You are on page 1of 19

Asian Journal of Biochemical and Pharmaceutical Research Issue 2 (Vol.

1) 2011

ISSN: 2231-2560
Research Article

Asian Journal of Biochemical and Pharmaceutical Research


Synthesis, Spectroscopic Characterization, DNA Cleavage And Antimicrobial Activity
Of Binuclear Copper(II), Nickel(II) And Oxovanadium(Iv) Schiff Base Complexes
N. Mahalakshmi and R.Rajavel*
Department of Chemistry, Periyar University, Salem, Tamilnadu, India.

Received: 3 May 2011; Revised: 23May 2011; Accepted: 28May. 2011

Abstract: A new series of transition metal complexes of Cu(II), Ni(II) and VO(II) have been synthesized from
the Schiff base (L) derived from 2-carboxybenzaldehyde and 3,3,4,4-tetraminobiphenyl. Structural features
were obtained from their elemental analyses, IR, magnetic susceptibility, molar conductance, UVVis and ESR
spectral studies. The data show that these complexes have composition of [M2(L)]X type. (Where M = (Cu(II),
Ni(II), and VO(II) X=ClO4-, SO42- L = binucleating tetradendate ligand). The UVVis, ESR, magnetic
susceptibility and spectral data of the complexes suggest a squareplanar geometry around the central metal ion
except VO(II) complex which has squarepyramidal geometry. The redox behavior of copper, Nickel and
vanadyl complexes was studied by cyclic voltammetry. The interaction study of the copper complex with CTDNA was carried out using cyclic voltammetry. The pUC18 DNA cleavage study was monitored by gel
electrophoresis method. The results suggest that binuclear Cu(II), Ni(II) and VO(II) complexes cleaves pUC18
DNA in presence of the oxidant H2O2. The invitro antimicrobial activities of the synthesized compounds have
been tested against the gram negative and gram positive bacterias. The binuclear Schiff base complexes were
found to be higher antibacterial activity than the free ligand.
Keywords: Schiff base, binuclear, pUC18 DNA cleavage, Antimicrobial, CT-DNA, 2-carboxybenzaldehyde.

INTRODUCTION:
Schiff bases and their metal complexes play an important role in the development of coordination
chemistry, resulting in an enormous number of publications, ranging from pure synthetic work to
physicochemical [1] and biochemical relevant studies of metal complexes and found wide range of
applications [2-6]. Mokhles M. Abd-Elzaher has been studied by spectroscopic characterization of
some tetradentate Schiff Bases and their complexes with Nickel, Copper and Zinc [7]. Schiff bases
derived from an amine and any aldehyde are a class of compounds which co-ordinate to metal ions via
the azomethine nitrogen [8]. Metal complexes of Schiff bases derived from substituted
salicylaldehydes and various amines have been widely investigated because of their wide applicability
[9-12].Chelating ligands containing O and N donor atoms show broad biological activity and are of
special interest because of the variety of ways in which they are bonded to metal ions [13]. It is well
known that several Schiff base complexes have anti-inflammatory, antipyretic, analgesic, anti-diabetic,
anti-bacterial, anti-cancer and anti-HIV activity [14, 15]. The interaction of transition metal complexes
with DNA has been extensively studied in the development of new tools for nanotechnology [16, 17].
In the present investigation we report here the synthesis, spectroscopic, redox, DNA cleavage studies,
antimicrobial and DNA cleavage studies and analytical characterizations of new tetradentate N2O2
525

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

donor type Schiff base and their metal complexes. Our general strategy for preparing the desired
ligand was based on the condensation of a suitable aldehyde precursor with aminobiphenyl. In this
paper the synthesis of new tetradentate N2O2 donor type Schiff base and its metal complexes (Cu(II),
Ni(II) and VO(II)) derived by the condensation of 3,3,4,4-tetraminobiphenyl and 2carboxybenzaldehyde is described.
EXPERIMENTAL:
All the chemicals used were chemically pure and AR grade. Solvents were purified and dried
according to standard procedures [18]. Metals were purchased from Merck. 3,3,4,4tetraminobiphenyl and 2- carboxybenzaldehyde were obtained from Aldrich. Other chemicals were
also purchased from Merck and Aldrich.
Synthesis: Caution! Perchlorate salts are potentially explosive and were handled only in small
quantities with care.
Physical measurements
The elemental analysis were carried out with a Carlo-Erba 1106-model 240 Perkin Elmer
analyzer. The solution conductivity measurements were performed to establish the charge type of the
complexes. The complexes were dissolved in MeCN/DMF/DMSO and molar conductivities of 10-3M
of their solutions at 29 0C were measured. Infrared spectra were recorded on the Perkin Elmer FT-IR8300 model spectrometer using KBr disc and Nujol mull techniques in the range of 4000-400 cm-1.
Electronic absorption spectra in the UV-Visible range were recorded on Perkin Elmer Lambda -25
between 200-800 nm by using DMF as the solvent Magnetic susceptibility data were collected on
powdered sample of the compounds at room temperature with PAR155 vibrating sample
magnetometer. EPR spectra were recorded on a Varian JEOL-JES-TE100 ESR spectrophotometer at
X-band microwave frequencies for powdered samples at room temperature. Cyclic voltammetry
studies were performed on a CHI760C electrochemical analyzer in single compartmental cells at 29 0C
with H2O/DMSO (95:5) solution using tetrabutylammonium perchlorate (TBAP) as a supporting
electrolyte.
Preparation of binucleating tetradentate Schiff base ligand
The binucleating tetradentate Schiff base was prepared by condensation of tetramine with
appropriate aldehydes (Scheme 1). 3,3,4,4-tetraminobiphenyl (0.214 g; 1 mmol) in 20 ml of
methanol was stirred with 2 carboxybenzaldehyde (0.600 g; 4 mmol) in 20 ml of methanol for 4 h.The
resulting yellow solid was separated and dried in vacuum. Yield: 80%.
Synthesis of binuclear Schiff base complexes
[Cu2(L)]4ClO4 (1)
[Ni2(L)]4ClO4 (2)
[VO2(L)]2SO4 (3)

526

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Metal(II) perchlorates of [Cu(II), Ni(II)] and [VO(II)] sulphate (0.2 mmol) and the potential
binucleating Schiff base ligand (0.1 mmol) were dissolved in DMF (20 ml) and the mixture was heated
to reflux for 3 h and the reactions were monitored by TLC. After partial evaporation of the solvent,
solid (6575%) metal(II) Schiff base complexes (Scheme 2) were separated and dried in vacuum. The
analysis results are in good consistency with proposed formulas in Table 1.
Cyclic voltammetry
All voltammetric experiments were performed with a CHI760C electrochemical analyzer, in
single compartmental cells using Tetrabutylammonium perchlorate as a supporting electrolyte. The
redox behavior of the complexes have been examined in absence and in presence of CT-DNA at a
scan rate 0.2 Vs1 in the potential range +1.2 to -2.0 V. A three-electrode configuration was used,
comprised of a glassy carbon electrode as the working electrode, a Pt-wire as the auxiliary electrode,
and an Ag/AgCl electrode as the reference electrode. The electrochemical data such as cathodic peak
potential (Epc) and anodic peak potential (Epa) were measured.
Gel Electrophoresis
The cleavage of pUC18 DNA was determined by agarose gel electrophoresis [19]. The gel
electrophoresis experiments were performed by incubation of the samples containing 40 M pUC18
DNA, 50 M metal complexes and 50 M H2O2 in Tris-HCl buffer (pH 7.2) at 37oC for 2 h. After
incubation, the samples were electrophoresed for 2 h at 50 V on 1% agarose gel using Tris-acetic acidEDTA buffer (pH 7.2). The gel was then stained using 1 g cm-3 ethidium bromide (EB) and
photographed under ultraviolet light at 360nm. All the experiments were performed at room
temperature.performed at room temperature unless otherwise mentioned.
Antimicrobial activity
The in vitro antibacterial activity of the ligand and the complexes were tested against the
bacterias Bacillus subtilis, Klebsiella pneumoniae, Escherichia coli and Staphylococcus aureus by
well diffusion method using nutrient agar as the medium. Streptomycin was used as standard for
bacteria. The stock solution (10-2 mol L-1) was prepared by dissolving the compound in DMF and the
solution was serially diluted in order to find minimum inhibitory concentration (MIC) values. In a
typical procedure, a well was made on the agar medium inoculated with microorganisms in a Petri
plate. The well was filled with the test solution and the plate was incubated for 24 h for bacteria at 35
o
C. During the period, the test solution diffused and the growth of the inoculated microorganisms was
affected. The inhibition zone was developed, at which the concentration was noted.
RESULTS AND DISCUSSION
The binuclear Schiff base ligand prepared and reacts with transition metals Cu(II), Ni(II), and
VO(II) . The Schiff base ligand has been synthesized from 2- carboxybenzaldehyde and 3,3,4,4tetraminobiphenyl (scheme 1) ( C44H30N4O8) in 2:1 mole ratio. The results of elemental analyses were
in good agreement with those required by the proposed formulae given in Table 1. The data in
527

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

consistent with the earlier reports support the proposed formulation of the binuclear complexes
(scheme 2). The higher conductance values (Table 2) of chelates support the electrolytic (1:2) nature
of metal complexes. The new binuclear complexes are stable, hygroscopic with higher melting points,
insoluble in water, soluble in acetonitrile, chloroform, DMF and DMSO.
IR spectra
The important IR absorption frequencies of the synthesized complexes are shown in Table 3.
The azomethine nitrogen C=N stretching frequency of the free ligand appears around 1626 cm-1, which
is shifted to lower frequencies in the spectra of all the complexes (1600-1610 cm-1). These bands are
shifted to lower wave numbers indicating the involvement of azomethine nitrogen in coordination to
the metal ion [20]. The above i.r. data clearly indicate that the carbonyl groups of 2carboxybenzaldehyde have reacted with the amine groups of 3,3,4,4-tetraminobiphenyl through the
condensation of the metal atoms. The ligand displays C-O absorptions at 1310 cm-1. On complexation
this band is shifted to higher frequency in the range of 1350-1375 cm-1 for all the Schiff base
complexes which suggest that the carbonyl group is involved in coordination in the enol form through
deprotonation [21, 22]. This is further supported by the disappearance of the OH in the range of 34003440 cm-1 in all the complexes. Accordingly, the ligand acts as a tetradendate chelating agent bonded
to the metal ion via two C-O groups & two C=N azomethine nitrogen atoms of the Schiff base
(scheme 2).Assignment of the proposed coordination sites is further supported by the appearance of
medium bands at 480-510cm-1 and 430-470 cm-1 which could be attributed to M-O, M-N respectively
[23]. In addition, the Oxovanadium complexes shows a band at 981 cm-1 attributed to V=O stretching
frequency [24]. A further examination of Infrared spectra of complexes shows the presence of a band
in the 1060-1110 cm-1 region. The strong band is ascribable to ClO4- & SO42- ions [25].
Electronic Epectra
The UV-visible spectra are often very useful in the evaluation of results furnished by other
methods of structural investigation. The electronic spectral measurements were used for assigning the
stereochemistries of metal ions in the complexes based on the positions and number of d-d transition
peaks. The electronic absorption spectra of the Schiff base ligand and its complexes were recorded in
DMF solution in the range of 200 to 800 nm regions and the data are presented in Table 4. The
absorption spectrum of free ligand consist of an intense bands centered at 330 nm attributed to n- *
transitions of the azomethine group. Another intense band in higher energy region of the spectra of the
free ligand was related to * transitions of benzene rings. These transitions are also found in the
spectra of the complexes, but they shifted towards lower frequencies, confirming the coordination of
the ligand to the metal ions. Further, the d-d transition of the complex showed a broad band centered at
530 nm for Cu(II) complex Fig. 1. This is due to 2B1g 2A1g transition [26]. The spectra of Ni(II)
complex in the visible region at about 475 and 490 nm is assigned to 1A1g 1A2g, 1A1g 1B1g,
transitions, suggesting an approximate square planar geometry of the ligand around the metal ions
[27]. The intense charge transfer band at 460-470 nm in Oxovanadium(IV) complex assigned to 2B2
2
A1, 2B2 2E transitions. This is due to electron delocalization over whole molecule on complexation.
Based on these data, a square planar geometry has been assigned to the complexes except VO(II)
528

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

complex which has square pyramidal geometry. These values are comparable with other reported
complexes [28].
Magnetic Properties
The magnetic moments of the solid state complexes were measured at room temperature. The
measured magnetic moments of mononuclear copper(II) complex 1.68B.M. Magnetic susceptibility
measurements shows that these complexes are paramagnetic, which corresponds to the +2 oxidation
state of copper(II) complexes. The magnetic moment of binuclear Cu(II) complex 1.59 B.M. It can be
observed that the magnetic moment values of binuclear copper(II) complexes are slightly lower than
the mononuclear copper(II) complexes. The strong antiferromagnetic coupling that was found for
binuclear copper(II) complexes were explained as the good super exchange properties [29]. Similarly,
the magnetic moment of binuclear Ni(II) complex has 2.54 B.M. indicating a geometry [30], and the
magnetic moment of binuclear VO(II) complex 4.40 B.M. indicating a squarepyramidal geometry
[31].
ESR spectra
The ESR spectra of transition metal(II) complexes provide information of importance in
studying the metal ion environment. The X-band ESR spectrum of the copper complex was recorded
on a powder solid at room temperature Fig. 2. It exhibits an axial signal which can be interpreted in
= 2.06,
terms of tetrahedral species with a strong signal in the low field region, corresponding to g
and a weak signal in the high field region due to g || = 2.12. Splitting of the signal in the high field
region may be due to a difference in surroundings between the two copper(II) centers suggesting a
is evidence for
binuclear structure for this complex [32]. As seen in Table 5, the presence of g || > g
square planar geometry around copper(II) atom [33]. The axial symmetry parameter (G) value of
Cu(II) complex (less than 4) show that the exchange interaction is negligible. The Cu(II) complex is
dimer with the unpaired electron lies in the dx2 - y2 orbital. Its interesting to note that the g || > g
>
2
2
2.0023, indicating that the ground state of Cu(II) is predominantly dx - y . The observed values of
Vanadyl complex g || = 2.04 > g = 1.98 indicates that the unpaired electron is present in the dxy
orbital with square pyramidal geometry around the VO(II) chelates [34].
Cyclic Voltammetry Studies
Electroanalytical techniques are the most effective and versatile methods available for the
mechanistic study of redox systems. The important parameters of a cyclic voltammogram are the
magnitudes of the anodic peak potential (Epa) and cathodic peak potential (Epc). Cyclic voltammetry
has been employed to study the interaction of complex with CT-DNA. The cyclic voltammogram of
complex 1 in the absence of CT-DNA shows in Fig. 3(a) reveals non-Nernstain but fairly
reversible/quasi-reversible one electron redox process involving Cu(II)/Cu(I) couple. The cyclic
voltammograms of complexes were obtained in H2O/DMSO (95:5) solution, at scan rate 0.2 Vs1 over
a potential range from +1.2 to -2.0 V. In the absence of CT-DNA, complex 1 and other complexes data
are listed in Table 6. the anodic peak potential (Epa) of complex 1 appeared at 0.410 V and the
cathodic (Epc) at 1.320 V. The cyclic voltammograms of complex 1 reveals a one electron
529

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

quasireversible wave attributed to the redox couple Cu(II)/Cu(I) with the formal electrode potential,
EO = -0.865 V, and Ep= - 0.910 V which is larger than the Nernstian value observed for the one
electron transfer couple. On addition of CT-DNA, the complex 1 Fig. (3b) shows a shift in Epc= (1.100 V), Epa= (+0.212 V) and Ep= (0.888 V) values indicating strong binding of binuclear complex
with CT-DNA. The decrease in ratio of anodic to cathodic peak signify that adsorption of Cu(I) is
enhanced in the presence of CT-DNA. Further, the shift in E0value and increase in peak heights
potentials suggest that both Cu(II) and Cu(I) form of complex 1 bind to CT-DNA [35].
Cleavage of Plasmid pUC18 DNA
DNA cleavage is controlled by relaxation of supercoiled circular conformation of pUC18 DNA
to nicked circular conformation and linear conformation. When circular plasmid DNA is conducted by
electrophoresis, the fastest migration will be observed for the supercoiled form (Form I). If one strand
is cleaved, the supercoils will relax to produce a slower-moving open circular form (Form II). If both
strands are cleaved, a linear form (Form III) will be generated that migrates in between. Figure 4
illustrates the gel electrophoresis experiments showing the cleavage of plasmid pUC18 DNA induced
by the three binuclear complexes. The control experiments did not show any apparent cleavage of
DNA (lane 1 & 2). Copper binuclear complex in the presence of H2O2 (lane 1) at higher concentration
(50M) shows more cleavage activity compared to binuclear Nickel and Oxovanadium(IV)
complexes. The supercoiled plasmid DNA was completely degraded. This shows that a slight increase
in the concentration over the optimal value led to extensive degradations, resulting in the
disappearance of bands on agarose gel [36]. Nickel binuclear complex in the presence of H2O2
resulting the conversion of supercoiled form (Form-I) into linear form (Form-III) (lane 4).
Oxovanadium binuclear complex in the presence of H2O2 (lane 5) at higher concentration (50M)
shows cleavage activity in which supercoiled DNA (Form-I) cleaved and supercoiled form converted
to open circular form (Form-II). The results revealed that the Cu(II), Ni(II) complexes have more
cleavage than VO(II) complex . Probably this may be due to the formation of redox couple of the
metal ions and its behaviour. Further the presence of a smear in the gel diagram indicates the presence
of radical cleavage [37].
Antimicrobial activity
The ligand and their complexes have been tested for invitro growth inhibitory activity against
gram-positive microbes Bacillus subtilis, staphylococcus aureus and gram-negative microbes
Klebsiella pneumonia, Escherichia coli by using well-diffusion method. As the test solution
concentration increases, the biological activity also increases. It is found that the activity increases
upon co-ordination. The increased activity of the metal chelates can be explained on the basis of
chelation theory [38]. The orbital of each metal ion is made so as to overlap with the ligand orbital.
Increased activity enhances the lipophilicity of complexes due to delocalization of pi-electrons in the
chelate ring [39]. In some cases increased lipophilicity leads to breakdown of the permeability barrier
of the cell [40, 41]. The results revealed that the metal complexes Cu(II), Ni(II) and VO(II) have
higher antimicrobial activity than the ligand are shown in Figs. 7(a), 7(b) and 7(c) and Table 8.

530

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

CONCLUSION:
The N2O2 type Schiff base ligand is synthesized from 2-carboxybenzaldehyde and 3,3,4,4tetraminobiphenyl. It acts as a tetradentate ligand and forms stable complexes with transition metal
ions such as Copper(II), Nickel(II), and Oxovanadium(IV). The ligand and its complexes are
characterized using spectral and analytical data. The interaction of these complexes with CT-DNA was
investigated by gel electrophoresis. All the transition metal complexes have higher activity than the
control CT-DNA. The Cu(II), Ni(II) complexes have more activity than VO(II) complex and the
control CT-DNA. . The metal complexes have higher antimicrobial activity than the free ligand.

Scheme 1 Structure of binucleating tetradendate Schiff base ligand

531

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Where, M = Cu(II), Ni(II), X = 4ClO4-

Scheme 2 Structure of binuclear Cu(II), Ni(II), VO(II) Schiff base complexes

532

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Table 1 Physical characterization, analytical data of the ligand and binuclear Schiff basecomplexes

Found (Calculated) (%)


M.
Complexes

point

Color

Yield
(%)
M

( C44H30N4O8)

240

Yellow

71.15

7.54

4.04

(71.56)

(7.68)

(4.06)

10.14

41.83

4.43

2.06

(10.45)

(41.90)

(4.46)

(2.25)

9.43

42.17

4.47

2.08

(9.47)

(42.24)

(4.52)

(2.10)

12.54

49.62

5.26

2.44

(12.59)

(49.66)

(5.29)

(2.47)

90

(Ligand )
Pale
[Cu2(L)]4ClO4

260

[Ni2(L)]4ClO4

248

[VO2 (L)]2SO4

279

green

Dark
green

Green

80

75

70

533

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Table 2 Molar conductance data of the binuclear Schiff base the complexes

Solvent

Molar conductance
m (ohm-1cm2mol-1)

Types of
electrolyte

Complexes
[Cu2(L)]4ClO4

MeCN

250

1:2

DMF

210

1:2

MeCN

100

1:2

DMSO

135

1:2

MeCN

154

1:2

DMSO

142

1:2

[Ni2(L)]4ClO4

[VO2 (L)]2SO4

534

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Table 3 Infrared spectral data for the ligand and binuclear Schiff base complexes

Complexes

(-C-O)
-1

(cm )

(-C=N)
-1

(cm )

(-OH)
-1

(cm )

(V=O)
-1

(cm )

(M-O)
-1

(cm )

(M-N)
-1

(cm )

ClO4/SO42(cm-1)

( C44H30N4O8)

1310

1626

3420

1350

1600

3390

510

460

1080

1610

3430

495

470

1110

1602

3385

981

430

1060

(Ligand)
[Cu2(L)]4ClO4

1362
[Ni2(L)]4ClO4
[VO2(L)]2SO4

1375

480

Table 4 Absorption spectral data of the ligand and binuclear Schiff base complexes

Absorption (max)(nm)
*

n*

d-d

Benzene/ imino

Azomethine

273,254

330

[Cu2(L)]4ClO4

550

279,245

334

430

[Ni2(L)]4ClO4

490

270,250

332

375

[VO2 (L)]2SO4

470

275,242

333

420

Complexes
( C44H30N4O8)
Ligand

535

LMCT

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Table 5 ESR spectral data of the binuclear Schiff base complexes

Complexes

g ||

g iso

[Cu2(L)]4ClO4

2.12

2.06

2.07

2.03

2.04

1.98

2.02

-1.44

[VO2 (L)]2SO4

Table 6 Cyclic voltammetric data of the binuclear Schiff base Complexes in DMSO solution.

Complexes

Couple

Epc (V)

Epa (V)

Ep(mv)

[Cu2(L)]4ClO4

Cu(II)/ Cu(I)

-0.563

-0.454

0.109

[Ni2(L)]4ClO4

Ni(II)/Ni(I)

1.60

1.85

0.25

0.61

0.75

0.14

-1.63

-0.52

1.11

[VO2 (L)]2SO4

VO(IV)/ VO(III)
VO(IV)/VO(V)

536

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Table 7 Antibacterial activity of the ligand and binuclear Schiff base complexes

Complexes

Klebsiella pneumoniae

Escherichia coli

(mm)

(mm)

25

50

75

100

25

(l)

50

75

Staphylococcus aureus

Bacillus subtilis

(mm)

100

25

(l)

50

75

(mm)

100

25

(l)

50

75

100

(l)

( C44H30N4O8)

10

12

12

14

11

12

14

14

11

13

14

16

11

13

15

18

[Cu2(L)]4ClO4

11

13

14

17

12

13

15

16

12

14

15

16

10

14

17

19

[Ni2(L)]4ClO4

12

15

17

18

13

14

15

18

13

15

16

19

12

15

17

20

[VO2(L)]2SO4

13

14

16

19

14

15

17

19

14

15

17

19

11

13

16

18

537

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Fig. 1 Absorption spectra of the binuclear Cu(II) Schiff base complex [Cu2(L)]ClO4.

Fig. 2 X-band ESR spectra of binuclear Cu(II) Schiff base complex [Cu2(L)]ClO4 at room temperature.

Fig. 3 Cyclic voltammogram (scan rate 0.2 Vs1, DMSO, 29 C) of


Fig. 3a Complex 1 alone

Fig. 3b Complex 1 in presence of CT-DNA [Complex 1] 1 103M, [DNA] 6 103 M.

538

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Fig 4. Changes in the agarose gel electrophoretic pattern of pUC18DNA induced by H2O2 and metal
complexes: Lane 1, DNA alone; Lane 2, DNA alone + H2O2; Lane 3, DNA + Cu binuclear complex +
H2O2; Lane 4, DNA + Ni binuclear complex + H2O2; Lane5, DNA + VO binuclear complex + H2O2.
Fig. 5 Difference between the antimicrobial activity of binuclear ligand & metal complexes

Fig. 5a (1) ligand ( C44H30N4O8), (2) [Cu2(L)]4ClO4 , (3) [Ni2(L)]4ClO4, (4) [VO2(L)]2SO4

539

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Staphylococcus aureus
Zone of inhibition (mm)

20
15

25 (l)
50 (l)

10

75 (l)
100 (l)

5
0
1

Complexes

Fig. 5b Difference between the antimicrobial activity of binuclear ligand & metal complexes (1) ligand (
C44H30N4O8), (2) [Cu2(L)]4ClO4 , (3) [Ni2(L)]4ClO4, (4) [VO2 (L)]2SO4. [X axis Zone of Inhibition (mm)]

Fig. 5c Difference between the antimicrobial activity of binuclear ligand & metal complexes (1) ligand
(C44H30N4O8), (2) [Cu2(L)]4ClO4 , (3) [Ni2(L)]4ClO4, (4) [VO2 (L)]2SO4. [X axis Zone of Inhibition (mm)]

540

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

Fig. 5d Difference between the antimicrobial activity of binuclear ligand & metal complexes (1) ligand
(C44H30N4O8), (2) [Cu2(L)]4ClO4 , (3) [Ni2(L)]4ClO4, (4) [VO2 (L)]2SO4. [X axis Zone of Inhibition

REFERENCES:
1.

X. F. Luo, X. Hu, X. Y. Zhao, S. H. Goh and X. D. Li., Polymer., 2003, 44, 5285.

2.

L. Streyer, Biochemistry, Freeman, New York, 1995.

3.

V. Razakantoanina, N. K. P. Phung and G. Jaureguiberry., Parasitol. Res., 2000, 86, 665.

4.

Q. X. Li, H. A. Tang, Y. Z. Li, M. Wang, L. F. Wang and C. G. Xia., J. Inorg. Biochem.,


2000, 78, 167.

5.

P. J. E. Quintana, A. De Peysder, S. Klatzke and H. J. Park., Toxicol. Lett., 2000, 85, 117.

6.

R. Baumgrass, M. Weivad and F. Erdmann., J. Biol. Chem., 2001, 276, 47914.

7.

R. Vafazadeh and M.Kashfi., Bull.Korean.Chem.Soc., 2007, 28, 1227.

8.

A. Anora and K. P. Sharma., Synth. React. Inorg. Met.-Org. Chem., 2000, 32, 913.

9.

E. Canpolat and M. Kaya., J. Coord. Chem., 2004, 57, 127.

10. A. P. Mishra, M. Khare and S. K. Gautam., Synth. React. Inorg. Met.-Org. Chem.,
2002, 32, 1485.
11. Y. Fan and C. Bi, J. Li., Synth. React. Inorg. Met.-Org. Chem., 2003, 33, 137.
541

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

12. E. Canpolat, M. kaya and A. Yazici., Spectrosc., Lett., 2005, 38, 35.
13. R. C. Maurya, P. Patel and S. Rajput., Synth. React. Inorg. Met.-Org. Chem., 2003, 33, 817.
14. A. P. Mishra and S. K. Gavtarm., J. Ind. Chem. Soc., 2004, 81, 324.
15. N. A. Venkariya, M. D. Khunt and A.P. Parikh., Ind. J. Chem., 2003, 42B, 421.
16. A. R. Banerjee, J. A. Jaeger and D. H. Turner., Biochemistry., 1993, 32, 153.
17. N. Y. Sardesai, K. Zimmerman and J. K. Barton., J. Am. Chem. Soc., 1994, 116, 7502.
18. A. I. Vogel, Text Book of Practical Organic Chemistry, 5th ed., Longman, London, 1989.
19. N. Ramana, R. Jeyamurugana, A. Sakthivel, L. Mitub., Spectrochim. Acta PartA., 2010,
75, 88.
20. S. Pal and S. Pal., J Chem Soc Dalton Trans., 2002, 2102.
21. E. Canpolat and M. Kaya., J.Coord.Chem., 2002, 55, 1419.
22. E. Canpolat and M. Kaya., Polish .J. Chem., 2003, 77, 961.
23. S. Yamada and A .Takeuchi., Coord Chem Rev.,1982, 43,187.
24. B. Kaitner and G. Parlovic., Croatica Chemica Acta., 1999,72, 607.
25. K. Nakamoto, Infrared and Raman spectra, of inorganic and coordination compounds, III
edition, John Wiley A(1978).
26. L. N. Sharadha and M. C. Ganorkar., Indian J Chem., 1988, 27A, 617.
27. A. B. P. Lever Crystal Field Spectra, Inorganic Electronic Specroscopy, first Ed.Elsevier.,
Amsterdam 1968, 249.
28. M. Sivasankaran Nair, G. Kalalakshmi and M. Sankaranarayana Pillai., J Indian Chem Soc.,
1999, 76, 310.
29. J. D. Ranford, J. J. Vittal and Y. M. Wang., Inorg. Chem., 1998, 37, 1226.
30. R. K. Agarwal, D. Sharma, L. Shing and H. Agarwal., Bioinorg. Chem. Appl, 2006, doi. 10.
1155/BCA/2006/29234.
31. R. Aurkie, S. Banerjee, S. Sen, R. J. Butcher, G. M. Rosair and M. T. Garland., Struct. Chem,
542

Asian Journal of Biochemical and Pharmaceutical Research

Issue 2 (Vol. 1) 2011

2008, 19, 209.


32. S. S. Tandon, L. C. Laurence, K. Thompson, S. P. Connors and J. N. Bridson., Inorg Chim
Acta., 1993, 213.
33. Y. Dong, L. F. Lindoy, P. Turner and G. Wei., Dalton Trans., 2004, 8, 1264.
34. L. Leelavathy, S. Anbu, M. Kandaswamy, N. Karthikeyan and N. Mohan., Polyhedron., 2009,
28, 903.
35. Z. S. Yang, Y. L. Wang and G. C. Zhao., Anal Sci., 2004, 20, 1127.
36. N. Raman, T. Baskaran and A. Selvan., J. Iran. Chem. Res., 2008, 1, 29.
37. A. M. Thomas, A. D. Naik, M. Nethaji and A. R. Chakravarty., Indian J Chem., 2004, A43, 691.
38. N. Raman, A. Kulandaisamy, A. Shanmugasundaramand, K. Jeyasubramanian., Trans.
Met.Chem., 2001, 26, 131.
39. R. S. Srivastava., Inorg. Chim. Acta., 1981, 56, 65.
40. N. Gupta, R. Swaroop and R. V. Singh., Main Group Met. Chem., 1997, 14, 387.
41. A. Cukurovali, I. Yilmaz, H. Ozmen and M. Ahmedzade., Trans. Met. Chem., 2002, 27, 171.

*Correspondence Author: R.Rajavel, Department of Chemistry, Periyar University, Salem,


Tamilnadu, India.

543

You might also like