR: ν (CvN) 1626 cm−1; ν (N–Zn) 552 cm−1;
0.25 mmol). The resulting mixture was stirred for 4 h at room ν (O–Zn) 468 cm−1. UV [nm (ε, M−1 cm−1)]: 244 (6.27 × 104),
temperature. The brilliant orange precipitate was washed with 304 (2.80 × 104), 384 (2.81 × 103).
cold ethanol and diethyl ether, and recrystallized from 50%
methanol–water solutions to afford the compound as an Computational details
orange solid (yield 106.0 mg, 48%). Molecular dynamics simulations. An alternating adenine–
1
H NMR δ ( ppm): 1.32–1.38 (m, 18H, CH3); 3.17 (d, 12H, J = thymine sequence of 12 base pairs, [dodeca(dA-dT)]2, was con-
9.6 Hz, CH2); 4.37 (s, 4H, CH2); 6.94 (d, 2H, J = 13.2 Hz, Ar); structed in the double-helix B-DNA conformation by the
7.43 (d, 2H, J = 12.9 Hz, Ar); 7.54–7.58 (m, 2H, Ar); 7.71 (s, 2H, NUCLEIC routine of the TINKER program package,55
Ar); 7.88–7.93 (m, 2H, Ar); 8.61 (s, 2H, Ar); 10.16–10.68 (s, 2H, as recently described.32 To create the intercalation pocket in
CH). the starting DNA structure, the torsion angles α–ζ and χ of the
13
C NMR δ ( ppm): 7.4 (CH3); 51.4 (CH2); 58.8 (CH2); 99.5 sugar phosphate backbone were opportunely modified.56 The
(Ar); 114.1 (Ar); 114.7 (Ar); 121.0 (Ar); 127.3 (Ar); 128.0 (Ar); naphthalene moiety of the zinc(II) complex 3 was intercalated
132.1 (Ar); 138.7 (Ar); 138.9 (Ar); 139.8 (Ar); 157.5 (Ar); 166.2 from the major groove side between the 6th and 7th base pairs
(CH). using a home-made routine. MD simulations on the 3/[dodeca-
Elemental analysis for C38Cl2H52N4O12Ni (1, 2ClO4−, 2H2O). (dA-dT)]2 complex were performed by the GROMACS 4.5.3
Found: C, 50.05%, H, 5.61%, N 5.90%; calc.: C, 51.49%, H, software package57,58 using the Amber99 force field59 with
5.91%, N, 6.32%. Parmbsc0 nucleic acid torsions.60,61 The starting geometry and
IR: ν (CvN) 1619 cm−1; ν (O–Ni) 506 cm−1; ν (N–Ni) the partial atomic charges of the metal complex were obtained
475 cm−1. UV [nm (ε, M−1 cm−1)]: 256 (5.93 × 104), 346 (2.55 × by DFT calculations (see below), while other intramolecular
104), 467 (7.67 × 103). force-field parameters were generated using the ACPYPE
2: 5-(triethylammoniummethyl)salicylaldehyde chloride software.62–64
(140.0 mg, 0.52 mmol) was dissolved in H2O–EtOH with two A triclinic box of TIP3P water molecules was added around
equivalents of NaOH (20.0 mg, 0.50 mmol). This solution was the oligonucleotides to a depth of 1.5 nm on each side of the
added to an ethanol solution of 2,3 naphthalendiamine solutes, for a total of 20 187 atoms, to obtain a solution
(39.8 mg, 0.25 mmol) and Cu(ClO4)2·6H2O (92.6 mg, density of about 1.02 g ml−1. 20 Na+ counterions were added
0.25 mmol). The resulting mixture was stirred for 4 h at room to neutralize the negative charges of the phosphate groups,
temperature. The brownish-green precipitate was washed with while other 18 Na+ and Cl− ions were added to set the solution
cold ethanol and diethyl ether, and recrystallized from 50% ionic strength to about 0.15 M (see ESI, Fig. S1†). van der
methanol–water solutions to afford the compound as a green Waals parameters for zinc (σ = 0.195998 nm, ε = 0.05230 kJ
solid (yield 165.3 mg, 74%). mol−1), sodium (σ = 0.332840 nm, ε = 0.0115897 kJ mol−1) and

6110 | Dalton Trans., 2014, 43, 6108–6119 This journal is © The Royal Society of Chemistry 2014
Dalton Transactions Paper

chlorine (σ = 0.440104 nm, ε = 0.418400 kJ mol−1) were taken (Tables S2–S4†). All calculations were performed using the
from the Amber99 force field implemented in GROMACS. Gaussian 09 program package.74
Explicit solvent simulations were performed in the isother-
mal–isobaric NPT ensemble, at a temperature of 300 K, under
the control of a velocity rescaling thermostat.65,66 The particle Results and discussion
mesh Ewald method67 was used to describe long-range electro-
Synthesis and characterization
static interactions. The timestep for integration was 2 fs and
all covalent bonds, including the four bonds between the 1–3 were synthesized as reported in Scheme 1 and character-
metal ion and the tetracoordinate Schiff base ligand, con- ized by NMR, IR and elemental analysis.
strained with the LINCS algorithm. There were two tempera- While the NMR spectra of the paramagnetic copper
ture coupling groups in these simulations, the first for the complex 2 were broadened by the magnetic interaction with
dodecanucleotide and, if present, for the metal complex, and the copper ion, the square planar N2O2 coordination around
the second for water and ions. Preliminary MD simulations the metal atom is confirmed by the NMR spectra obtained for
showed that the structure of the isolated metal complex is 1 on the basis of the diamagnetic nature of the compound,
maintained in solution. Preliminary energy minimizations typical of planar nickel compounds with a d8 electronic con-
were run for 5000 steps with the steepest descent algorithm. figuration. According to the literature, the diamagnetism of
During the equilibration, the oligonucleotide and the metal NiII-Salen complexes is also maintained in solutions of coordi-
complex/oligonucleotide system were harmonically restrained nating and non-coordinating solvents.75 It should be noted
with a force constant of 1000 kJ mol−1 nm−2, gradually relaxed that elemental analysis suggests that the nickel(II) and
into five consecutive steps of 100 ps each, to 500, 200, 100 copper(II) complexes, 1 and 2, present two crystallization water
and 50 kJ mol−1 nm−2. The MD simulations were carried out molecules.
for 50 ns.
DFT/MM calculations. The relaxed geometries of the com- Experiments in aqueous solution
plexes between [dodeca(dA-dT)]2 and the metal complexes 1–3 The metal complexes under investigation share an intense
were optimized by two-layer quantum mechanics/molecular absorption band at about 250 nm. A characteristic absorption
mechanics (QM/MM) hybrid calculations, as implemented in band, related to metal perturbed infra-ligand electronic tran-
the ONIOM method,68,69 with the aim to perform a high-level sition, is noticeable in 1 (346 nm), 2 (316 nm) and 3 (304 nm).
calculation on the intercalation pocket and to take into Finally, a weak transition band can be observed at 467 nm for
account the constraining effects of the double-helical structure 1, which is remarkably stronger at 406 nm for 2 and at 384 nm
at lower levels of theory. The M06-2X70 DFT functional and the for 3 (black lines in Fig. 1a–c). Such spectra are significantly
dzvp basis set71 were used in the higher QM layer, to suitably modified by the addition of increasing amounts of calf thymus
model the hydrogen bonding and π–π stacking interactions DNA. In detail, the absorption bands of 1 at 346 nm, of 2 at
between the sixth and seventh Watson–Crick base pairs. As 316 nm and of 3 at 304 nm are red shifted by 5, 8 and 11 nm
recently described,32 the Amber99 force field was used in the and show hypochromism of about 42%, 35% and 26%,
lower MM layer of the DFT/MM calculations. The highest respectively (Fig. 1a–c). These results, attributable to stacking
layer of the model includes the sixth and seventh base interaction between the aromatic naphthalene rings of the
pairs and the cationic complexes 1–3, with charge set to +2 complexes and the base pairs of DNA, collectively suggest that
and spin multiplicity 1 or 2, the latter in the presence of the the three metal complex cations act as DNA intercalators.4,76,77
copper ion. Default atomic partial charges were used for the To determine the intrinsic binding constant, Kb, and the
dodecanucleotide atoms, implicitly included in the force field stoichiometry, s, of the metal complex–DNA systems, the quan-
parameters. tity (εa − εf )/(εb − εf ) at 346 nm for 1, at 316 nm for 2 and at
Vibration frequency calculations, within the harmonic 304 nm for 3, has been plotted as a function of the molar con-
approximation, were performed on the optimized geometries centration of DNA (inset in Fig. 1a–c), and analyzed by eqn
by the same DFT/MM method. Solvent effects were evaluated (1a) and (1b):78
by performing M06-2X/dzvp single point calculations on the 1=2
2K b 2 Ct DNAphosphate

high layer model extracted by the DFT/MM optimized geome-

b b2
try, with the implicit water solvent reproduced by the polariz- εa εf s
¼ ð1aÞ
able continuum model (PCM),72,73 using default settings for εb εf 2K b Ct
PCM cavities. Standard enthalpy and Gibbs free energy values,
K b DNAphosphate
 
at 298.15 K, of each energy minimum structure, both in vacuo b ¼ 1 þ K b Ct þ ð1bÞ
and in solution, were calculated by adding the thermal correc- 2s
tion obtained by vibration frequency analysis of the DFT/MM where Ct is the total concentration of the metal complex, εf,
systems to the DFT energy calculated for the high layers both determined by a calibration curve of the isolated metal com-
in vacuo and in water solution (see ESI, Table S1†). The PCM plexes in aqueous solution. εb was determined from the
energy data contain also non-electrostatic effects. Cartesian plateau of the plot, where further addition of DNA did not
coordinates of the optimized structures are reported in the ESI cause any changes in the absorption spectrum. Finally, εa was

This journal is © The Royal Society of Chemistry 2014 Dalton Trans., 2014, 43, 6108–6119 | 6111
Paper Dalton Transactions

Table 1 DNA-binding constant (Kb) and stoichiometry (s) of the three


metal complexes

1 2 3

Kb (1.03 ± 0.05) × 107 (2.9 ± 0.2) × 106 (6.2 ± 0.3) × 105


s 0.74 ± 0.06 0.87 ± 0.07 1.01 ± 0.07

Fig. 2 Absorption spectra of 1 in the presence of increasing amounts of


[ poly(dA-dT)]2 (a) and of [ poly(dG-dC)]2 (b) in Tris-HCl 1 mM buffer. [1] =
Fig. 1 Absorption spectra of 1 (a), 2 (b) and 3 (c) in the presence of 9 μM, [DNAphosphate] = 0–80 μM. The arrows indicate the trend of the
increasing amounts of ct-DNA in Tris-HCl 1 mM buffer. (a) [1] = 8 μM, spectral change upon polynucleotide addition.
[DNAphosphate] = 0–100 μM. (b) [2] = 13 μM, [DNAphosphate] = 0–60 μM. (c)
[3] = 35 μM, [DNAphosphate] = 0–193 μM. The arrows indicate the trend of
the spectral change upon DNA addition.

Analogous spectrophotometric titrations of the nickel(II)


complex 1 were carried out with synthetic polynucleotides of
alternating AT and GC sequences, as shown in Fig. 2, with the
determined as the ratio between the measured absorbance aim of determining its preferential affinity toward specific
and the analytical molar concentration of 1–3. base pair sequences.
The values of Kb and s are reported in Table 1. These The values of Kb and s obtained, reported in Table 2, show
results confirm that each metal complex tightly binds DNA that the binding affinity toward AT sequences is the same as
and that the binding strength decreases in the order 1 > 2 > 3, that toward ct-DNA, within the associated error and is one
previously observed for similar Salphen-like complexes of the order of magnitude stronger than that toward GC sequences.
same three metals, presenting the same triethylammonium- These results suggest that the strong affinity of the title metal
methyl groups.54,79 Finally, the complexes show an interaction complexes toward ct-DNA is essentially attributable to their
stoichiometry that also increases going from about 0.7 with 1 binding with AT sequences. Interestingly, the base-pair compo-
to 1 with 3. sition of native ct-DNA is roughly 60% AT and 40% GC.80

6112 | Dalton Trans., 2014, 43, 6108–6119 This journal is © The Royal Society of Chemistry 2014
Dalton Transactions Paper

Table 2 Binding constant (Kb) and stoichiometry (s) of compound 1 up to [ML2+]/[DNAphosphate] molar ratios of approximately 0.6,
with [ poly(dA-dT)]2 and [ poly(dG-dC)]2 0.6 and 0.4 respectively. The CD of native DNA is drastically
modified by the addition of increasing amounts of the three
[Poly(dA-dT)]2 [Poly(dG-dC)]2
metal complexes. In particular, a decrease and a blue shift of
Kb (1.3 ± 0.3) × 107 (2.0 ± 0.2) × 106 the positive CD band of DNA is observed in both titrations of 1
s 0.70 ± 0.03 0.73 ± 0.01 (Fig. 3a) and 2 (Fig. 3b).
The positive CD band of DNA becomes negative at metal
complex concentrations of 1 and 2 higher than 5 μM and
10 μM, respectively (Fig. 3a–b). Induced CD bands are also
evident for both complexes, at 356 and 479 nm for 1 (Fig. 3a)
and at 325 and 426 nm for 2 (Fig. 3b). Moreover, the DNA
dichroic band at 275 nm (black line in Fig. 3c) is mono-
tonously increased and split into two bands at 266 and 287 nm,
by the addition of increasing amounts of 3. Induced CD bands
appear at approximately 405, 447 and 513 nm. These results
indicate that deep conformational changes occur in the
double-helical structure of DNA, due to the interaction with
the metal complexes. The appearance of induced CD bands
shows that each of the three metal complexes supplies a
further chromophore tightly appended to the chiral backbone
of the DNA double helix.84 Interestingly, the CD spectrum
recorded in the presence of 1 and 2 at concentrations higher
than 15 μM is similar to that observed for the condensed ψ-
DNA forms.85 This would indicate that the nickel(II) and
copper(II) complexes at higher concentrations induce the for-
mation of supramolecular DNA aggregates. On the other hand,
the zinc complex 3 does not induce such supramolecular DNA
organization, possibly following its relatively lower DNA-
binding affinity. Similar results were obtained with analogue
intercalator Salphen-like nickel(II), copper(II) and zinc(II)
complexes.54,77,79
Thermal denaturation profiles of calf thymus DNA solu-
tions, in the presence of increasing amounts of 1–3, are shown
in Fig. 4 and were obtained by plotting the absorbance
at 258 nm as a function of temperature. As a matter of fact,
with the increasing of temperature the double stranded DNA
gradually dissociates into single strands: the DNA melting

Fig. 3 Circular dichroism spectra of ct-DNA (black) in the presence


of increasing amounts of 1 (a), 2 (b) and 3 (c) in 1.0 mM Tris-HCl. (a)
[DNAphosphate] = 50 μM, [1] = 0–30 μM. (b) [DNAphosphate] = 50 μM, [2] =
0–26 μM. (c) [DNAphosphate] = 50 μM [3] = 0–20 μM. The arrows indicate
the trend of the spectral change upon DNA addition.

Circular dichroism (CD) is a useful technique to monitor


changes in the DNA morphology due to drug–DNA inter-
Fig. 4 Examples of thermal denaturation profiles of ct-DNA 50 μM in
actions, since CD signals are sensitive to small variations in
the presence of the three complexes 1 (diamonds), 2 (triangles) and 3
the chiral conformation of DNA.81–83 For this reason, CD (circles), in 1.0 mM Tris-HCl, at the metal complex/DNAphosphate molar
spectra of ct-DNA 50 µM, in 1.0 mM Tris-HCl (Fig. 3), were ratio, R1, 0.10. The thermal denaturation profile of ct-DNA is represented
recorded in the presence of increasing amounts of 1, 2, and 3, by squares.

This journal is © The Royal Society of Chemistry 2014 Dalton Trans., 2014, 43, 6108–6119 | 6113
Paper Dalton Transactions

Table 3 Increase of the DNA melting temperature (ΔTm) in the pres- increase in viscosity of DNA since it increases separation of
ence of the three metal complexes at the indicated metal complex/ base pairs at intercalation sites, hence an increase in overall
DNAphosphate molar ratio (R1)
double-helical axial length occurs.88 Therefore these results
R1 1 2 3 confirm that the three metal complexes are DNA-intercalators.
Moreover, the slope of the three linear trends decreases in the
0.05 10 °C 9 °C 6 °C order 1 > 2 > 3. More in detail, the three correlation coeffi-
0.10 19 °C 14 °C 13 °C
cients of the linear fit of the data are higher than 0.99 and the
slopes of the three linear trends are 0.074, 0.042 and 0.023, in
temperature (Tm) is defined as the temperature where half of the presence of 1, 2 and 3, respectively. These results indicate
the total base pairs is unpaired,86 and corresponds to the that the relative viscosity raise in the presence of the nickel(II)
inflection point in the sigmoidal plots of Fig. 4. The DNA and copper(II) complexes, 1 and 2, is more than three times
melting temperature is therefore strictly related to the stability and roughly two times, respectively, than that observed in the
of the helix and any interaction of chemicals with DNA alters presence of the zinc(II) complex 3. In conclusion, the interca-
its Tm, stabilizing or destabilizing the final complex. Intercala- lating affinity of the three complexes perfectly follows the same
tion of any compound between DNA base pairs results in a decreasing binding affinity order found with both spectro-
quite strong stabilization of the helix that produces an increase photometric and thermal denaturation experiments.
of melting temperature of DNA.87 Moreover, the presence of
positive charges on the DNA-intercalator, precisely on the two
MD simulations and DFT/MM calculations
triethylammoniummethyl groups, besides providing a remark-
able water solubility to the three complexes, further improves MD simulations have been performed on the intercalation
the interaction by means of attractive interactions with the complex 3/[dodeca(dA-dT)]2. The root mean square deviation
negatively charged sugar-phosphate backbone of DNA. In (RMSD) along the simulation for all non-hydrogen atoms is
detail, the melting temperature of ct-DNA 50 μM in Tris-HCl shown in the ESI, Fig. S1.† The metal complex remains inside
1 mM (63.5 ± 1 °C, black line in Fig. 4) increases to 10, 9 and the binding pocket for the whole simulation time.
6 °C, at the metal complex/DNAphosphate molar ratio, R1 = 0.05, The equilibrium geometry, after about 50 ns, has been used
and of 19, 14 and 13 °C at R1 equal to 0.10, for 1, 2 and 3, as a starting point for further geometry optimizations, by
respectively (Table 3). hybrid two-layer QM/MM calculations, using DFT as the QM
These results are indicative of a strong metal complex–DNA method and the Amber99 force field as the MM method, as
interaction, which leads to a stabilization of the native DNA recently reported,32 of the intercalation complexes of the three
conformation. They also confirm the order of DNA-binding metal complexes 1–3 with [dodeca(dA-dT)]2 (Fig. 6).
strength obtained for the three metal complexes by spectro- The Cartesian coordinates of the three intercalation com-
photometric titrations: 1 > 2 > 3. plexes are reported in the ESI.† The higher layer of the DFT/
The relative viscosity of a solution of DNA 5 × 10−5 M line- MM structures shown in Fig. 6 involves the four nitrogen bases
arly increases with the addition of 1, 2 and 3, at metal of the intercalation pocket and the metal complex. The
complex/DNAphosphate molar ratios equal to 0, 0.05, 0.10, 0.15 selected DFT functional, M06-2X, allows to reliably describe
and 0.20, where η0 is the viscosity of the ct-DNA solution weak interactions in implicit water solution, notably H-bond
(Fig. 5). It is known that an intercalative interaction causes an and π–π stacking89 that are fundamental in the binding
between the metal complex and the DNA biomolecule.
Significant structural details can be obtained by the ana-
lysis of the optimized structures reported in Fig. 6, which
provide atomic level models explaining the strong DNA-
binding experimentally detected. In particular, the following
three are the complementary contributions ruling the DNA–
metal complex interaction: (1) the electrostatic attraction
between the two positively charged residues of the metal com-
plexes, the triethylammoniummethyl groups and the two nega-
tively charged phosphate groups; (2) the intercalation of the
naphthalene moiety between the stacked and H-bonded nitro-
gen bases of the intercalation pocket; (3) the metal coordi-
nation by exocyclic donor atoms of the nitrogen bases, as
summarized in Table 4.
Concerning the first contribution, electrostatic stabilizing
attraction, it is interesting to see that the triethylammonium-
Fig. 5 Relative viscosity of ct-DNA solutions 50 μM in the presence of
the three metal complexes 1 (green diamonds), 2 (blue triangles) and 3
methyl groups are at a short distance from the nearest-
(red circles), in 1.0 mM Tris-HCl, at the indicated metal complex/ neighbor phosphate groups of the DNA backbone, with
DNAphosphate molar ratio, R1. minimum P–N distances (of the phosphate and of the

6114 | Dalton Trans., 2014, 43, 6108–6119 This journal is © The Royal Society of Chemistry 2014
Dalton Transactions Paper

Table 5 Formation energy,a in kJ mol−1, in terms of standard enthalpy


(ΔH°) and Gibbs free energy (ΔG°), calculated at 298.15 K for the com-
plexes of 1, 2 and 3 with [dodeca(dA-dT)]2

ΔH° ΔG° ΔH° ΔG°


Model system (vacuo) (vacuo) (water) (water)

1/[dodeca(dA-dT)]2 −141.1 −96.5 −102.8 −58.2


2/[dodeca(dA-dT)]2 −153.3 −89.5 −112.8 −49.0
3/[dodeca(dA-dT)]2 −192.8 −171.2 −54.3 −32.7
a
The formation energy was evaluated by the following equation, where E
can be either H or G: ΔE° = E°(ML2+/[dodeca(dA-dT)]2) − E°([dodeca(dA-
dT)]2) − E°(ML2+).

the hypothesis that, in the intercalation complexes with DNA,


while the copper and zinc ions are axially coordinated in an
asymmetrical octahedral environment, the nickel ion is not and
preserves its square planar coordination geometry (Table 4).
Interestingly, the analysis of the optimized geometries
allows to nicely show that the decrease in the DNA binding
strength monotonously follows the distortion from the co-
planarity of the MO2N2 atoms in the three metal complexes, as
it often occurs in nickel(II), copper(II) and zinc(II) complexes of
the same chelating ligand.4
Fig. 6 Two different views of the intercalation site of the supramolecu-
lar complexes between the three metal complexes 1–3 with [dodeca- Standard enthalpy and Gibbs free energy values, calculated
(dA-dT)]2, highlighting the interatomic distances of the metal ion with at 298.15 K, were used to evaluate, in vacuo and in solution,
the O4 keto atom of thymine {in red, 3.02 Å (1), 2.57 Å (2) and 2.94 Å (3)} the formation energy of the supramolecular complexes of 1, 2
and with the N6 amine atom of adenine {in blue, 3.46 Å (1), 3.05 Å (2) and 3 with [dodeca(dA-dT)]2 (Table 5).
and 2.41 Å (3)}. DFT and MM regions are represented by “ball and
The analysis of the data in Table 5 allows us to make inter-
stick” and “sticks” styles, respectively.
esting considerations on the energetic contributions involved
in the DNA binding of the title metal complexes. First, the for-
Table 4 Calculated coordination distances (Å) of the metal ion of 1, 2 mation of the intercalation complexes is always exothermic,
and 3 (M = Ni, Cu, Zn) with the oxygen keto atom of thymine O4 and in vacuo and in solution. However, both entropy and solvation
with the amine nitrogen atom of adenine N6, in the intercalation pocket
seem to destabilize the DNA-binding energy. The structure and
of [dodeca(dA-dT)]2
electronic properties of the three intercalation complexes may
1 2 3 provide an explanation of the observed trends. The role of the
polar solvent can be rationalized taking into account that there
O4 3.02 2.57 2.94
N6 3.46 3.05 2.41
is a considerable electrostatic character in the interaction
energy that is screened going from the gas phase to water solu-
tion. However, the solvent destabilization of the zinc(II)
triethylammoniummethyl groups, respectively) in the range complex 3 is much larger than those of the nickel(II) and
5–11 Å for the three metal complex–dodecanucleotide systems. copper(II) complexes 1 and 2. This result finds nice support in
With regard to the intercalation of the naphthalene residue the calculated APT charge on the zinc atom, which is about 1.4,
between the stacked and H-bonded DNA bases, the pictures in also in the intercalation complexes, while those on the nickel and
Fig. 6 show that such a process is more pronounced for the copper atoms are about 1.0 and 1.2, respectively. The formation
nickel complex 1 and, in decreasing order, for the copper and free energy is always smaller than the formation enthalpy, both
the zinc ones 2 and 3. This trend is evidenced by the distortion in vacuo and in solution, indicating that the entropic contri-
from linearity of the N–M–O angle between the metal ion and bution, in the equation ΔG° = ΔH° − TΔS°, is always negative.
the axial exocyclic atoms N and O of the amine and keto However, such entropic destabilization is slightly larger for the
groups interatomic distances, highlighted in blue and red, intercalation complexes of [dodeca(dA-dT)]2 with 1, 2 and smaller
respectively. In detail, this angle assumes the following values: for that with the zinc complex 3. This result also follows the DNA-
172°, 175° and 168° for 3, 2 and 1, respectively. binding affinity, which is greater for 1 and 2 and smaller for 3.
Finally, concerning metal coordination, it has been Remarkably, excellent agreement exists between the calcu-
reported that N,O coordination distances within 2.5–3.0 Å, lated formation free energy values of the intercalation com-
revealed by X-ray crystallography for copper and zinc ions in plexes with [dodeca(dA-dT)]2 and the analogous formation free
active sites of biomolecules, are considered weak coordination energy values experimentally obtained from the DNA-binding
bonds.90 Following this rule, our DFT/MM calculations support constants reported in Table 1, as ΔG° = −RT ln(Kb). In

This journal is © The Royal Society of Chemistry 2014 Dalton Trans., 2014, 43, 6108–6119 | 6115
Paper Dalton Transactions

Fig. 7 Linear correlation plot between the experimental and calculated


formation free energies for 1, 2 and 3 with native ct-DNA and with
[dodeca(dA-dT)]2, respectively.

particular, the latter are −40.0, −36.9 and −33.1 kJ mol−1, for
1, 2 and 3, respectively. Moreover, the experimental DNA-
affinity trend 1 > 2 > 3 is reproduced by the calculation. In fact,
a good linear correlation (R = 0.995) was found between the
experimental and calculated formation Gibbs free energies, as
shown in Fig. 7. In conclusion, although it is known that there
are ten possible ways in which the four nitrogen bases can
produce intercalation sites,80 the DNA model selected for our
calculations correctly describes the DNA-binding interaction
mechanism of the title metal complexes.
Fig. 8 Zn K-edge XANES (a) and FT-EXAFS (b) of 3 457 μM (blue) and in
X-ray absorption spectroscopy the presence of ct-DNA 1.692 mM (green) in Tris-HCl 1 mM aqueous
solutions.
The XANES spectra of the zinc(II) complex 3 in aqueous solu-
tion, isolated (blue line) and in the presence of ct-DNA (green
line) are reported in Fig. 8a. The corresponding Fourier trans-
formed EXAFS (FT-EXAFS) spectra for the same samples are
shown in Fig. 8b. By comparison with the literature,91 it can be
concluded that the near-edge features in the XANES spectra of
both 3 and of 3/ct-DNA solutions are typical of Zn2+ ions in an
octahedral coordination environment.
The FT-EXAFS spectrum of 3 (Fig. 8b, blue line) shows a
broad first coordination peak, between 1.4 and 2.0 Å (un-
corrected for phase shift), and several other contributions
between 2 and 3 Å, arising from the multiple scattering paths
of the C, O and N atoms of the chelating ligand. Such features
are nicely reproduced by using as guess structure the opti-
mized geometry of the diaquo complex reported in the inset of
Fig. 9. There is in fact good agreement between the refined
structures obtained from EXAFS and DFT simulations: the Fig. 9 FT EXAFS data (blue), model (red) and residual (brown) of 3
457 μM in Tris-HCl 1 mM aqueous solutions. The DFT optimized geome-
experimental equatorial distances, Zn–N and Zn–O, are 2.1
try of the distorted octahedral complex 3, with two apically coordinated
and 1.9 Å, respectively, while those calculated are 2.1 and water molecules, is also shown.
2.0 Å, respectively. The DFT calculations predict the two water
molecules to be placed at 2.1 and 2.2 Å, tilted away from the che-
lating atoms, while the model used to fit the EXAFS shows the [DNAphosphate] molar ratio of about 0.27. Under these con-
best agreement with the apical O of the water molecules at 2.1 Å. ditions, the solution experiments reported above have shown
The EXAFS spectrum of the isolated complex is significantly the metal complex to be fully intercalated. Analogous aqueous
modified by the addition of an excess of DNA, at a [3]/ solutions of the nickel(II) and copper(II) complexes 1 and 2

6116 | Dalton Trans., 2014, 43, 6108–6119 This journal is © The Royal Society of Chemistry 2014
Dalton Transactions Paper

decreasing in the order Ni > Cu > Zn and with selective affinity


of the three metal complexes toward AT with respect to GC
sequences. DFT/MM calculations provided detailed local infor-
mation on the DNA-binding site, which consists of (1) electro-
static attraction between the lateral cationic groups of the
metal complexes and the anionic phosphate groups of the
DNA backbone, (2) the DNA-intercalation of the planar aro-
matic fragment of the chelating Shiff-base ligand, (3) the metal
ion coordination by the exocyclic keto-oxygen and amine-nitro-
gen of the nitrogen bases. The values of the DNA-binding con-
stants and their decreasing trend in the order 1 > 2 > 3 are
correctly reproduced by the calculated formation Gibbs free
energy values of the supramolecular complexes of 1, 2 and 3
with [dodeca(dA-dT)]2 in solution.
Fig. 10 FT EXAFS data (green), model (magenta) and residual (brown) Remarkably, for the first time, the occurrence of DNA–
of 3 457 μM in the presence of ct-DNA 1.692 mM in Tris-HCl 1 mM metal ion coordination in aqueous solution was confirmed for
aqueous solutions.
the zinc(II) complex 3 by the modifications of the experimental
EXAFS spectra of the metal complex in the presence of DNA
undergo photoreduction and/or photodegradation to some and by the excellent agreement between the first shell coordi-
extent, both as isolated species and in the presence of DNA, nation distances, found by analysis of the EXAFS, with those
when exposed to the synchrotron X-ray beam. The presence of obtained in the structure of 3/[dodeca(dA-dT)]2, optimized by
metal–metal distances (either from the formation of metal DFT/MM calculations.
nanoparticles or from polymerization of the complexes) was
detected in the spectra, which were not analyzed further.
In the presence of DNA, an overall change in the local struc- Acknowledgements
ture of Zn2+ is evident. Firstly, the first coordination shell loses
one or two apical H2O molecules, as witnessed by the decrease We gratefully acknowledge the CINECA award N. IsB05, year
in the broad contribution around 1.5 Å (Fig. 8b). Moreover, a 2012, under the ISCRA initiative, for the availability of high
sharp increase of the feature around 2.4 Å is visible. The high- performance computing resources and support, and the Euro-
layer structure optimized by DFT/MM calculations (Fig. 6) has pean Synchrotron Radiation Facility (ESRF) for the provision of
been used to reproduce the simulated FT-EXAFS of 3/ct-DNA synchrotron beam time (experiment SC-3581) and for financial
solutions (Fig. 10). support. We thank Dr S. Nikitenko, Dr D. Banerjee and the
Interestingly, the peak at 2.4 Å can be modelled by two N staff of the BM26 beamline of ESRF for assistance during the
and two O atoms at 2.1 Å and 2.0 Å, respectively, corres- XAS measurements.
ponding to the four equatorial atoms of the Schiff base, plus
one further N atom at about 2.5 Å and one O atom around
2.9 Å (Fig. 10). The six coordination distances are in excellent Notes and references
agreement with the calculated distances for Zn–N and Zn–O,
in 3/[dodeca(dA-dT)]2 (Fig. 6), in particular, with the Zn–N6 1 A. C. Komor and J. K. Barton, Chem. Commun., 2013, 49,
and Zn–O4 apical distances, 2.41 and 2.94 Å, respectively, 3617–3630.
involving the amine group of adenine and the keto group of 2 P. C. Bruijnincx and P. J. Sadler, Curr. Opin. Chem. Biol.,
thymine (Table 4). The latter prove the occurrence of a further 2008, 12, 197–206.
stable coordination of the metal ion with respect to the metal 3 J. Reedijk, Metallomics, 2012, 4, 628–632.
complex alone. 4 G. Barone, A. Terenzi, A. Lauria, A. M. Almerico, J. M. Leal,
N. Busto and B. García, Coord. Chem. Rev., 2013, 257, 2848–
2862.
Conclusions 5 A. Terenzi, L. Tomasello, A. Spinello, G. Bruno,
C. Giordano and G. Barone, J. Inorg. Biochem., 2012, 117,
Detailed information on the DNA-binding of three square 103–110.
planar nickel(II) (1), copper(II) (2) and zinc(II) (3) Schiff-base 6 A. Terenzi, C. Ducani, V. Blanco, L. Zerzankova,
complexes was obtained through the complementary appli- A. F. Westendorf, C. Peinador, J. M. Quintela,
cation of experimental investigations in water solution P. J. Bednarski, G. Barone and M. J. Hannon, Chem.–Eur. J.,
and quantum mechanics/molecular mechanics (DFT/MM) 2012, 18, 10983–10990.
calculations. 7 A. Terenzi, M. Fanelli, G. Ambrosi, S. Amatori, V. Fusi,
Spectroscopic and viscometry measurements confirmed L. Giorgi, V. T. Liveri and G. Barone, Dalton Trans., 2012,
that 1, 2 and 3 are DNA-intercalators with DNA-affinity 41, 4389–4395.

This journal is © The Royal Society of Chemistry 2014 Dalton Trans., 2014, 43, 6108–6119 | 6117
Paper Dalton Transactions

8 A. Lauria, I. Abbate, C. Patella, A. Martorana, G. Dattolo 33 C. M. Nunn and S. Neidle, J. Med. Chem., 1995, 38, 2317–
and A. M. Almerico, Eur. J. Med. Chem., 2013, 62, 416–424. 2325.
9 A. Lauria, C. Patella, I. Abbate, A. Martorana and 34 J. R. Quintana, A. A. Lipanov and R. E. Dickerson, Biochem-
A. M. Almerico, Eur. J. Med. Chem., 2012, 55, 375–383. istry, 1991, 30, 10294–10306.
10 A. Lauria, C. Patella, I. Abbate, A. Martorana and 35 H. Niyazi, J. P. Hall, K. O’Sullivan, G. Winter, T. Sorensen,
A. M. Almerico, Eur. J. Med. Chem., 2013, 65, 381–388. J. M. Kelly and C. J. Cardin, Nat. Chem., 2012, 4, 621–
11 M. J. Hannon, Chem. Soc. Rev., 2007, 36, 280–295. 628.
12 B. M. Zeglis, V. C. Pierre and J. K. Barton, Chem. Commun., 36 M. S. Searle, A. J. Maynard and H. E. L. Williams, Org.
2007, 4565–4579. Biomol. Chem., 2003, 1, 60–66.
13 T. Takagaki, T. Bando and H. Sugiyama, J. Am. Chem. Soc., 37 F. Coste, J.-M. Malinge, L. Serre, M. Leng, C. Zelwer,
2012, 134, 13074–13081. W. Shepard and M. Roth, Nucleic Acids Res., 1999, 27, 1837–
14 T. Yoshidome, M. Endo, G. Kashiwazaki, K. Hidaka, 1846.
T. Bando and H. Sugiyama, J. Am. Chem. Soc., 2012, 134, 38 B. Spingler, D. A. Whittington and S. J. Lippard, Inorg.
4654–4660. Chem., 2001, 40, 5596–5602.
15 H.-C. Tai, R. Brodbeck, J. Kasparkova, N. J. Farrer, 39 O. Julien, J. R. Beadle, W. C. Magee, S. Chatterjee,
V. Brabec, P. J. Sadler and R. J. Deeth, Inorg. Chem., 2012, K. Y. Hostetler, D. H. Evans and B. D. Sykes, J. Am. Chem.
51, 6830–6841. Soc., 2011, 133, 2264–2274.
16 J. Pracharova, L. Zerzankova, J. Stepankova, O. Novakova, 40 Y. Wu, D. Bhattacharyya, C. L. King, I. Baskerville-
N. J. Farrer, P. J. Sadler, V. Brabec and J. Kasparkova, Chem. Abraham, S.-H. Huh, G. Boysen, J. A. Swenberg, B. Temple,
Res. Toxicol., 2012, 25, 1099–1111. S. L. Campbell and S. G. Chaney, Biochemistry (Moscow),
17 R. B. Sears, L. E. Joyce, M. Ojaimi, J. C. Gallucci, 2007, 46, 6477–6487.
R. P. Thummel and C. Turro, J. Inorg. Biochem., 2013, 121, 41 B. Andersen and E. Sletten, J. Inorg. Biochem., 2000, 79,
77–87. 353–358.
18 M. P. Barrett, C. G. Gemmell and C. J. Suckling, Pharmacol. 42 A. A. Hummer and A. Rompel, in Advances in Protein Chem-
Ther., 2013, 139, 12–23. istry and Structural Biology, ed. C. Z. Christov, Academic
19 X. Cai, P. J. Gray Jr. and D. D. Von Hoff, Cancer Treat. Rev., Press, 2013, vol. 93, pp. 257–305.
2009, 35, 437–450. 43 H. Song, D. L. Wilson, E. R. Farquhar, E. A. Lewis and
20 H.-K. Liu and P. J. Sadler, Acc. Chem. Res., 2011, 44, 349– J. P. Emerson, Inorg. Chem., 2012, 51, 11098–11105.
359. 44 R. Ortega, Metallomics, 2009, 1, 137–141.
21 N. J. Wheate, C. R. Brodie, J. G. Collins, S. Kemp and 45 W. Shi, M. Punta, J. Bohon, J. M. Sauder, R. D’Mello,
J. R. Aldrich-Wright, Mini Rev. Med. Chem., 2007, 7, 627– M. Sullivan, J. Toomey, D. Abel, M. Lippi, A. Passerini,
648. P. Frasconi, S. K. Burley, B. Rost and M. R. Chance, Genome
22 L. González-Bulnes and J. Gallego, Biopolymers, 2012, 97, Res., 2011, 21, 898–907.
974–987. 46 R. W. Strange, M. Ellis and S. S. Hasnain, Coord. Chem.
23 D. Xu, X. Wang, D. Fei and L. Ding, Nucleosides Nucleotides Rev., 2005, 249, 197–208.
Nucleic Acids, 2010, 29, 854–866. 47 P. McPhie, Methods Enzymol., 1971, 22, 23–32.
24 S. Zhou, Y. Fu, X. Fan, Y. Zhang and C. Li, Med. Chem. Res., 48 S. D. Kennedy and R. G. Bryant, Biophys. J., 1986, 50, 669–
2012, 22, 2862–2869. 676.
25 J. B. Chaires, in Anthracycline Antibiotics, American Chemical 49 V. I. Ivanov and E. E. Minyat, Nucleic Acids Res., 1981, 9,
Society, 1994, vol. 574, pp. 156–167. 4783–4798.
26 E. J. Gabbay, D. Grier, R. E. Fingerle, R. Reimer, R. Levy, 50 R. B. Gennis and C. R. Cantor, J. Mol. Biol., 1972, 65, 381–
S. W. Pearce and W. D. Wilson, Biochemistry, 1976, 15, 399.
2062–2070. 51 B. Ravel and M. Newville, J. Synchrotron Radiat., 2005, 12,
27 C. Temperini, L. Messori, P. Orioli, C. D. Bugno, F. Animati 537–541.
and G. Ughetto, Nucleic Acids Res., 2003, 31, 1464–1469. 52 K. V. Klementev, J. Phys. Appl. Phys., 2001, 34, 209–217.
28 K. J. Kilpin, C. M. Clavel, F. Edafe and P. J. Dyson, Organo- 53 A. L. Ankudinov, B. Ravel, J. J. Rehr and S. D. Conradson,
metallics, 2012, 31, 7031–7039. Phys. Rev. B: Condens. Matter, 1998, 58, 7565–7576.
29 L. W. Chung, H. Hirao, X. Li and K. Morokuma, Wiley Inter- 54 A. Silvestri, G. Barone, G. Ruisi, D. Anselmo, S. Riela and
discip. Rev. Comput. Mol. Sci., 2012, 2, 327–350. V. T. Liveri, J. Inorg. Biochem., 2007, 101, 841–848.
30 F. H. Wallrapp and V. Guallar, Wiley Interdiscip. Rev. 55 J. W. Ponder, TINKER, 4.2 Software Tools for Molecular
Comput. Mol. Sci., 2011, 1, 315–322. Design, http://dasher.wustl.edu/tinker
31 H. Paulsen, A. X. Trautwein, P. Wegner, C. Schmidt, 56 W. Saenger, Principles of Nucleic Acid Structure, Springer
A. I. Chumakov and V. Schünemann, ChemPhysChem, 2011, Verlag, New York, 1984.
12, 3434–3441. 57 D. Van Der Spoel, E. Lindahl, B. Hess, G. Groenhof,
32 A. Spinello, A. Terenzi and G. Barone, J. Inorg. Biochem., A. E. Mark and H. J. C. Berendsen, J. Comput. Chem., 2005,
2013, 124, 63–69. 26, 1701–1718.

6118 | Dalton Trans., 2014, 43, 6108–6119 This journal is © The Royal Society of Chemistry 2014
Dalton Transactions Paper

58 B. Hess, C. Kutzner, D. van der Spoel and E. Lindahl, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar,
J. Chem. Theory Comput., 2008, 4, 435–447. J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene,
59 J. Wang, P. Cieplak and P. A. Kollman, J. Comput. Chem., J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo,
2000, 21, 1049–1074. R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin,
60 A. Pérez, I. Marchán, D. Svozil, J. Sponer, T. E. Cheatham R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin,
III, C. A. Laughton and M. Orozco, Biophys. J., 2007, 92, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador,
3817–3829. J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas,
61 A. T. Guy, T. J. Piggot and S. Khalid, Biophys. J., 2012, 103, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaus-
1028–1036. sian 09, Revision A.1, Gaussian Inc., Wallingford CT, 2009.
62 A. W. Sousa Da Silva, W. F. Vranken and E. D. Laue, 75 M. D. Hobday and T. D. Smith, Coord. Chem. Rev., 1973, 9,
ACPYPE – Antechamber Python Parser Interface, http:// 311–337.
code.google.com/p/acpype 76 T. Uno, K. Hamasaki, M. Tanigawa and S. Shimabayashi,
63 J. Wang, R. M. Wolf, J. W. Caldwell, P. A. Kollman and Inorg. Chem., 1997, 36, 1676–1683.
D. A. Case, J. Comput. Chem., 2004, 25, 1157–1174. 77 A. Terenzi, C. Ducani, L. Male, G. Barone and
64 J. Wang, W. Wang, P. A. Kollman and D. A. Case, J. Mol. M. J. Hannon, Dalton Trans., 2013, 42, 11220–11226.
Graph. Model., 2006, 25, 247–260. 78 M. T. Carter, M. Rodriguez and A. J. Bard, J. Am. Chem.
65 G. Bussi, D. Donadio and M. Parrinello, J. Chem. Phys., Soc., 1989, 111, 8901–8911.
2007, 126, 014101-1–014101-7. 79 G. Barone, N. Gambino, A. Ruggirello, A. Silvestri,
66 M. Parrinello and A. Rahman, J. Appl. Phys., 1981, 52, A. Terenzi and V. Turco Liveri, J. Inorg. Biochem., 2009, 103,
7182–7190. 731–737.
67 T. Darden, D. York and L. Pedersen, J. Chem. Phys., 1993, 80 G. Barone, C. Fonseca Guerra and F. M. Bickelhaupt, Chem.
98, 10089–10092. Open, 2013, 2, 186–193.
68 M. Svensson, S. Humbel, R. D. J. Froese, T. Matsubara, 81 M. J. Waring, J. Mol. Biol., 1965, 13, 269–274.
S. Sieber and K. Morokuma, J. Phys. Chem., 1996, 100, 82 A. Rodger and B. Nordén, Circular dichroism and linear
19357–19363. dichroism, Oxford University Press, Oxford, New York,
69 T. Vreven and K. Morokuma, J. Comput. Chem., 2000, 21, 1997.
1419–1432. 83 J. Kypr, I. Kejnovská, D. Renčiuk and M. Vorlíčková, Nucleic
70 Y. Zhao and D. G. Truhlar, Theor. Chem. Acc., 2008, 120, Acids Res., 2009, 37, 1713–1725.
215–241. 84 M. J. Carvlin, N. Dattagupta and R. J. Fiel, Biochem.
71 N. Godbout, D. R. Salahub, J. Andzelm and E. Wimmer, Biophys. Res. Commun., 1982, 108, 66–73.
Can. J. Chem., 1992, 70, 560–571. 85 J. E. B. Ramos, R. de Vries and J. R. Neto, J. Phys. Chem. B,
72 J. Tomasi, B. Mennucci and R. Cammi, Chem. Rev., 2005, 2005, 109, 23661–23665.
105, 2999–3093. 86 Y.-J. Liu, H. Chao, F.-L. Tau, Y.-X. Yuan, W. Wei and L.-N. Ji,
73 G. Scalmani and M. J. Frisch, J. Chem. Phys., 2010, 132, J. Inorg. Biochem., 2005, 99, 530–537.
114110. 87 J. M. Kelly, A. B. Tossi, D. J. Mcconnell and C. Ohuigin,
74 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, Nucleic Acids Res., 1985, 13, 6017–6034.
M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, 88 H.-L. Huang, Y.-J. Liu, C.-H. Zeng, L.-X. He and F.-H. Wu,
B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, DNA Cell Biol., 2010, 29, 261–270.
X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, 89 R. F. Ribeiro, A. V. Marenich, C. J. Cramer and
J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, D. G. Truhlar, Phys. Chem. Chem. Phys., 2011, 13, 10908–
J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, 10922.
H. Nakai, T. Vreven, J. A. Montgomery Jr., J. E. Peralta, 90 M. M. Harding, Acta Crystallogr. Sect., 2000, 56, 857–867.
F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, 91 C. Hennig, K.-H. Hallmeier, G. Zahn, F. Tschwatschal and
K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, H. Hennig, Inorg. Chem., 1999, 38, 38–43.

This journal is © The Royal Society of Chemistry 2014 Dalton Trans., 2014, 43, 6108–6119 | 6119

You might also like