Vous êtes sur la page 1sur 9

Bioorganic & Medicinal Chemistry 17 (2009) 6054–6062

Contents lists available at ScienceDirect

Bioorganic & Medicinal Chemistry


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

Characterization and DNA-interaction studies of 1,1-dicyano-2,2-ethylene


dithiolate Ni(II) mixed-ligand complexes with 2-amino-5-methyl thiazole,
2-amino-2-thiazoline and imidazole. Crystal structure of [Ni(i-MNT)(2a-5mt)2]
Philip J. Cox a, George Psomas b, Christos A. Bolos b,*
a
School of Pharmacy, The Robert Gordon University, Schoolhill, Aberdeen AB10 1FR, Scotland, UK
b
Laboratory of Inorganic Chemistry, Chemistry Department, Aristotle University of Thessaloniki, GR-541 24, Greece

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

Article history: A series of mixed-ligand neutral nickel(II) complexes of the general formula [Ni(i-MNT)(2a-5mt)2] (1),
Received 14 April 2009 [Ni(i-MNT)(2a-2tzn)2] (2) and [Ni(i-MNT)(Im)2] (3), [where i-MNT2 = the dianion of 1,1-dicyano-2,2-
Revised 18 June 2009 ethylenedithiolate, 2a-5mt = 2-amino-5-methyl thiazole, 2a-2tzn = 2-amino-2-thiazoline and Im = imid-
Accepted 20 June 2009
azole] were prepared and characterized with elemental analyses, spectroscopic (IR, UV–vis) methods,
Available online 30 June 2009
magnetic susceptibility, molar conductivity and cyclic voltammetry measurements. The magnetic data,
the electronic spectra and the electrical conductivity measurements indicated mononuclear neutral com-
Keywords:
plexes with square-planar geometry. The X-ray analysis of [Ni(i-MNT)(2a-5mt)2] shows the nickel atom
Ni(II) complexes
i-MNT
being fourfold coordinated with the two sulfur atoms of the dithiolate (i-MNT) ligand and the endocyclic
2-Amino-5-methyl-thiazole nitrogen atoms from the two 2a-5mt ring giving rise to a slightly distorted square-planar arrangement.
2-Amino-2-thiazoline The cyclic voltammograms of the complexes have been recorded and the corresponding redox potentials
Imidazole have been estimated. The DNA-binding studies of the complexes has been evaluated by examining their
Interaction with calf-thymus DNA ability to bind to calf-thymus DNA (CT DNA) with UV spectroscopy and cyclic voltammetry. Both studies
Cyclic voltammetry have shown that the complexes can bind to CT-DNA by the intercalative and the electrostatic binding
mode. Competitive binding studies with ethidium bromide (EB) with fluorescence spectroscopy have also
shown that the complexes exhibit the ability to displace the DNA-bound EB indicating that they can bind
to DNA in strong competition with EB.
Ó 2009 Elsevier Ltd. All rights reserved.

1. Introduction In our previous studies a series of mixed-ligand Cu(II) coordina-


tion compounds, with bidentate, tridentate and polydentate
A large number of coordination compounds with metal ions amines and their Schiff bases,7,8 trihalides,9 1,2-dithiolates,10 and
have been investigated for their anti-proliferative activity and used 1,3-thiazoles, thiazolines and imidazole11–14 have been utilized
in cancer therapy.1 However their several side-effects leave room as models to investigate their toxicity,10,12 anti-inflammatory
for the selection of other metals and additional ligands, known activity,12 antimicrobial activity against Gram(+) and Gram() bac-
for their bioactivity.2 A principal target of metals in the cancer cells teria9,11,14 and anti-proliferative activity in vitro9,11,13,14 and
is their coordination with DNA, bound selectively to it either in vivo10,12 towards various cancer cell lines, for example, HeLa,
through the oxygen of phosphates or to heterocyclic nitrogen T47D, HT-29, MCF7, MRC5 and P388 leukemia. A relationship be-
atoms (N1, N3 and N7) of DNA bases, or both.3 Nickel is recognized tween medicinal effectiveness and physicochemical, structural
as an essential trace element for bacteria and plants, but its role in properties of the ligands and their Cu(II) complexes, have been car-
animal biochemistry is not well defined.4 Some of the Ni(II) com- ried out. Among the ligands characterized for individual anti-pro-
plexes with dimethyldithiophosphates ligands is reported to be ac- liferative activity are those of S,N-1,3-thiazoles and especially
tive against Walker 256 carcinosarkoma tumor system, but they their Cu(II) complexes with 2-amino-5-methyl thiazole (2a-5mt)
were found to be inactive against L1210 leukemia.5 Regarding which were found to meet the necessary criteria established by Na-
the DNA interaction with Ni(II) complexes showed intercalative tional Cancer Institute.15 It has also been demonstrated that these
behavior and also DNA cleavage ability.6 Cu(II) complexes induce apoptosis via mitochondrial pathway. In
these studies, although the molar ratio of the reactants (Ligand/
Cu(II)/2a-5mt) was 1:1:2, the isolated complexes conform to
* Corresponding author. Tel.: +30 2310 997705; fax: +30 2310 997738.
1:1:1 stoichiometries. Our efforts to insert one more 2a-5mt
E-mail address: bolos@chem.auth.gr (C.A. Bolos). molecule into the coordination sphere of Cu(II), which might

0968-0896/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.bmc.2009.06.058
P. J. Cox et al. / Bioorg. Med. Chem. 17 (2009) 6054–6062 6055

differentiate further the bioactivity of the compounds under study, 1630 cm1. It can be observed that the lower limit is slightly
failed. This disadvantage of Cu(II) led us to carry out similar reac- shifted to higher frequencies upon coordination, while the upper
tions with the Ni(II) since this metal adopts more easily various limit is shifted to lower frequencies by 1–17 cm1. This picture
geometries and characteristic magnetic and redox properties. In may be tentatively explained in terms of the highly mobile p-elec-
spite of the toxicity of Ni(II) complexes, ternary Ni(II) complexes trons within the thiazole and imidazole conjugated systems upon
are formed with DNA, oligopeptides and proteins under physiolog- the coordination through the endocyclic nitrogen.8,21 The IR peaks
ical conditions.16,17 The latter had reversible Ni(III, II) redox behav- of all complexes are almost similar, indicating that the complexes
ior, whereas Ni(II) complexes with bases were able to induce, may have similar structures.
cooperatively, conformational changes in synthetic DNA.18 In ef-
fect, what we wanted to address in this project was the synthesis 2.1.2. UV–vis spectral studies
and characterization of neutral mixed-ligand dithiolato Ni(II) com- The band maxima of the electronic excitation spectra of the
plexes with 2a-5mt, 2a-2tzn and imidazole, with the aim of exam- complexes in DMSO solution were studied in the visible and UV re-
ining (i) the possible coordination of a second 2a-5mt molecule gions and are given in the experimental part. The spectra of com-
with Ni(II), (ii) the resultant geometry and (iii) the biological activ- plexes 1–3 exhibit a broad absorption band of low intensity
ity of the new Ni(II) compounds. Prior to the examination of their (e = 38–45 M1 cm1) in the range 622–634 nm. This band could
possible bioactivity we thought that it would be important to per- be attributed to d?d transition for the central Ni(II), probably of
2
form various studies of the compounds, such as reactivity with the dz ?dx2 y2 type.22 The band or shoulder at ca. 457 nm for all
calf-thymus (CT) DNA with cyclic voltammetry (CV) and UV and complexes could be characterized as charge transfer band. The
fluorescence spectroscopy. The results of these studies are dis- transfer could take place from the metal to the dithiolate ligand
cussed in details and found to be promising for biological studies. (metal-to-ligand charge transfer) a typically d(M)?p* (S2C@) tran-
sition (most probable) or, on the contrary, from the ligand to the
2. Results and discussion metal (ligand-to-metal charge transfer). This band is due to the
extensive conjugation of the dithiolate ligand and is red-shifted
The reaction of [Ni(L)4Cl2] with Na2(i-MNT)3H2O affords brown in the spectra of the complexes, compared to that of the free ligand
mixed-ligand complexes corresponding to the general formulas (410 nm). Credence to the above arguments is given by the ob-
[Ni(i-MNT)(2a-5mt)2], [Ni(i-MNT)(2a-2tzn)2] and [Ni(i-MNT)- served strong shift of the IR m(CS2) band at higher wavenumbers
(Im)2]. The new chelates are slightly soluble in methanol, soluble (vide supra). Finally, the band at 340 nm is due to internal transi-
in acetonitrile and are very soluble in acetone, DMSO and DMF. tions of the ligands and can be attributed to n?p* or p?p* transi-
The analytical data C, H, N and Ni confirm the composition of the tions.23 The similarity of the position of these bands in the spectra
complexes and the stoichiometry proposed. Evidence of the mode of the complexes should be noted, indicating similar chromophore
of bonding of the ligands was gathered from the IR spectra, while with different N-coordinated ligands.
their geometry was proposed by electronic excitation spectra and
confirmed with X-ray determination of [Ni(i-MNT)(2a-5mt)2]. 2.1.3. Molar conductivity measurements
The molar conductivity KV values of the complexes in 103 V
2.1. Spectroscopic data DMSO solutions vary from 4 to 8 lS cm1, thus revealing their
non-electrolytic character.24
2.1.1. IR spectral studies
The most significant bands emerging from the functional group 2.1.4. Magnetic measurements
of i-MNT and from the 2a-5mt, 2a-2tzn and imidazole and their The low magnetic measurement values (0.2 BM) of the Ni(II)
complexes are given in the experimental section. The dianion of complexes shows that they are diamagnetic giving rise to a slightly
the 1,1-dicyano-2,2-ethylene dithiolate contain the C„N, the distorted square-planar arrangement. This is a natural conse-
C@CS2, and the C-SS functional groups and while the absorption quence of the low spin Ni(II) d8 configuration and the eight elec-
due to the multiple bond in each group is easily distinguishable. trons being paired, can occupy the four low lying d orbitals.
The very strong bands at 2175 cm1 and in the region 1356–
1436 cm1 have been assigned to m(C„N) and m(C@C), respec- 2.2. Description of the structure of [Ni(i-MNT)(2a-5mt)2]
tively, of the i-MNT ligand.19 Upon coordination, these bands are
shifted to higher wavenumbers: the former by 28–53 cm1 and The molecular structure of the [Ni(i-MNT)(2a-5mt)2] complex
the latter by 35–93 cm1 compared to the free i-MNT ligand. These along with the atomic numbering scheme is shown in Figure 1, se-
shifts reveal an increase of the order of the C„N and ˜C@C— bonds, lected bond lengths and angles are given in Table 1.
since the planar extensive p-conjugated system of the i-MNT li- The nickel atom is four-coordinated with the two sulfur atoms
gand facilitate a drift of electron density towards the sulfur atoms of the dithiolate ligand (i-MNT) and the ring-nitrogen atoms of
that results in a shortening of the C„N and ˜C@C— bonds.20 The the two 2a-5mt ligands giving rise to a slightly distorted square-
medium intensity bands at 943–989 cm1 and 794–896 cm1 in planar arrangement with the maximum deviation of the fitted
the IR spectra are assigned to mas(CS2) and msym(CS2), respectively. atoms from the mean S2N2 plane being 0.084(2) Å. The central
These two bands are shifted to lower (13–38 cm1) or higher (9– nickel atom lies in this plane within experimental error, its out-
28 cm1) wavenumbers in the infrared spectra of the complexes. of-plane distance being 0.011(1) Å. Severe distortion with respect
The variation in these shifts could be due to the different donation to the idealized square-planar geometry is observed in the coordi-
ability of the 2a-5mt, 2a-2tzn and imidazole ligands. Other infrared nation angles. The small S–Ni–S chelation angle of 78.85(2)° indi-
data further confirming the coordination of the ligands, localized in cates that it is determined by the geometry of the rigid i-MNT
the S–Ni–S and the Ni–N groups are the new medium or weak ligand.19,20
intensity peaks appearing at 478–486 cm1 and 387–396 cm1, On the other hand, as a result of the relatively small separation
which may be attributed to the m(Ni–N) and m(Ni–S), respectively. of the coordinated nitrogen atoms of the flexible 2a-5mt ligands
The bands in the range 1631–1531 cm1 in the spectra of the free (2.851 Å in comparison to 3.10 Å, which is the sum of their van
thiazole, thiazoline or imidazole ligands can be assigned to m(C@N) der Waals radii) the opposite N–Ni–N angle appears greater than
and m(C@C) of their rings. In the spectra of the complexes this re- the ideal value of 90° [95.26(7)°]. The two Ni–S distances are very
gion gives rise to a number of bands located in the range 1532– close to each other and have a mean value of 2.189(1) Å, which is in
6056 P. J. Cox et al. / Bioorg. Med. Chem. 17 (2009) 6054–6062

cule in one layer and the N atoms of the i-MNT ligands of two other
molecules, lying either side of this layer.

2.3. Cyclic voltammetry for the complexes

The complete scan in the range 1.5 V to +1.0 V of complex 1 in


0.4 mM DMSO solution shows two cathodic waves at 720 mV and
1360 mV and three anodic waves at +90 mV, +565 mV and
+850 mV (Fig. 2A). These two one-electron cathodic waves suggest
that the reduction of the mononuclear complex 1 takes place in
two steps.31 In the reverse scan, one new cathodic wave at
+10 mV appears. Scanning in the range +1.0 to 1.0 V the anodic
wave at +80 mV disappears, confirming that this wave can be as-
signed to the process [Ni0]?[NiI]. Scanning in the range 1.0 to
+0.5 V the cathodic wave at +10 mV and the anodic wave at
+880 mV disappear, confirming that these waves can be assigned
to the couple Ni(II)/Ni(III).
Figure 1. The molecular structure and atom labeling of [Ni(i-MNT)(2a-5mt)2].
In 0.4 mM 1/2 DMSO/buffer solution, the waves are not distin-
guished as clear as in DMSO solution. The cyclic voltammogram
Table 1
Selective bond lengths (Å) and angles (°) for complex 1

Bond lengths (Å) Bond angles (°) A -200


Ni(1)–N(3) 1.9225(15) N(3)–Ni(1)–N(1) 95.26(7)
Ni(1)–N(1) 1.9368(15) N(3)–Ni(1)–S(3) 170.35(5) 0
Ni(1)–S(3) 2.1839(5) N(1)–Ni(1)–S(3) 93.76(5)
Ni(1)–S(4) 2.1934(5) N(3)–Ni(1)–S(4) 92.30(5)
S(3)–C(9) 1.7140(18) N(1)–Ni(1)–S(4) 172.07(5)
I (μA) 200
S(4)–C(9) 1.7200(18) S(3)–Ni(1)–S(4) 78.847(18)
N(5)–C(11) 1.148(2) C(9)–S(3)–Ni(1) 86.50(6) 400
N(6)–C(12) 1.149(2) C(9)–S(4)–Ni(1) 86.05(6)
C(9)–C(10) 1.383(3) C(10)–C(9)–S(3) 125.17(14)
C(10)–C(12) 1.422(3) C(10)–C(9)–S(4) 126.71(14) 600
C(10)–C(11) 1.424(3) S(3)–C(9)–S(4) 108.09(10)
800
1.0 0.5 0.0 -0.5 -1.0 -1.5

very good agreement with the corresponding ones observed in E (V)


other dithiolato complexes, having the Ni(II) atom in a square-pla-
nar coordination.25–28 The first 2a-5mt ring, comprising the atoms B -100
N1,C1,C2,C3 and S1 is strictly planar within the experimental error
(3r), the maximum deviation from the mean plane being -50
0.005(2) Å. On the other hand, the second thiazole ring, comprising
I (μA)

the atoms N3, C5, C6, C7 and S2, exhibits a small puckering, the 0
maximum deviation from the mean plane being 0.013(2) Å.
The i-MNT ligand is planar [maximum deviation of the fitted 50
atoms from their best plane: 0.087(2) Å], having the approximate
symmetry C2v and forming with the plane of the chromophore 100
NiS2N2 an angle of 10.20(3)°. The nickel—i-MNT ligand bond for-
mation does not affect the average C–S bond length or influences 150
0.5 0.0 -0.5 -1.0 -1.5
appreciably the C@CS2 interatomic distance in comparison with
E (V)
the observed ones in the free [S2C@C(CN)2]2 dianion.21,29,30 The
molecules are arranged in the unit cell having the approximate
mirror plane of the i-MNT ligand almost normal to the c-axis and C -200
are stacked almost normal to this axis. These layers are held to- -100
gether mainly through H-bonds (Table 2), formed between the H
atoms of the two amino groups of a [Ni(i-MNT)(2a-5mt)2] mole- 0
I (μA)

100

200
Table 2
300
Hydrogen bonds for complex 1 (Å and °)

D–HA d(D–H) (Å) d(HA) (Å) d(DA) (Å) \(DHA) (°) 400

N(2)–H(2A)N(6)#1 0.88 2.16 3.010(3) 163.3 500


N(2)–H(2B)N(5)#2 0.88 2.19 3.057(2) 167.3 1.0 0.5 0.0 -0.5 -1.0 -1.5
N(4)–H(4A)N(5)#3 0.88 2.24 3.018(3) 146.9 E (V)
N(4)–H(4B)N(6)#4 0.88 2.17 3.009(2) 158.0
Figure 2. Cyclic voltammogram of 0.4 mM DMSO solution of (A) [Ni(i-MNT)(2a-
Symmetry transformations used to generate equivalent atoms: #1 x, y, z + 2, 5mt)2], 1, (B) [Ni(i-MNT)(2a-2tzn)2], 2 and (C) [Ni(i-MNT)(Im)2], 3. Scan rate = 100
#2 x + 1/2, y + 1/2, z + 5/2, #3 x + 1, y, z + 2, #4 x + 1/2, y + 1/2, z + 3/2. mV s1. Supporting electrolyte = TEAP, 0.1 M.
P. J. Cox et al. / Bioorg. Med. Chem. 17 (2009) 6054–6062 6057

of complex 1 shows the existence of two cathodic waves at Table 3


850 mV ([NiII]?[NiI]) and 1400 mV ([NiI]?[Ni0]) and three ano- Redox couples of the complexes and their cathodic and anodic potentials (in mV) in
DMSO and in 1/2 DMSO/buffer solution in the absence and presence of CT DNA
dic waves at 670 mV ([Ni0]?[NiI]), 290 mV ([NiI]?[NiII]) and
+770 mV ([NiII]?[NiIII]), while a third cathodic wave attributed to Ni(II)/Ni(I) Ni(I)/Ni(0) Ni(II)/Ni(III)
[NiIII]?[NiII] process appears at +20 mV. However, the current Epc1 Epa1 Epc2 Epa2 Epc3 Epa3
intensities attributed to the couple Ni(II)/Ni(I) are much higher 1 720a +565a 1360a +90a +850a +10a
than the others which are sometimes difficult to distinguish. Scan- 850b 290b 1400b 670b +770b +20b
ning in the range +1.0 to 1.0 V, the waves attributed to the pro- 825c 360c
I 0
cess ½Ni ¡½Ni  disappear once again (inset in Fig. 3A). +25d 70d

Quite similar is the redox behavior of complexes 2 and 3 2 850a +570a 1030a +195a +1260a +30a
(Fig. 2B–C and insets in Fig. 3B–C) and the corresponding redox 870b 410b n.d.b,e n.d.b,e +420b +40b
750c 485c
potentials are given in Table 3. Considering all the above processes
+120d 75d
3 930a +540a 1340a +130a +1430a +10a
705b 480b 1480b 640b n.d.b,e n.d.b,e
700c 510c
A -200
+5d 30d
a
-200 In 0.4 mM DMSO solution.
I (μA)

0 b
In 0.4 mM DMSO/buffer (1/2) solution (Efpc=a ).
c
In presence of CT DNA (Ebpc=a ) in 0.4 mM DMSO/buffer (1/2) solution.
-150 200 d
DEpc=a ¼ Ebpc=a  Efpc=a .
e
n.d. = not determined.
-100 400
I (μA)

1.0 0.5 0.0 -0.5 -1.0


E(V)
-50
and taking into account the literature,32–34 the complete pattern of
0
the redox behavior of complexes 1–3 could be described in the fol-
lowing equations:
50
0.0 -0.2 -0.4 -0.6 -0.8 -1.0
Epc1 Epc2
E (V)
[NiII] + e [NiI] [Ni0]
-400
B e
-200
Epa1 Epa3
I (μA)

-200 0 Epa2
[Ni0] [NiI] [NiII] [NiIII]
200

-100 400 +e +e +e
I (μA)

1.0 0.5 0.0 -0.5 -1.0


E (V)
Epc3
0 [NiIII] + e [NiII]

100
2.4. Interaction with DNA
0.0 -0.2 -0.4 -0.6 -0.8 -1.0
E (V) Transition metal complexes can bind to DNA via covalent and/
or noncovalent interactions.35,36 In covalent binding mode, a labile
C -400 ligand of the complexes is usually replaced by a nitrogen base of
-200 DNA such as guanine N7. On the other hand, the noncovalent
-200
DNA interactions include intercalative, electrostatic and groove
I (μA)

0
(surface) binding of metal complexes along outside of DNA helix,
200 along major or minor groove.36
-100
400
1.0 0.5 0.0 -0.5 -1.0
I (μA)

E (V)
2.4.1. Spectroscopic study of the DNA-binding
0 Electronic absorption spectroscopy is an effective method to
examine the binding mode of DNA with metal complexes.37 The
absorption spectra of the interaction of CT DNA with the com-
100 plexes have been recorded for a constant CT DNA concentration
in different [complex]/[DNA] mixing ratios (r) values and they
are shown in Figure 4. The changes observed in the absorption
0.0 -0.2 -0.4 -0.6 -0.8 -1.0
E (V) spectra of CT DNA in the presence of complexes 1–3, that is, the de-
crease of the intensity at kmax = 258 nm, which, for complexes 2
Figure 3. Cyclic voltammogram of 0.4 mM 1/2 DMSO/buffer (containing 150 mM and 3, is accompanied with a red-shift of the kmax up to 266 nm,
NaCl and 15 mM trisodium citrate at pH 7.0) solution of (A) [Ni(i-MNT)(2a-5mt)2] 1, indicate that the interaction with CT DNA results in the direct for-
(B) [Ni(i-MNT)(2a-2tzn)2] 2 and (C) [Ni(i-MNT)(Im)2], 3 in the absence (insets) or
presence of CT DNA in increasing amounts. The arrows show the changes upon
mation of a new complex with double-helical CT DNA,38 which
increasing amounts of CT DNA. Scan rate = 100 mV s1. Supporting electrolyte = simultaneously may cause the slight change of the conformation
buffer solution. of DNA39 in the case of complexes 2 and 3. The observed hypochro-
6058 P. J. Cox et al. / Bioorg. Med. Chem. 17 (2009) 6054–6062

A A

{[DNA]/(εa-εf)} x 10 , (M cm)
1.0 4

2
0.8 0.6 3

9
0.6 2
A

0.4
0.4 1
0 5 10 15 20 25
6

A
[DNA] x10 , (M)
0.2
0.2

0.0
250 300 350
λ (nm)
0.0
300 350 400 450 500
B 2.0 λ (nm)

{[DNA]/(εa-εf)} x10 , (M cm)


1.5 B

2
9
1.0 1.6
A

0.6
1.4
0.5 0.5
1.2
0.4
1.0
0.0 0 5 10 15 20 25
0.3
A
250 300 350 [DNA] x 10 , (M)
6

λ (nm)
0.2

C 2.0
0.1

0.0
1.5 300 350 400 450 500
λ (nm)

1.0
A

{[DNA]/(εa-εf)} x10 , (M cm)

C
2

2.4
1.5
0.5
9

2.0

0.0 1.0 1.6


250 300 350
λ (nm) 1.2
A

Figure 4. UV spectra of CT DNA in buffer solution (150 mM NaCl and 15 mM 5 10 15 20 25 30


0.5 6
trisodium citrate at pH 7.0) in the presence of (A) [Ni(i-MNT)(2a-5mt)2], 1 [DNA]x10 , (M)
([DNA] = 0.146 mM), (B) [Ni(i-MNT)(2a-2tzn)2], 2 ([DNA] = 0.164 mM) and (C)
[Ni(i-MNT)(Im)2], 3 ([DNA] = 0.264 mM) at diverse r values (r = [complex]/[DNA])
(r = 0–0.08). The arrows show the intensity changes upon increasing concentration
0.0
of complexes. 300 350 400 450 500
λ (nm)

mism is due to stacking interaction between the chromophores of Figure 5. UV spectra of (A) [Ni(i-MNT)(2a-5mt)2], 1, (B) [Ni(i-MNT)(2a-2tzn)2], 2
and (C) [Ni(i-MNT)(Im)2], 3 in DMSO solution in the presence of CT DNA at
the complexes and DNA base pairs probably consistent with the
increasing amounts. [Complex] = 40 lM, [CT DNA] = 0-40 lM. The arrows show the
intercalative binding mode.37,40 Additionally, the observed red- intensity changes upon increasing concentration of CT DNA. Insets: plots of ð½DNA
ea ef Þ
shift of 8 nm is an evidence of the stabilization of the CT DNA versus [DNA].
duplex.41
Figure 5 illustrates the spectral changes occurred in DMSO solu-
tion of complexes 1–3 during the titration upon addition of crease in the presence of CT DNA, resulting in hyperchromism
increasing amounts of CT DNA. The initial spectrum refers to the (Fig. 5A). The hyperchromic effect observed might be ascribed to
spectrum of a fixed concentration (4  105 M) of the free complex external contact (electrostatic binding)42 or that the complex could
in the absence of CT DNA. In the UV region, the intense absorption uncoil the helix structure of DNA and made more bases embedding
bands observed in the spectra of the complexes are attributed to in DNA exposed.43 On the other hand, the intensity of the band of 1
the intraligand p?p* transition of the coordinated groups of corre- centered at 455 nm has been found to initially decrease slightly in
sponding ligands.36 the presence of DNA, resulting in a slight hypochromism (Fig. 5A).
Even though no appreciable change in the position of the intral- Furthermore, two distinct isosbestic points at 445 nm and 465 nm
igand band of the complex is observed by addition of CT DNA, the exist in the spectrum and accompany the hypochromism as the
intensity of the band of 1 centered at 341 nm has been found to in- DNA concentration increases.
P. J. Cox et al. / Bioorg. Med. Chem. 17 (2009) 6054–6062 6059

For complex 2, the intensity of all the observed bands (293 nm, potentials are given in Table 3. No new redox peaks appeared after
343 nm, 373 nm and 455 nm) increase in the presence of CT DNA, the addition of CT DNA to each complex, but the current intensity
resulting in hyperchromism (Fig. 5B). Additionally, a blue-shift of of all the peaks decreased significantly, suggesting the existence of
5 nm is observed for bands centered at 293 nm (up to 288 nm) an interaction between each complex and CT DNA. The decrease in
and 373 nm (up to 368 nm) suggesting a conformational change current intensity can be explained in terms of an equilibrium mix-
of the DNA duplex.41 In the case of 3, hyperchormism is observed ture of free and DNA-bound complex to the electrode surface.52
for both the observed bands at 279 nm and 374 nm (Fig. 5C), while It can be observed (Table 3) that complexes 1–3 exhibit similar
for the latter a slight blue-shift of 3 nm (to 371 nm) is also present. electrochemical behavior upon addition of CT DNA at diverse r val-
The above experimental results derived from the UV titration ues. For increasing amounts of CT DNA, the cathodic potential Epc
experiments suggest that all the complexes can bind to CT DNA, shows a positive shift (DEpc = +25 mV for 1, DEpc = +120 mV for 2
maybe by electrostatic interaction.44 Nevertheless, the existing li- and DEpc = +5 mV for 3) while the anodic potential Epa shifts to
gands in combination with the formed chelate rings provide an ex- more negative values (DEpa = 70 mV for 1, DEpa = 75 mV for 2
tended aromatic moiety to the complexes so their simultaneous and DEpa = 30 mV for 3) ( Fig. 3). These shifts of the potentials
intercalative binding mode to the base pairs of DNA may not be ru- show that complexes 1, 2 and 3 can bind to DNA by both interca-
led out.45 lation and electrostatic interaction.33,34,50,51
The binding strength of the complexes with CT DNA is mirrored
in the intrinsic binding constant Kb, which represents the binding 2.4.3. Competitive DNA-binding studies with EB
constant per DNA base pair and can be obtained by monitoring A competitive ethidium bromide (EB) binding study has been
the changes in the absorbance at 341 nm for 1, 343 nm for 2 and undertaken with fluorescence spectroscopic titration46,47,51 in or-
374 nm for 3, with increasing concentrations of CT DNA, according der to investigate if the complexes could displace EB from its EB-
to the following equation (Eq. (1)):44 DNA complex. EB (=3,8-diamino-5-ethyl-6-phenylphenanthridini-
um bromide), a phenenthridine fluorescence dye, is a typical indi-
½DNA ½DNA 1
¼ þ ð1Þ cator of intercalation,53 forms soluble complexes with nucleic acids
ðea  ef Þ ðeb  ef Þ K b ðeb  ef Þ
and emits intense fluorescence in the presence of CT DNA due to
where [DNA] is the concentration of DNA in base pairs, the apparent the intercalation of the planar phenenthridinium ring between
absorption coefficients ea, ef and eb correspond to Aobsd/[complex], the adjacent base pairs on the double helix.54
extinction coefficient for the free complex and the extinction No fluorescence has been observed for the complexes at room
coefficient for the complex in the fully bound form, respectively. In temperature in solution or in the presence of CT DNA and the bind-
plots ð½DNA ing of the complexes and DNA cannot be directly predicted through
ea e Þ versus [DNA], Kb is given by the ratio of the slope to the y
f
the emission spectra. Thus, competitive EB binding studies have
intercept (insets in Fig. 5). The determined Kb values for complexes
been performed in order to gain support for the extent of binding
1–3 are given in Table 4. These values suggest a relatively strong bind-
of each complex with DNA. EB does not show any appreciable
ing of the complexes to CT DNA and are lower than the EB binding
emission in buffer solution due to fluorescence quenching of the
affinity for CT DNA, (Kb = 1.23 ± 0.07  105 M1 bp1),46 suggesting
free EB by the solvent molecules55 while in the presence of CT
that electrostatic and/or intercalative interaction may affect EB
DNA, the fluorescence intensity of EB is highly enhanced due to
partial displacement.33,34,46,47 The intrinsic binding constant (Kb)
its strong intercalation between the adjacent DNA base pairs.56
obtained for complex 3 (=4.94 ± 0.05  104 M1) is higher than the
Addition of a second molecule, which can also bind to DNA, can de-
other two complexes indicating its stronger ability to bind to CT
crease the DNA-induced EB emission.57 Two mechanisms have
DNA.38
been proposed to account for this reduction in the emission inten-
sity: the replacement of molecular fluorophores (EB in this case)
2.4.2. Electrochemical study of the DNA-binding
and/or electron transfer.58
The electrochemical investigations of metal–DNA interactions
The emission spectra of EB bound to CT DNA in the absence and
can provide a useful complement to spectroscopic methods, for
presence of each complex have been recorded for [EB] = 2
example, for non absorbing species, and yield information about
 105 M, [DNA] = 2.6  105 M and increasing amounts of each
interactions with both the reduced and oxidized form of the me-
complex. The emission band at 592 nm of the DNA-EB system
tal.48 The electrochemical potential of a small molecule will shift
decreased in intensity upon addition of each complex at diverse r
positively when it intercalates into DNA double helix, while a neg-
values. This decrease of EB fluorescence (up to 20% of the initial
ative shift of the potential is observed for an electrostatic interac-
EB-DNA fluorescence intensity) (Fig. 6A) indicates the competition
tion to DNA.49,50 Additionally, in the case of more potentials than
of the complexes with EB in binding to DNA. The quenching of
one, a positive shift of Ep1 and a negative shift of Ep2 could imply
DNA–EB fluorescence for complexes 1–3 suggests that they are
that the molecule can bind to DNA by both intercalation and elec-
able to displace EB from the DNA–EB complex and they can inter-
trostatic interaction.38,51
act with CT DNA probably by the intercalative mode.47,59
In Figure 3, the cyclic voltammograms of the complexes in the
The quenching efficiency for each complex is evaluated by the
presence of CT DNA in diverse r values are shown. The quasi-
Stern–Volmer constant KSV, which varies with the experimental
reversible redox couple Ni(II)/N(I) for each complex in 1/2
conditions according to Eq. (2):
DMSO/buffer solution has been studied upon addition of CT DNA
and the shifts of the corresponding cathodic Epc and anodic Epa Io
¼ 1 þ K SV ½complex ð2Þ
I
where Io and I are the emission intensities in the absence and the
Table 4 presence of the complex, respectively. Figure 6B shows the Stern–
The intrinsic binding constants (Kb) of complexes 1–3 with CT DNA and the Stern–
Volmer plots for the complexes and the KSV values calculated for
Volmer constants (KSV)
the complexes are shown in Table 4. The Stern–Volmer plot of
Complex Kb (M1) KSV (M1) DNA–EB illustrates that the quenching of EB bound to DNA by each
4
1 2.94(±0.28)  10 5.39(±0.32)  105 complex is in good agreement (R = 0.99) with the linear Stern–Vol-
2 1.68(±0.20)  104 4.83(±0.29)  105 mer equation (Eq. (2)), which proves that the partial replacement of
3 4.94(±0.05)  104 4.29(±0.28)  105
EB bound to DNA by each complex results in a decrease in the fluo-
6060 P. J. Cox et al. / Bioorg. Med. Chem. 17 (2009) 6054–6062

A 100 4. Experimental

80 4.1. Materials and instrumentation


1
% EB fluorescence

2
60 4.1.1. Materials
3
All chemicals were purchased from Aldrich, EDH Chemicals and
40 Merck and were used as received. CT DNA and EB were purchased
from Sigma, NaCl and all solvents were purchased from Merck, tri-
20 sodium citrate was purchased from Riedel-de Haen. All the chem-
icals and solvents were reagent grade and were used as purchased.
0 Tetraethylammonium perchlorate (TEAP) was purchased from Car-
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 lo Erba and, prior to its use, it was recrystallized twice from ethanol
r and dried under vacuum.
DNA stock solution was prepared by dilution of CT DNA to buffer
B 6 (containing 150 mM NaCl and 15 mM trisodium citrate at pH 7.0)
followed by exhaustive stirring at 4 °C for three days, and kept at
5 4 °C for no longer than a week. The stock solution of CT DNA gave
a ratio of UV absorbance at 260 and 280 nm (A260/A280) of 1.89, indi-
4 cating that the DNA was sufficiently free of protein contamination.60
The DNA concentration was determined by the UV absorbance at
Io/I

3 1 260 nm after 1:20 dilution using e = 6600 M1 cm1.61


2
2 3
4.1.2. Instrumentation—physical measurements
1 Elemental analysis was performed on a Perkin–Elmer 240B ele-
mental analysis instrument, whereas Ni analyses were performed
0 2 4 6 8 10
6 with a Perkin–Elmer model 403 Atomic Absorption Spectrometer.
[Complex] x 10 , (M)
Molar conductivities were measured on a WTW conductivity bridge
Figure 6. (A) Plot of EB relative fluorescence intensity (%) versus r (=[complex]/ employing a calibrated dip type cell and 103 M solutions in DMSO.
[DNA]) for complexes 1–3 in buffer solution (150 mM NaCl and 15 mM trisodium Magnetic susceptibility measurements of powdered samples were
citrate at pH 7.0). (B) Stern–Volmer quenching plot of EB bound to CT-DNA by performed at room temperature (25 °C) utilizing a Johnson–Mat-
complexes 1–3. [EB] = 20 lM, [DNA] = 26 lM, kem = 592 nm.
they balance, calibrated with Hg[Co(NCS)4]. Infrared spectra were
recorded throughout the 4000–225 cm1 region, with samples pre-
pared as KBr or polyethylene discs, with a Perkin–Elmer FT-IR Spec-
rescence intensity.52 The high KSV values of the complexes show trum One Spectrophotometer equipped with a CsI beam splitter.
that they can be bound very tightly to the DNA.33,34,45–47 Among UV–vis (UV–vis) spectra were recorded in DMSO solution at con-
the three complexes, 1 has the highest Ksv suggesting its stronger centrations in the range 105–103 M on a Hitachi U-2001 dual
ability to displace EB from the EB–DNA complex than the other beam spectrophotometer. Fluorescence spectra were recorded in
two complexes. solution on a Hitachi F-7000 fluorescence spectrophotometer.
Cyclic voltammetry studies were performed on an Eco chemie
3. Conclusions Autolab Electrochemical analyzer. Cyclic voltammetric experi-
ments were carried out in a 30 mL three-electrode electrolytic cell.
The mixed-ligands Ni(II) complexes afforded by the reaction of The working electrode was platinum disk, a separate Pt single-
[Ni(L)4Cl2] with dianion of 1,1-dicyano-2,2-ethylenedithiolate li- sheet electrode was used as the counter electrode and a Ag/AgCl
gand (i-MNT). The crystal and molecular structures of the studied electrode saturated with KCl was used as the reference electrode.
Ni(II) compounds reveal the same distorted square-planar geome- The cyclic voltammograms of the complexes were recorded in
try for all complexes. The (i-MNT) ligand is coordinated with the 0.4 mM DMSO solutions and in 0.4 mM 1/2 DMSO/buffer solutions
two sulfur atoms and the endocyclic nitrogen atom of the two at m = 100 mV s1 where TEAP and the buffer solution were the
2a-5mt or 2a-2tzn or Im ligands. The isolated neutral, diamagnetic supporting electrolytes, respectively. Oxygen was removed by
and of square-planar geometry complexes are of interest, since purging the solutions with pure nitrogen previously saturated with
they meet the necessary properties for various biological tests. In solvent vapors. All electrochemical measurements were performed
the cyclic voltammograms of the complexes recorded in DMSO at 25.0 ± 0.2 °C.
solution and in 1/2 DMSO/buffer solution, the expected waves
attributed to redox couples and the corresponding potentials, char- 4.2. Synthesis of the dithiolate ligands
acteristic for Ni(II) complexes have been recorded.
The study of the interaction of the complexes with CT DNA has The disodium salt of 1,1-dicyano-2,2-ethylenedithiolate trihy-
been performed with UV spectroscopy and cyclic voltammetry and drate Na2(i-MNT)3H2O (Scheme 1) was prepared by established
has revealed that the complexes can bind to DNA. UV spectroscopic procedures.62–64
titrations have been used in order to calculate the binding strength
of the complexes with CT DNA which is mirrored in the intrinsic
binding constant Kb. Cyclic voltammetric studies have shown that NaS C N
complexes 1–3 can bind to CT DNA by both intercalation and elec- C C 3H2O
trostatic interaction. Competitive binding studies with EB with NaS C N
fluorescence spectroscopy have shown that the interaction be-
Na2(i–MNT)·3H2O
tween DNA–EB complex and the complexes can release EB from
its DNA complex, indicating that they can bind to DNA probably Scheme 1. Structure of disodium salt of 1,1-dicyano-2,2-ethylenedithiolate trihy-
via the intercalative mode. drate ligand.
P. J. Cox et al. / Bioorg. Med. Chem. 17 (2009) 6054–6062 6061

N N Table 5
N
Crystal data and structure refinement for complex 1
H3C NH2 NH2
S S NH Molecular formula C12H12N6NiS4
Formula weight 427.23
2a–5mt 2a–2tzn Im Temperature (K) 120(2)
Wavelength (Å) 0.71073
Scheme 2. Structural formulae of 2-amino-5-metyl-thiazole, 2-amino-2-thiazoline Crystal system Monoclinic
and Imidazole. Space group P21/n
a (Å) 8.1524(2)
4.3. Synthesis of the starting compounds b (Å) 14.0906(3)
c (Å) 15.5424(2)
a (°) 90
The starting compounds used [Ni(L)4Cl2] were prepared by the b (°) 95.3390(10)
reaction of NiCl26H2O with L = 2a-5mt, 2a-2tzn and Im in 1:4 mo- c 90
lar ratio.65 The formulae of the 2-amino-5-methyl-thiazole, 2-ami- Volume (Å3) 1777.64(6)
no-thiazoline and imidazole are given in Scheme 2. Z 4
Density (calcd) (Mg/m3) 1.596
The mixed-ligand chelates of the general type [Ni(i-MNT)(L)2]
Absorption coefficient (mm1) 1.566
were prepared by treating the starting compounds [Ni(L)4Cl2] with F(0 0 0) 872
dithiolate i-MNT in 1:1 molar ratio in methanol.19 A representative Crystal size (mm3) 0.10  0.04  0.02
reaction of the synthesis of [Ni(i-MNT)(2a-5mt)2] is given in h range for data collection (°) 3.00–27.48
Index ranges 10 6 h 6 10, 18 6 k 6 18,
Scheme 3.
20 6 l 6 20
Reflections collected 27,062
4.4. Synthesis of [Ni(i-MNT)(2a-5mt)2], 1 Independent reflections 4069 [R(int) = 0.0482]
Completeness to h = 27.48° 99.8%
A solution of [Ni(2a-5mt)4Cl2] (0.699 g, 1 mmol) in 20 mL MeOH Max. and min. transmission 0.9693 and 0.8591
Refinement method Full-matrix least-squares on F2
was mixed slowly with a solution of Na2(i-MNT) (0.24 g, 1 mmol) in
Data/restraints/parameters 4069/0/210
10 mL MeOH. After 1 h of stirring the resulting dark-brown solid was Goodness-of-fit on F2 1.017
collected by filtration, washed with methanol three times and ether Final R indices [I > 2r(I)] R1 = 0.0278, wR2 = 0.0571
and dried in air. Weight: 0.304 g. Yield: 72%. Slow evaporation of R indices (all data) R1 = 0.0428, wR2 = 0.0612
Final weighting scheme calcd w ¼ 1=½r2 ðF 2o Þ þ ð0:0250PÞ2
[Ni(i-MNT)(2a-5mt)2] solution in CH3CN/CH2Cl2 at low temperature
þ0:7523P, where P ¼ ðF 2o þ 2F 2c Þ=3
(4 °C) yielded brown crystals suitable for X-ray diffraction analysis. Largest diff. peak and hole (e Å3) 0.290 and 0.299
Anal. Calcd for C12H12N6S4Ni: C, 33.74; H, 2.83; N, 19.67; Ni, 13.74.
Found: C, 33.68; H, 2.83; N, 19.65; Ni, 13.68. IR: mmax/cm1;
m(C„N): 2211(vs); m(C@CS2): 1431(vs); mas(CS2): 941(m); ms(CS2):
813(m); m(Ni–N): 478(w); m(Ni–S): 392(w) (KBr disk); UV–vis: C10H8N6S2Ni: C, 35.85; H, 2.41; N, 25.08; Ni, 17.52. Found: C,
kmax/nm (log e) (in DMSO): 622 (1.64), 457 (4.10), 340 (4.40). 35.81; H, 2.44; N, 24.95; Ni, 17.49. IR: mmax/cm1; m(C„N):
2228(s); m(C@C): 1376(vs); mas(CS2): 916(w); ms(CS2): 898(w);
4.5. Synthesis of [Ni(i-MNT)(2a-2tzn)2], 2 m(Ni–N): 486(w); m(Ni–S): 387(w) (KBr disk); UV–vis: kmax/nm
(log e) (in DMSO): 634 (1.25), 456 (4.12), 340 (4.41).
Complex 2 was obtained by slow addition of Na2(i-MNT)3H2O
(0.24 g, 1 mmol) in MeOH (20 mL) to a solution of [Ni(2a-2tzn)4Cl2] 4.7. DNA-binding studies
(0.538 g, 1 mmol) in MeOH (20 mL). The brown solid was filtered,
washed successively with 5 mL of MeOH and Tt2J and dried in The interaction of complexes 1–3 with CT DNA has been studied
air. Weight: 0.26 g, Yield: 65%. Anal. Calcd for C10H12N6S4Ni: C, with UV spectroscopy and CV in order to investigate the possible
29.79; H, 3.00; N, 20.84; Ni, 14.56. Found: C, 29.81; H, 2.91; N, binding modes to CT DNA and to calculate the binding constants
20.80; Ni, 14.35%. IR: mmax/cm1; m(C„N): 2203(vs); m(C@CS2): to CT DNA (Kb). In UV titration experiments, the spectra of CT
1434(vs); mas(CS2): 936(w); ms(CS2): 901(w); m(Ni–N): 481(w); DNA in the presence of each complex have been recorded for a con-
m(Ni–S): 396(w) (KBr disk); UV–vis: kmax/nm (log e) (in DMSO): stant CT DNA concentration in diverse [complex]/[CT DNA] mixing
625 (1.57), 457 (4.26), 341 (4.53). ratios (r). The intrinsic binding constants, Kb, of the complexes with
CT DNA have been determined using the UV spectra of the com-
4.6. Synthesis of the complex [Ni(i-MNT)(Im)2], 3 plexes recorded for a constant concentration in the absence or
presence of CT DNA for diverse r values. Control experiments with
Complex 3 was prepared by gradual addition of Na2(i- DMSO were performed and no influence of the spectra was
MNT)3H2O (0.24 g, 1 mmol) in MeOH (20 mL) to a solution of observed.
[Ni(Im)4Cl2] (0.40 g, 1 mmol) in MeOH (20 mL). The brown solid The interaction of the complexes with CT DNA has been also
was isolated by filtration, washed with 5 mL of MeOH and Tt2J investigated by monitoring the changes observed in the cyclic vol-
and dried in air. Weight: 0.20 g, Yield: 60%. Anal. Calcd for tammogram of a 0.40 mM 1/2 DMSO/buffer solution of the com-

CH3
S
H 2N
N S C N
[Ni(2a-5mt)4Cl2] + Na2(i-MNT).3H2O Ni C C
N S C N

H3C + 2NaCl + 2a-5mt +3H2O


S NH2

Scheme 3. Representative reaction of the synthesis of [Ni(i-MNT)(2a-5mt)2].


6062 P. J. Cox et al. / Bioorg. Med. Chem. 17 (2009) 6054–6062

plexes upon addition of CT DNA at diverse r values. The buffer was 14. Chaviara, A. T.; Cox, P. J.; Repana, K. H.; Pantazaki, A. A.; Papazisis, K. T.;
Kortsaris, A. H.; Kyriakidis, D. A.; Nikolov, G. S.; Bolos, C. A. J. Inorg. Biochem.
also used as the supporting electrolyte and the cyclic voltammo-
2005, 99, 467.
grams were recorded at m = 100 mV s1. 15. Goldin, A.; Sofina, Z.; Syrum, A. Natl. Cancer Inst. Monograph 1980, 55, 25.
The competitive studies of each complex with EB have been 16. Ciccarelli, R. B.; Wetterhahn, K. E. Cancer Res. 1984, 44, 3892.
investigated with fluorescence spectroscopy in order to examine 17. Misra, M.; Olinski, R.; Dizdaroglu, M.; Kasprzak, K. S. Chem. Res. Toxicol. 1993, 6,
33.
whether it is able to displace EB from its CT DNA–EB complex. 18. Nieboer, E.; Tom, R. T.; Rossetto, F. E. Biol. Trace Elem. Res. 1989, 21, 23.
The CT DNA–EB complex was prepared by adding 20 lM EB and 19. Alkam, H.; Hatzidimitriou, A.; Hadjikostas, C. C.; Tsiamis, C. Inorg. Chim. Acta
26 lM CT DNA in buffer (150 mM NaCl and 15 mM trisodium cit- 1997, 256, 41.
20. Hadjikostas, C. C.; Alkam, H. H.; Bolos, C. A.; Christidis, P. C. Polyhedron 2001,
rate at pH 7.0). The intercalating effect of complexes 1–3 with 20, 395.
the DNA–EB complex was studied by adding a certain amount of 21. Nakamoto, K. Infrared and Raman spectra of Inorganic and Coordination
a solution of the compound step by step into the solution of the Compounds, 4th ed.; J. Wiley & Sons, Inc.: NY, 1986.
22. Lever, A. B. P. Inorganic Electronic Spectroscopy, 2nd ed.; Elsevier Science
DNA–EB complex. The influence of the addition of each compound Publishers B.V.: NY, 1984.
to the DNA–EB complex solution has been obtained by recording 23. Macias, B.; Villa, M. V.; Chicote, E.; Martin-Velasco, S.; Castineiras, A. J.; Borras,
the variation of fluorescence emission spectra. A. Polyhedron 2002, 21, 1899.
24. Jeary, W. J. Coord. Chem. Rev. 1971, 7, 81.
25. Coucouvanis, D.; Hollander, F. J.; Caffery, M. I. Inorg. Chem. 1976, 15, 1853.
4.8. Crystal structure determination of [Ni(i-MNT)(2a-5mt)2] 26. Alkam, H. H.; Kanan, K. M.; Hadjikostas, C. C. Polyhedron 2005, 24, 2944.
27. Liu, B. X.; Chen, G. H.; Lin, Y. Y.; Yu, Y. P. Acta Crystallogr., Sect. E 2007, 63,
m1971.
Crystal data, data collection and refinement details are given in
28. Zhang, L. J.; Liu, B. X.; Ge, H. Q.; Xu, D. J. Acta Crystallogr., Sect. E 2006, 62,
Table 5. X-ray intensity data were collected at 120 K on a Bruker- m2180.
Nonius diffractometer graphite monochromated MoKa radiation. 29. Hummel, H. U. Acta Crystallogr., Sect. C 1985, 41, 1591.
A total of 27062 reflections were measured (3.00° < h < 27.48°) 30. Hummel, H. U. Acta Crystallogr., Sect. C 1987, 43, 41.
31. Dendrinou-Samara, C.; Psomas, G.; Raptopoulou, C. P.; Kessissoglou, D. P. J.
which were further merged to 4069 independent reflections Inorg. Biochem. 2001, 83, 7.
(Rint = 0.0482), of which 3304 were considered as observed 32. Thirumavalavan, M.; Akilan, P.; Kandaswamy, M. Polyhedron 2005, 24, 1781.
[I > 2r(I)]. Multi-scan absorption corrections were made with the 33. Psomas, G. J. Inorg. Biochem. 2008, 102, 1798.
34. Skyrianou, K. C.; Raptopoulou, C. P.; Psycharis, V.; Kessissoglou, D. P.; Psomas,
program SADABS.66 The structure was solved with direct methods G. Polyhedron, 2009, 28, in press, (doi: 10.1016/j.poly.2009.04.002)
using the program67 SIR97 and refined by full-matrix least-squares 35. Zhang, Q. L.; Liu, J. G.; Chao, H.; Xue, G. Q.; Ji, L. N. J. Inorg. Biochem. 2001, 83, 49.
refinement with the program SHELXL97,68 while geometric calcula- 36. Zhang, Q. L.; Liu, J. G.; Liu, J.; Xue, G. Q.; Li, H.; Liu, J. Z.; Zhou, H.; Qu, L. H.; Ji, L.
N. J. Inorg. Biochem. 2001, 85, 291.
tions were performed with the program PLATON.69 37. Kelly, T. M.; Tossi, A. B.; McConnell, D. J.; Strekas, T. C. Nucleic Acids Res. 1985,
13, 6017.
Acknowledgements 38. Son, G. S.; Yeo, J. A.; Kim, M. S.; Kim, S. K.; Holmen, A.; Akerman, B.; Norden, B. J.
Am. Chem. Soc. 1998, 120, 6451.
39. Song, Y. M.; Wu, Q.; Yang, P. J.; Luan, N. N.; Wang, L. F.; Liu, Y. M. J. Inorg.
We thank Professor P.C. Christidis for his valuable discussion. Biochem. 2006, 100, 1685.
40. Pyle, A. M.; Rehmann, J. P.; Meshoyrer, R.; Kumar, C. V.; Turro, N. J.; Barton, J. K.
J. Am. Chem. Soc. 1989, 111, 3053.
Supplementary data 41. Long, E. C.; Barton, J. K. Acc. Chem. Res. 1990, 23, 271.
42. Pasternack, R. F.; Gibbs, E. J.; Villafranca, J. J. Biochemistry 1983, 22, 2406.
Complete lists of atomic coordinates, anisotropic displacement 43. Pratviel, G.; Bernadou, J.; Meunier, B. Adv. Inorg. Chem. 1998, 45, 251.
44. Janeso, A.; Nagy, L.; Moldrheim, E.; Sletten, E. J. Chem. Soc., Dalton Trans. 1999,
parameters, bond lengths and angles, hydrogen bonds have been 1587.
deposited at the Cambridge Crystallographic Data Centre [CCDC 45. Chauhan, M.; Banerjee, K.; Arjmand, F. Inorg. Chem. 2007, 46, 3072.
No. 716372], 12 Union Road, Cambridge, CB2 1EZ, UK (fax: +44 46. Dimitrakopoulou, A.; Dendrinou-Samara, C.; Pantazaki, A. A.; Alexiou, M.;
Nordlander, E.; Kessissoglou, D. P. J. Inorg. Biochem. 2008, 102, 618.
123 336033; e-mail: deposit@ccdc.cam.ac.uk or www://www.ccdc. 47. Tarushi, A.; Psomas, G.; Raptopoulou, C. P.; Kessissoglou, D. P. J. Inorg. Biochem.
cam.ac.uk). Supplementary data associated with this article can be 2009, 103, 898.
found, in the online version, at doi:10.1016/j.bmc.2009.06.058. 48. Carter, M. T.; Bard, A. J. J. Am. Chem. Soc. 1987, 109, 7528.
49. Zhang, S. S.; Niu, S. Y.; Qu, B.; Jie, G. F.; Xu, H.; Ding, C. F. J. Inorg. Biochem. 2005,
99, 2340.
References and notes 50. Carter, M. T.; Rodriguez, M.; Bard, A. J. J. Am. Chem. Soc. 1989, 111, 8901.
51. Jiao, K.; Wang, Q. X.; Sun, W.; Jian, F. F. J. Inorg. Biochem. 2005, 99, 1369.
1. Wang, D.; Lippard, S. J. Nat. Rev. 2005, 4, 307. 52. Tabassum, S.; Parveen, S.; Arjmand, F. Acta Biomater. 2005, 1, 677.
2. Zhong, X.; Wei, H.-L.; Liu, W.-S.; Wang, D.-Q.; Wanga, X. Bioorg. Med. Chem. Lett. 53. Wilson, W. D.; Ratmeyer, L.; Zhao, M.; Strekowski, L.; Boykin, D. Biochemistry
2007, 17, 3774. 1993, 32, 4098.
3. Juris, A.; Balzani, F.; Barigelletti, S.; Campagna, S.; Belser, P.; Von Zelewsky, A. 54. Reihardt, C. G.; Krugh, T. R. Biochemistry 1978, 17, 4845.
Coord. Chem. Rev. 1988, 84, 85. 55. Dhar, S.; Nethaji, M.; Chakravarty, A. R. J. Inorg. Biochem. 2005, 99, 805.
4. Halcrow, M. A.; Christou, G. Chem. Rev. 1994, 94, 2421. 56. Olmsted, J.; Kearns, D. R. Biochemistry 1977, 16, 3647.
5. Buschini, A.; Pinelli, S.; Pellakani, C.; Giordani, F.; Belicchi, F.; Bisceglie, F.; 57. Baguley, B. C.; Lebret, M. Biochemistry 1984, 23, 937.
Giannetto, M.; Pelosi, G.; Tarasconi, P. J. Inorg. Biochem. 2009, 103, 666. 58. Pasternack, R. F.; Cacca, M.; Keogh, B.; Stephenson, T. A.; Williams, A. P.; Gibbs,
6. Barone, G.; Cambino, N.; Ruggirello, A.; Silvestri, A.; Terenzi, A.; Turco Liveri, V. F. J. J. Am. Chem. Soc. 1991, 113, 6835.
J. Inorg. Biochem. 2009, 103, 731. 59. Novakova, O.; Chen, H.; Vrana, O.; Rodger, A.; Sadler, P. J.; Brabec, V.
7. Bolos, C. A.; Nikolov, G. S.; Ekateriniadou, L.; Kortsaris, A.; Kyriakidis, D. A. Biochemistry 2003, 42, 11544.
Metal-Based Drugs 1998, 5, 323. 60. Marmur, J. J. Mol. Biol. 1961, 3, 208.
8. Bolos, C. A.; Fanourgakis, P. V.; Christidis, P. C.; Nikolov, G. S. Polyhedron 1999, 61. Reichmann, M. F.; Rice, S. A.; Thomas, C. A.; Doty, P. J. Am. Chem. Soc. 1954, 76,
18, 1661. 3047.
9. Chaviara, A. T.; Kioseoglou, E. E.; Pantazaki, A. A.; Tsipis, A. C.; Karipidis, P. A.; 62. Gompper, R.; Topfl, W. Chem. Ber. 1962, 95, 2851.
Kyriakidis, D. A.; Bolos, C. A. J. Inorg. Biochem. 2008, 102, 1749. 63. Werden, B. G.; Billig, E.; Gray, H. B. Inorg. Chem. 1966, 5, 78.
10. Bolos, C. A.; Chaviara, A. T.; Mourelatos, D.; Iakovidou, Z.; Mioglou, E.; 64. Jensen, K. A.; Hendriksen, L. Acta Chem. Scand. 1968, 22, 1107.
Chrysogelou, E.; Papageorgiou, A. Bioorg. Med. Chem. 2009, 17, 3142. 65. Raper, E. S. Coord. Chem. Rev. 1994, 129, 91.
11. Chaviara, A. T.; Cox, P. J.; Repana, K. H.; Papi, R. M.; Papazisis, K. T.; Zambouli, 66. Blessing, R. H. J. Appl. Crystallogr. 1997, 30, 421.
D.; Kortsaris, A. H.; Kyriakidis, D. A.; Bolos, C. A. J. Inorg. Biochem. 2004, 98, 67. Altomare, A.; Burla, M. C.; Camali, M.; Casscarano, G. L.; Giacovazzo, C.;
1271. Guagliardi, A.; Moliterni, A. G. G.; Polidori, G.; Spagna, R. J. Appl. Crystallogr.
12. Chaviara, A. T.; Christidis, P. C.; Papageorgiou, A.; Chrysogelou, E.; Hadjipavlou- 1999, 32, 115.
Litina, D. J.; Bolos, C. A. J. Inorg. Biochem. 2005, 99, 2102. 68. Sheldrick, G. M. SHELXL-97 Program for the Refinement of Crystal Structures;
13. Bolos, C. A.; Papazisis, K. T.; Zambouli, D.; Voyatzi, S.; Kortsaris, A. H.; University of Göttingen: Germany, 1997.
Kyriakidis, D. J. Inorg. Biochem. 2002, 88, 25. 69. Spek, A. I. J. Appl. Crystallogr. 2003, 36, 7.

Vous aimerez peut-être aussi