chapter 1
chapter 1
CHAPTER 1
1. Introduction
Coordination chemistry has become an interesting field of research rather than a closed
body of knowledge. Nowadays a large number of articles on coordination compounds are being
published. New synthetic techniques allow the identification of a huge number of fascinating
complexes and in the laboratory they continue to pose synthetic challenges. Metal complexes
play an important and diversified roles in biological systems. The role of chlorophyll,
haemoglobin, carbonic anhydrase, vitamin B12, xanthine oxidase and haemocyanin shows the
intimate linkage between inorganic chemistry and biology. Studies of these types of metal
complexes are now a part of the highly demanding field, bio-inorganic chemistry.
Metal complexes find attractive applications in the field of biology. Metal complexes are
used as diagnostic and therapeutic agents. Studies on metal based anticancer drugs and
antiarthritic agents are some of the currently active topics of investigation in bio-inorganic
chemistry. Many metal complexes, especially those of Schiff bases, have been studied from the
point of view of using them as antibacterial, antifungal and anticancer drugs. In this chapter, a
review of Schiff bases, their metal complexes and their general applications is presented.
Schiff bases are an important class of ligands due to their synthetic flexibility, selectivity
and sensitivity towards the central metal atom and structural similarities with natural biological
substances. Schiff base ligands considered as ‘privileged ligands’1 contain the azomethine group
(-CH=N-) and they are prepared by the condensation of a primary amine and a carbonyl
compound. The azomethine group is particularly suited for binding to metal ions via the lone pair
2
on nitrogen. When the Schiff bases contain one or more donor atoms in addition to -C=N- group,
they act as polydentate chetating ligands or macrocycles. Because of the versatility of the Schiff
bases, a large number of complexes with interesting structures are synthesized even now. Schiff
bases derived from aliphatic aldehydes are unstable and are readily polymerizable2 while those
derived from aromatic aldehydes are more stable. The common Schiff bases are crystalline and
feebly basic in nature. In general, aldehydes react faster than ketones in condensation reactions,
leading to the formation of Schiff bases as the reaction centre of an aldehyde is sterically less
hindered than that of a ketone. Furthermore, the extra carbon of ketone donates electron density
to the azomethine carbon and thus makes the ketone less electrophilic compared to aldehyde3.
Schiff bases can act as bidentate, tridentate, tetradentate or polydentate ligands depending
upon the other functional groups present and they are capable of forming very stable complexes
with transition metals. If they bear a functional group, usually the hydroxyl, sufficiently near the
site of condensation in such a way that a five or six membered ring can be formed when reacting
Schiff bases derived from aromatic amines and aromatic aldehydes have a wide variety of
OH
C
R N
OH
NR12
C N
R R1
Bidentate Tridentate
CH3
OH HO R R
C C
C N N C N OH N
R R
OH HO
Tetradentate Pentadentate
Generally Schiff bases are prepared under acid or base catalysis or with heat. When two
equivalents of salicylaldehyde are condensed with a diamine, a particular type of chelating Schiff
base is produced, the so-called Salen ligands, with four coordinating sites and two axial sites
open to ancillary ligands, are very much like porphyrins, but can be more easily prepared. The
term Salen was used only to describe the tetradentate Schiff bases derived from ethylenediamine
(Figure. 1.2).
R
H
N OH
N OH
N OH 4R=H
N OH H 5 R = tBu
R
R
Salen Chiran salen
Desai et al6. Recent report says, some Schiff bases (Figure 1.3) are also employed as fluorescent
N N
N N
NaO3S
(I) (II)
NO2 NO2
N N
NaO3S
(III) (IV)
Cl Cl
N N
NaO3S
(V) (VI)
Schiff base metal complexes are generally prepared by treating metal halides with
Schiff base ligands under suitable experimental conditions. In special cases metal alkoxides,
metal amides, metal alkyls or metal acetates have been used for the synthesis. There are
numerous literature reviews on the synthesis and characterization of Schiff base metal
complexes8-10.
Generally transition metal ions are used to prepare coordination complexes, as there is
and tetradentate Schiff base ligands are reported in literature11-13. The –OH or –SH groups when
present ortho to the azomethine moiety in the Schiff base can induce tautomerism in the
The transition metals, especially first row transition metal ions, are well known for their
ability to form wide range of coordination complexes in which octahedral, tetrahedral, and
square planar geometries predominate. Copper(II) is a typical transition metal ion to form
complexes, but less typical in its reluctance to take up a regular octahedral (or) tetrahedral
geometry. The magnitude of the splitting of the electronic energy levels in copper(II) complexes
tend to be larger than other first row transition metals due to the presence of large Jahn-Teller
distortion. Copper is one of the essential trace elements present in living organisms. A number of
important redox enzymes like hemocyanins, superoxide dismutase, blue copper proteins, etc.,
Copper(II) complexes with amino acids are cited as having potent anti-inflammatory and
anti- ulcer activity14. Copper ions are found to be present in the active sites of many number of
metalloproteins, which are involved in important biological electron transfer reactions as well as
in molecular oxygen redox reactions15-16. There has been a substantial interest in the rational
design of novel transition metal complexes, which bind and cleave duplex DNA with high
metal complexes has been substantially aided by the DNA cleavage chemistry that is associated
pair stack of DNA has been an area of interest for some time. It is clear that a number of factors
affect both the sequence selectivity of intercalators binding to DNA and the resulting twist angle,
and therefore it is necessary to understand the structural features of intercalators that control the
specificity of their binding. This, inturn should lead to the design of more specifically targeted
intercalators. The synthesis and investigation of synthetic reversible dioxygen carriers have
attracted substantial interest in recent years and their physico-chemical properties are sufficiently
favourable for application in industries and medicine17-18. Synthetic oxygen carriers have been
studied extensively over several decades for two main reasons viz (i) to understand the
mechanism of oxygen binding proteins and (ii) to design complexes suitable for practical
applications. Co(II) complexes of porphyrins, Schiff bases and tetraaza systems (Figure 1.4 (A-
C) are usually studied as models for oxygen carriers. Schiff bases used for the studies of oxygen
carrying properties are generally tetradentate, of which at least two of the ligating atoms should
be nitrogen, with the others being nitrogen, oxygen, sulphur (or) combination of the three.
Tian et al19 reported the nuclease activity (DNA cleavage) of Cobalt(II) complexes with
OH HO
N N
C Co C
H N N H
A
7
Figure 1.4. (A-C) Co(II) complexes of porphyrins, Schiff bases and tetraaza systems
pyrazole derivative in humans and due to its antipyretic property he named the compound
"Antipyrine". Transition metal complexes of pyrazolone derivatives are of huge interest due to
their biological activities, especially pyrazolone Schiff-base derivatives. Among the pyrazolone
derivatives, 4-aminoantipyrine forms a variety of Schiff bases with aldehydes/ketones, and they
applications21. 4-Aminoantipyrine has an N-phenyl group and a –CH group on either side of a
polar carbonyl group, thus resembling N-substituted amides. Coordination chemists, medicinal
chemists and analytical chemists have extensively studied 4-aminoantipyrine. The carbonyl
group in 4-aminoantipyrine is a potential donor due to the large dipole moment (5.48 D) and
Three modes of coordination must be considered for the system, unidentate with bonding
through either oxygen or the amino nitrogen or chelation utilizing both these donors. The
majority of the Schiff bases of 4-aminoantipyrine derivatives are obtained in good yield by
protonated) of the ligand precursor, it is possible to get either neutral or protonated forms of the
ligands. Since 4-aminoantipyrine has an additional possible coordination site in the amino
nitrogen, it was considered really valuable to study the complexes of this ligand.
Patel25 carried out the physico-chemical studies on antipyrine complexes of CoCl2, CoBr2,
CuCl2, and CuBr2 of the type MX2.2AAP. Both electrical conductance and cryoscopic
electrolytes.
Bose and Patel26 prepared the Cu(II) nitrate and bromide complexes of 4-
prepared antipyrine complexes of Fe(III) chloride and thiocyanate. The decrease in value of the
C=O stretch in IR spectra predict coordination through the carbonyl oxygen to Fe(III).
(n=1,X=Cl, Br, NCS or NO3; n=2,X=I or ClO4, L=HNAAPS or CAAPS) and characterized
1.5).
N N
OH H
H O O
N N
N N N
N N N
NH2 NH2
HNAAPS CAAPS
Kuncheria et al29 synthesized ten new complexes of thorium(IV) nitrate with Schiff bases
elemental analysis, molecular weight determination, conduction and spectral studies. Mn(II),
10
Co(II) and Ni(II) complexes with tridentate Schiff bases derived from 4-aminoantipyrine and
prepared. For all complexes, spectroscopic results suggested an octahedral geometry30. Savant
and Ramamurthy31 studied antipyrine complexes of Ti(II), Zr(II), Th(IV) and U(IV) perchlorates.
They are stable at room temperature and decomposed exothermally at 300°C. Spectral studies
indicated the bonding of metal to ligand through carbonyl oxygen. Nair et al synthesized new
lanthanide nitrate complexes of the Schiff base derived from 4-aminoantipyrine and 2-
hydroxybenzaldehyde32, 33
viz. 4-[N-(2-hydroxybenzylidene)amino]antipyrine. Gadre et al34
reported isolation and structure elucidation of complexes of the Schiff-base ligands, salicylidene-
group with the cyclic carbonyl present in the 4-aminoantipyrine36. Ismail et al37 synthesized
complexes derived from catecholamine and 4-aminoantipyrine. These were characterized by IR,
UV-Vis, magnetic and thermal studies which indicated that Fe(III) forms 1 : 2 (M :
reveal the presence of Fe(III) chelates in octahedral geometry while the Cu(II) chelates are
N N O N
O N O N
N
HN NH HN
NC
R OH O R O OH2
R
Fe Cu
O HO O
OH2 OH2
conductance and infrared studies. Thermal stabilities of these complexes were investigated by
thermogravimetric analysis. Issaadi et al40 synthesized an tetradentate N2O3 type Schiff base and
its complexes with Cu(II), Co(II), Zn(II) and Cd(II) which were characterized by spectroscopic
determinations and cyclic voltammetry. The spectral techniques indicated that Cu(II) and Cd(II)
complexes are mononuclear while Co(II) and Zn(II) complexes appeared to be binuclear.
such as perchlorate, nitrate, thiocyanate, chloride and bromide have been prepared by Linert et
analysis. Kinetic parameters like activation energy, pre-exponential factor and entropy of
H
C N
N N
N O O N
Issa et al42 synthesized five Schiff bases derived from 4-amino antipyrine and
1
benzaldehyde derivatives; their UV–Vis, IR, H NMR and fluorescence spectra were
delocalization molecular orbital theory. The square planar and octahedral geometry of the 4-
O H
NH2
OH
N CH O
N O
N . nH2O
N Cu
N O X
Copper salt
metal salt Cl
Cl
N Cl N
M
N O Cl
N
N
Octahedral geometry
Donia and Ebeid43 synthesized Ni(II) complexes using ligands such as 4-[N-(2-hydroxy-
were studied using differential thermal analysis, electronic and IR spectroscopy, X-ray powder
diffraction and electrical conductivity. Thermochromism in these complexes has been attributed
to dehydration.
Liang et al44 synthesized Cu(II) complexes of the Schiff base derived from
analysis, IR, 1H NMR and single crystal XRD studies. Alaudeen et al45 carried out transition
metal chelates of Mn(II), Cr(II), Fe(III), Co(III), Zn(II) and Cd(II) with 4-aminoantipyrine,
measurements, IR, UV-Vis., and Mossbauer spectral studies, which suggested an octahedral
of Schiff bases derived from 4-aminoantipyrine and some aromatic aldehydes, which were
electronic spectra and magnetic susceptibilities and were also screened for antibacterial and
antifungal properties.
Sinha et al49 synthesized Pd(II) and Ag(II) complexes derived from 4-aminoantipyrine
and imidazole. These complexes were characterized by UV-Vis, IR and 1H-NMR spectral
studies. Rosu et al50 synthesized Cu(II) complexes derived from 4-aminoantipyrine and
14
Determinations of the antimicrobial activity of the ligands and the complexes were carried out
against some bacteria. Complexes of uranyl(VI) nitrate with functional Schiff bases derived from
synthesized from neutral or alkaline media and their spectra-structure correlation obtained. The
: 1 : 1 molar ratio. Microanalytical and spectral techniques were used to confirm the structures.
techniques. All the complexes were monomeric and neutral with square pyramidal geometry.
Nair and Mathew54,55 reported mercury(II)/dioxotungston chelates of azo dyes derived from 4-
against some bacteria for antibacterial activity. Nair et al56 synthesized Cu(II) complexes derived
nickel(II), copper(II) and zinc(II) complexes derived from 2-formylpyridine, N-antipyrinyl and
thiosemicarbazone. Ligand field parameters were calculated and the proposed stereochemistry is
15
based on various physical and spectral methods. Bhattacharaya et al58 prepared adducts of some
aryltinhalides with antipyrine. The IR data of this adduct indicated coordination through the
carbonyl oxygen of the ligand. Bailey and Peterson59 observed that 4-aminoantipyrine was a
monodentate ligand bonding through nitrogen in complexes of Cu(II), Co(II) and Ni(II) only
when bromide anion was coordinated to the metal. This may be due to the greater ability of
bromide to accept electrons from the metal ion via back bonding, thereby reducing the electron
density on the metal. This effect could be counteracted by donation from nitrogen rather than
oxygen. A large number of compounds of antipyrine have been shown to be unidentate; oxygen
V. K. Revankar et al63 reported Co(II), Ni(II), Cu(II) and Zn(II) complexes derived from
techniques. All the complexes were monomeric and neutral with octahedral geometry (Figure
1.9).
O NH2 N OH
N N
O O
N N O N N
N N N
N H H O HO
OO N Reflux in ethanol
Reflux in ethanol
H for 3 h for 3 h
metal salt
metal salt
N N
N
N O
N O N
O NH N
N M . 2H2O
O N O
. H2O O N
Cl M
Cl H2O N
Prakash et al reported64 Cr(III), Mn(II), Co(III), Ni(II), Cu(II), Zn(II) and Cd(II) chelates
of Schiffs base derived from vanillin and 4-aminoantipyrine. All the complexes are shown to be
The chelates of VO(II), Cr(III), Mn(II), Fe(III), Co(II), Ni(II), Cu(II) and Zn(II) with 4-
magnetic measurements and spectral studies. The ligand field parameters for VO(IV), Cr(III),
Co(II) and Ni(II) chelates were calculated. Infrared spectra of 4AAP and its duplexes are quite
The carbonyl group of 4-aminoantipyrine was also condensed by primary amino groups.
Ni(ClO4)2.2L. The complexes were characterized on the basis of elemental analyses, molecular
weight, conductivity, magnetic moment, IR and electronic spectral data. The coordination
number of Ni2+ in these complexes is presumed to be six with octahedral geometry. In all the
Thermogravimetric data of the Nickel complex indicated that the complex is stable to ca 140°C,
complexes were made on the basis of elemental analysis, molecular weight, magnetic moment,
were also investigated68. M.M. Omar et al69-70 synthesized complexes of Schiff base derived
Chandra et al reported71 Co(III), Ni(II) and Cu(II) complexes of Schiff base derived from
characterized by spectral methods. The nickel(II) complexes were found to have octahedral
O O
S
NH HN
N N
N N N
H2C CH CH2
3 H3 C
Neutral tetradentate complexes of Cu(II), Ni(II), Co(II), Mn(II), Zn(II) and VO(IV) were
aminoantipyrine in ethanol and characterized by microanalytical data, IR, UV-Vis, 1H-NMR and
ESR spectra. The IR and UV-vis spectra suggest that all the complexes have square-planar
geometry except vanadyl and manganese complexes which show square-pyramidal and
octahedral geometries, respectively. The redox behavior of copper and vanadyl complexes has
been studied by CV and ESR spectra of copper and vanadyl complexes were also discussed.
Antimicrobial activity of the Schiff base and complexes have been extensively studied on some
microorganisms. The general structure of Schiff base metal complexes is shown in Figure 1.11.
18
H
R CH3
the Cu(II) complexes suggest square planar geometry. The DNA binding properties of copper
complexes were studied by using electronic absorption spectra, viscosity and thermal
denaturation experiments.
Raman et al75 synthesized neutral tetradentate complexes of Cu(II), Ni(II), Co(II), Mn(II),
VO(IV) and Zn(II) using Schiff bases derived from Salicylaldehyde with 4-aminoantipyrine and
exhibit square-planar geometry except the Mn(II) and VO(IV) complexes. The Mn(II) chelates
show an octahedral environment and the VO(IV) chelates exhibit square pyramidal geometry.
Neutral complexes Cu(II), Ni(II) and Zn(II) derived from salicylaldehyde/o-vanillin with
4-aminoantipyrine and 2-aminobenzoic acid. Their structural features and other properties were
deduced from elemental analysis, magnetic susceptibility and molar conductivity as well as from
mass, IR, UV–vis, 1H NMR and EPR spectral studies. All the complexes have square planar
geometry. DNA binding and cleavage studies are also reported76. Raman et al77 synthesized
aminoacid based Cu(II), Ni(II) and Zn(II) complexes of tetradentate Schiff base ligand derived
19
from salicylaldehyde with 4-aminoantipyrine and alanine/valine. All the complexes have square
planar. Antibacterial, antifungal, DNA binding and cleavage studies were also made. Raman et al
aminoantipyrine and 2-aminobenzoic acid. Cu(II), Co(II), Ni(II) and Zn(II) complexes are
octahedral geometry (Figure 1.12) and the DNA studies were carried out by absorption spectra,
N R
N O N O
N M
N
M
N O N
N
O
O
R
O
R = H and OCH3 R = CH3 and (CH3)2
M = Cu(II), Ni(II) and Zn(II) M = Cu(II), Ni(II) and Zn(II)
Figure 1.12. Schiff base metal complexes derived from 2-aminobenzoic acid/aminoacid
Cu(II) complexes have been synthesized from the Schiff base derived from salicylidene-
structural features have been determined from microanalysis and spectra, which show that all the
complexes were square planar. The antimicrobial activity and powder XRD pattern of all the
complexes were also reported21. Joseph et al79 synthesized copper complexes of Schiff base
Spectral results showed that all complexes are square planar. The antimicrobial activity and
N O
Cu
N N Cl
N
X
X = H, Cl, CH3 and NO2
forming cationic complexes with Cu(II), Ni(II), Co(II) and Zn(II) salts in ethanol. All the
synthesized complexes were characterized by spectral techniques. The IR and UV-vis spectra
suggested that all complexes were square planar80-82. The Schiff base derived from the
was a tetradentate ligand and all the complexes are square planar except cobalt complex which
has octahedral geometry83. Prakash et al84 reported Mn(II), Co(II), Ni(II), Cu(II) and Zn(II)
phenylenediamine. The data of the complexes suggested square planar geometry for the metals
with primary valency two. They also reported Mn(II), Co(II), Ni(II) and Cu(II) complexes of
Schiff base from vanillin with 4-aminoantipyrine and o-phenylenediamine85. Spectra results
showed that all complexes have square planar geometry (Figure 1.14).
21
R R
CH CH
N N
M
N N
N N N N
Figure 1.14. Schiff base derived from o-phenylenediamine with different aldehydes
Co(ClO4)2.2L. All the complexes were six coordinated according to magnetic and electronic
spectral measurements. The IR data of the complexes indicated that both thiosemicarbazones
were neutral tridentate (N, N, S) ligands86. Raman et al87 reported Cu(II) and Zn(II) complexes
voltammetric measurements and EPR spectral studies. All the complexes were monomeric and
Versatility of Schiff base ligands and biological, analytical and industrial applications of
their complexes make further investigations in this area highly desirable. The applications of
Schiff base complexes88-89 of various transition metals have been investigated for their
applications in various fields. Oxygen and nitrogen donor Schiff bases (salen type) are of
particular interest because of their ability to form transition metal complexes with unusual
coordination polymers and helical assemblies92-94. They have been found to be associated with
Co(II), Ni(II), Cu(II) and Zn(II) complexes of the Schiff base derived from vanillin and
DL-α-aminobutyric acid found to exhibit higher antibacterial activity compared to the Schiff
base98. Copper, nickel, iron and zinc complexes of the Schiff base derived from salicylaldehyde
with 2-aminobenzoic acid show good antibacterial activity against several pathogenic bacteria
Pseudomonas aeruginosa, and Staphylococcus Pyogones and fungal activity against candida
albicans100. N-5-chloro-salicylidiene tauriene Schiff bases and its copper and nickel complexes
Copper, cobalt and nickel complexes of Schiff base from diacetyl with ethylenediamine
and benzoic acid showed antibacterial activity against Bacillus subtilis and Escherichia coli and
antifungal activity against Aspirgillus niger and Aspirgillus flavus102. Mn(II), Co(II), Ni(II) and
23
Cu(II) complexes of Schiff base derived from Pyrrole-2-aldehyde and ethylenediamine exhibit
antibacterial activity against Escherichia coli and Staphylococcus aureus103 (Figure 1.15).
C N N C
N H H N
H H
1.4.2. Biological applications of 4-aminoantipyrine based Schiff bases and their metal
complexes
aeruginosa and Bacillus subtilis. The complexes showed higher inhibitory activity than the
ligands and have higher activity than ampicillin, except for Klebsilla pneumoniae and
Pseudomonas aeruginosa21.
Rosu et al. reported biological activity of certain ligands and their copper complexes50 on
samples of E. coli, K. pneumoniae, A. boumanii, S. aureus, P. aeruginosa and Candida sp. The
Schiff-base ligand and its copper complex exhibited high bactericidal activity towards E. coli and
A. boumanii, good activity against S. aureus and E. coli, exhibiting their potential usefulness as
broad spectrum antimicrobial agents. The biocidal activity studies of the oxovanadyl complexes
and the free ligands towards three gram-positive bacteria: B. subtilis, Sarcina lutea and S.
aphylococcus; seven gram negative bacteria viz E. coli, K. pneumoniae, P. aeruginosa, S. typhi,
Serratia marcescens, Shigella sonnie and Proteus mirabilis; three fungal species: Aspergillus
24
flarus, Penicillium chrysogenum and candida albicans, have been made by Ismail et al37; the
activity is related to the nature and structure of the complexes. The copper complex was the most
promising broad spectrum antimicrobial agent due to the presence of coordinated anion and
bridged –OH- with higher antimicrobial activity than the other complexes. It is found to be
superior to all other complexes against all the test organisms except P. aeruginosa.
microorganisms, and some other biotests have also been performed. In most tests, the activity of
activity increased but there is no evidence of further clinical tests. The efficacy of metal-based
therapeutic agents changes considerably by making small changes in the Schiff-base ligand
and VO(II) and MnL.2H2O was tested against bacteria such as S. typhi, S. aureus, K.
pneumoniae, B. subtilis, S. flexneri and P. aeruginosa and the fungi A. niger and R. bataicola.
From the data the inhibition zone of the metal chelates is found to be higher than that of the
ligand. Such increased activity of the metal chelates is due to the lipophilic nature of the metal
ion in complexes. Furthermore, the mode of action of the compounds may involve the formation
of a hydrogen bond through the azomethine nitrogen atom with the active centers of all the
constituents, resulting in interference with normal cell process. The minimum inhibitory
aminoantipyrine and aniline derivatives were determined for five bacteria viz. S. aureus, K.
pneumoniae, S. typhi, P. aeruginosa and B. subtilis by serial dilution. A comparative study of the
ligands and their copper complexes indicates that the copper complexes exhibit slightly higher
25
antibacterial activity than the free ligands and ampicillin21. Such increased activity of the
complexes can be explained on the basis of Overtone’s concept and Tweedy’s concept. Raman et
al105 synthesized a tetraaza macrocyclic Schiff base and its metal complexes derived from 1,2-
1.4.3.1.Structure of DNA
DNA is Deoxyribonucleic acid, the chemistry of life. This nucleic acid is made of long
chain of nucleotides, which are complex molecules present in the nucleus of all cellular forms of
life and many viruses and in the cytoplasm of single celled bacteria which do not have a nucleus.
DNA carrier have a nucleus. DNA carriers along its length has a series of coded chemicals
called genes, which give instructions for passing hereditary characteristics, such as leaf shape,
Each nucleotide consists of 1) a sugar with five carbon atoms, either (a) Deoxyribose,
present in DNA or (b) Ribose, present in RNA. 2) One of the three phosphates with four oxygen
atoms, two of which are negatively charged attached to the 5’ carbon atom of the deoxyribose
sugar and covalently to the 3’ of the next. (One electron from each atom joins two together). 3) A
base: One of four kinds of nucleobases (a base). A base is a ring structure containing nitrogen
and is attached to the 1’ carbon atom of the deoxyribose sugar. The four bases in DNA are
Adenine (A) and Guanine (G) which are purines; Thymine (T) and Cytosine (C) which are
pyrimidines. The most widely accepted model for the structure of DNA molecule was proposed
by Watson and Crick in 1953 for which they were awarded the Nobel Prize for Medicine in
1962. According to the model DNA molecule is a double helix (Figure 1.16). The molecule is
26
formed by two antiparallel polynucleotide strands which are spirally coiled round each other in a
right-handed helix. The two strands are held together by hydrogen bonds. The double stranded
helical structure has alternated major and minor grooves. Each strand is a long polynucleotide of
deoxyribonucleotides. The two strands are complementary to each other with regards to the
arrangement of the bases in the two strands. Thus, in the double helix, purines and pyrimidines
exist in base pairs, i.e., (A and T) and (G and C). As a result, if the base sequence of one strand
of DNA is known, the base sequence of its complementary strand can be easily deduced. The
backbone of the strand is formed by alternately arranged deoxyribose sugar and phosphate
The DNA molecule that Watson and Crick described was in B form. However, DNA can
exist in other form also. A and B forms have right handed helix while Z form has left handed
helix. B is the major form that is found in the cell. The following table summarises the features
of the different forms of DNA and Figure 1.17 shows different forms of DNA.
27
The interaction of small molecules like metal complexes with DNA has been an active
area of research at the interface of chemistry and biology106-111. These small molecules are
stabilized in binding to DNA through a series of weak interactions, such as the p-stacking
interactions associated with intercalation of a planar aromatic group between the base pairs,
hydrogen-bonding and van der Waals interactions of functionalities bound along the groove of
the DNA helix112 and the electrostatic interaction of the cation with phosphate group of DNA113.
Small molecules (metal complexes) bind to DNA double helix by three distinguished
Groove binding
Intercalative binding results when small molecules or the drug intercalate into the
nonpolar interior of the DNA helix. Aromatic group is stacked between the base pairs in this type
of binding and this happens when ligands of an appropriate size and chemical nature fit
themselves in between base pairs of DNA. The ligands suitable for intercalation are mostly
polycyclic, aromatic and polar and therefore often make good nucleic acid stains. There is a
current interest in designing and synthesizing DNA strand, as these molecules might function as
Groove binding interactions involve direct interactions of the bound molecule with edges
of base pairs in either of the major (G-C) or minor (A-T) grooves of the nucleic acids. The
antibiotic netropsin is a model groove binder in which methyl group prevents intercalation116.
Binding within the major groove of the double helix is rare for small molecules (Figure 1.19).
29
molecules. They electrostatically interact with the negatively charged phosphate backbone of
DNA chain. Electrostatic attraction is generally weak under physiological conditions. Cations
Varying the substitutive group or substituent position in the intercalative ligand can
create some interesting differences in the space configuration and the electron density
distribution of transition metal complexes, which will result in some differences in spectral
properties and the DNA binding behavior of the complexes and will be helpful to clearly
DNA cleavage by metal complexes generally proceeds via two major pathways:
oxidative pathway and hydrolytic pathway. The DNA cleavage activity of metal complexes can
be targeted towards different constituents of DNA: the heterocyclic bases, deoxyribose sugar
moiety and phosphodiester linkage. Oxidative cleavage of DNA takes place in the presence of
Many metal complexes have been studied to understand their capability in the hydrolytic
required for hydrolytic cleavage of phosphodiester bond due to unusual stability of the diester
bond in DNA. Among several types of DNA cleavage reactions, those occurring under
The reagents showing photo induced DNA cleavage have major advantage over chemical
nucleases, as the latter requires a reducing agent or H2O2 for its activity. The reagents cleaving
DNA on photoactivation generally show localized effect as they are otherwise non toxic and
such compounds should be useful in the photodynamic therapy (PDT), which has emerged as a
promising tool against cancer. The Food & Drug Administration (FDA) approved PDT drug
photophrin, which is a mixture of hematoporphyrin derivatives and is currently used for the
The design of molecules that exhibit strong binding affinity to DNA is a challenging area
of research. Such molecules can act as excellent chemotherapeutic reagents that exert their
biological activity through interactions with DNA120-125. Interactions with DNA are not the only
31
factors that determine the biological activity of these molecules, but their reactivity and
selectivity are often correlated with their mode of binding with DNA. Therefore, a better
understanding of the factors that govern the interactions of small molecules with DNA is an
important in the rational design of various DNA-targeted chemotherapeutic agents and molecular
probes for DNA. Of these small molecules, the bifunctional derivatives that can undergo photo
induced electron-transfer processes have attracted much attention in recent years for their use in
DNA detection, analysis and cleavage126-131. Stable and inert complexes containing active metal
having potential NNO-tridentate donor Schiff bases derived from the condensation of 2,6-
aminopyridine and 1,10-phenanthroline. Cu(II) complexes of this Schiff base interact with native
calf thymus DNA by groove or intercalating binding mode138. Binuclear copper(II) complexes
propanediamine, are found to be effective in the cleavage of plasmid DNA in the presence of
H2O2 at pH = 7.2 and 37°C. Cu(II) Schiff base complexes derived from diethylenetriamine and
with DNA through a simple mode of coordination139. Copper (II) complex of Schiff bases
activity140. Complete cleavage of double stranded pUC19 DNA by the Cu(II) complex.
2+
N N H2O N N
Cu
N N N N
found to be intercalators of DNA due to the extensively π-conjugated and planar structure of this
carbohydrazone interact with calf-thymus DNA through groove binding143. Metal complexes of
copper, cobalt, nickel and zinc are used now-a-days to study the metal complex-DNA
interactions.
Nair et al studied DNA cleavage with Schiff base derived from indole-3-carboxaldehyde
and m-amino benzoic acid/glycylglycine144,145. They reported metal complexes of Schiff base
Cu(II) ion this ligand undergoes hydrolytic cleavage to form ethylenediamine copper(II)
A knowledge of interaction of DNA with small molecules are basic in the design of new
types of pharmaceutical products. Some metal complexes interact with DNA and induce
breakage of DNA strands. Thus, to cancer genes, after DNA strand cleaves, the DNA double
strands break. The replication ability of the cancer gene is destroyed. Many complexes including
the platinum group have been synthesized and tested in a number of biological systems after the
been studied in view of their possibility to lead to advanced functional materials, tuning the
redox potentials, affinity towards DNA, and specificity for the DNA base sequence recognition.
DNA cleavage was conducted using CT DNA by gel electrophoresis with the corresponding
metal complexes in the presence of H2O2 as oxidant. Raman et al carried out CT DNA cleavage
studies of 4-aminoantipyrine derived Schiff bases and their complexes82. Cu(II), Ni(II) and
Co(II) complexes cleave DNA as compared to control DNA, while Mn(II), Zn(II), VO(IV),
Hg(II) and Cd(II) complexes do not cleave DNA in the presence of H2O2. System comprising,
H2O2-Zn(II) complex were studied under the same conditions using electrophoresis and the result
reveals that cleavage of DNA in Cu(II) system is more efficient than other systems. All the other
systems showed the same electrophoretic behavior and less cleavage activity against CT DNA147.
Similar investigations were made with three oxovanadium(IV) Schiff-base complexes52 viz:
along with reference DNA. DNA-H2O2-1a complex, DNA-H2O2-1b complex and DNA-H2O2-1c
complex were prepared under the same conditions. Damage of DNA in 1a was attributed to the
cleavage of DNA. The 1b and 1c systems showed the same electrophoretic behavior and less
cleavage activity against CT DNA. The available literature reveals that the cleavage efficiency of
the complexes compared to controls is due to their efficient DNA–binding ability. The metal
complexes are able to convert super coiled DNA into open circular DNA. General oxidative
mechanisms proposed for DNA cleavage by hydroxyl radicals via abstraction of a hydrogen
from sugar units predict the release of specific residues arising from transformed sugars,
depending on the position from which the hydrogen is removed148. The cleavage is inhibited by
34
free radical scavengers implying that hydroxyl radical or peroxy derivatives mediate the
species generated from the co-reactant H2O2. Raman et al reported plasmid pUC19 super coiled
(SC) DNA cleavage studies of metal complexes derived from 3,4-dimethoxybenzaldehyde with
Jayabalakrishnan et al151 reported the synthesis and DNA cleavage studies of three hexa-
PPh3/AsPh3/py and L = monobasic tridentate Schiff base ligand derived by the condensation of
REFERENCES
2. Z. Li, K.R. Conser and E.N. Jacobson, J. Am. Chem. Soc., 115, 5326 (1993).
6. M.N. Desai, J.D. Talati and N.K. Shah, Anti-Corrosion Methods and Materials 55, 27
(2008).
8. M. Kojima, H. Taguchi, M. Tsuchimoto and K. Nakajima, Coord. Chem. Rev., 237, 183
(2003).
9. A. Syamal and M.R. Maurya, Coord. Chem. Rev., 95, 183 (1989).
10. J. Costamage, J. Vargas, R. Latorre, A. Alvarado and G. Mena, Coord. Chem. Rev.,119,
67 (1992).
11. P.A. Vigato and S. Tamburini, Coord. Chem. Rev., 248, 1717 (2004).
12. A.A. Solimann and W. Linert, Montash. Chem., 138, 175 (2007).
13. E. Kim, E.E. Chufan, K. Kamraj and K.D. Karlin, Chem. Rev., 104, 1077 (2004).
14. J.R.J. Sorenson, Progress in Medicinal Chemistry, G.P. Ellis and G.B. West, Ed.,
15. R.H. Holm, P. Kennepohl and E. I. Solomon, Chem. Rev., 96, 2239 (1996).
36
16. K.D. Karlin and A.D. Zuber buhler, Bioinorganic Catalysis, 2nd Ed., J. Reediijk, E.
17. S. Srinivasan, J. Annaraj and P.R. Athappan, J. Inorg. Biochem., 99, 876 (2005).
18. D.H. Busch and N.W. Alcock, Chem. Rev., 94, 585 (1994).
19. J-L. Tian, L. Feng, W. Gu, G-J. Xu, S-P. Yan, D-Z. Liao, Z-H. Jiang and P. Cheng, J.
21. N. Raman, A. Kulandaisamy and C. Thangaraja. Transition Met. Chem., 29, 129 (2004).
23. N. Raman, S. Johnson Raja, J. Joseph and J. Dhaveethu Raja. J. Chil. Chem. Soc., 52,
1138 (2007).
24. S. Guru and D.V. Rao. J. Indian Chem. Soc., 45, 160 (1968).
25. J. Gopalakrishnan, C.C. Patel and A. Ravi, Bull. Chem. Soc. Jpn., 40, 791 (1967).
26. K.S. Bose and C.C. Patel, Indian J. Chem., 8A, 557 (1970).
27. C.P. Prabhakaran and C.C. Patel, Indian J. Chem., 10A, 438 (1972).
28. R.K. Agarwal and S. Prasad, J. Iran. Chem. Soc., 2, 168 (2005).
31. V.V. Savant, P. Ramamurthy and C.C. Patel, J. Less Common Met., 22, 474 (1970).
32. M.L.H. Nair and C.P. Radhakrishnan, Indian J. Chem., 39A, 989 (2000).
33. M. Thomas, M.K.M. Nair and P.K. Radhakrishnan, Synth. React. Inorg. Met.-Org.
34. J.N. Gadre, M. Mulay and C. Vaze, Indian J. Heterocycl. Chem., 13, 335 (2004).
37
35. R.K. Agarwal, H. Agarwal and A.K. Manglik, Synth. React. Inorg. Met.–Org. Chem., 26,
1163 (1996).
36. A.Y. Deshmukh, P.B. Rahuwanshi and A.G. Dashi, Asian J. Chem., 14, 185 (2002).
37. K.Z. Ismail, A.E. Dissouky and A.Z. Shehada, Polyhedron, 16, 2909 (1997).
38. G.G. Mohamed, M.A. Zayed, F.A. Nour El-Dien and R.G. El-Nahas, Spectrochemica
39. R.K. Agarwal, R. Garg and S.K. Sindhu, Bull. Chem. Soc. Ethiopia, 19, 185 (2005).
40. S. Issaadi, D. Haffar, T. Douadi, S. Chafaa, D. Se´raphin, M.A. Khan and G. Bouet.
42. R.M. Issa, A.M. Khedr and H.F. Rizk, Spectrochemica Acta Part A, 62, 621 (2005).
43. A.M. Donia and E.M. Ebeid. Thermochim. Acta, 131, 1 (1988).
44. H. Liang, Q. Yu, R. Xiang Hu, Z. Yuan Zhou and X. Ge Zhou, Transition Met. Chem.,
45. M. Alaudeen, P.G. Sushama and A.M. Dorothy, Indian J. Chem., 42A, 1617 (2003).
47. R.K. Agarwal, N. Goel and A.K. Sharma, J. Indian Chem. Soc., 78, 39 (2001).
48. R.K. Agarwal, L. Singh and D.K. Sharma, Bioinorg. Chem. Appl., 2006, 1 (2006).
49. S. Senapati, S.K. Jasimuddin and C. Sinha, Indian J. Chem., 45A, 1153 (2006).
50. T. Rosu, S. Pasculescu, V. Lazar, C. Chifiriuc and R. Cernat, Molecules, 11, 904 (2006).
52. N. Raman, S.J. Raja, J. Joseph and J.D. Raja, Russian J. Coord. Chem., 33, 7 (2007).
53. H.K. Nair and S. Thomas, Asian J. Chem., 17, 2657 (2005).
54. H.K. Nair and G. Mathew, Asian J. Chem., 17, 323 (2005).
55. H.K. Nair and G. Mathew, Asian J. Chem., 17, 1729 (2005).
56. H.K. Nair, G. Mathew and M.R. Sudarsana Kumar, Indian J. Chem., 44A, 85 (2005).
57. A.K. El-Sawaf, D.X. West, F.A. El-Saied and R.M. El-Bahnasawy, Transition Met.
58. S.N. Bhattacharya and K. Meena, Indian J. Chem., 16A, 272 (1978).
59. R.A. Bailey and T.R. Peterson, Can. J. Chem., 47, 168 (1969).
60. G. Shankar and S.K. Ramalingam, Transition Met. Chem., 9, 449 (1984).
61. V.N. Krishnamurthy and S. Soundarrajan, Proc. Indian Acad. Sci., 65, 148 (1967).
62. C.P. Prabhakaran and C.C. Patel, J. Inorg. Nucl. Chem., 30, 867 (1968).
63. G.S. Kurdekar, M.P. Sathisha, S. Budagumpi, N.V. Kulkarni, V.K. Revankar and D.K.
64. M. S. Suresh and V. Prakash, Int. J. Phy. Sci., 5(14), 2203 (2010).
65. M.J. Kharodawala and A.K. Rana, Asian J. Chem., 14, 967 (2002).
66. N. Raman, C. Thangaraja and S. Johnson Raja, Central Euro. J. Chem., 3, 537 (2005).
68. J.R. Chopra, D. Uppal, U.S. Arora and S.K. Gupta, Asian J. Chem., 12, 1277 (2000).
69. G. G. Mohamed, M.M. Omar and A.A. Ibrahim, Eur.J. Med. Chem., 44, 4801 (2009).
70. M.M. Omar, G. G. Mohamed and A.A. Ibrahim, Spectrochimica Acta Part A, 73, 358
(2009).
71. S. Chandra, D. Jain, A.K. Sharma and P. Sharma, molecules, 14, 174 (2009).
39
72. N. Raman, A. Kulandaisamy and K. Jeyasubramanian, Indian J. Chem., 41A, 942 (2002).
74. K. Nagashria, J. Josepha and C. Justin Dhanaraj, Appl. Organometal. Chem., 25, 704
(2011).
76. N. Raman and S. Sobha, Spectrochimica Acta Part A, 85(1), 223 (2012).
78. N. Raman, S. Sobha and L. Mitu, Journal of Saudi Chemical Society, Article In Press.
79. J. Joseph, K. Nagashri and G.A.B. Rani. Journal of Saudi Chemical Society, Article In
Press.
80. N. Raman, S. Syed Ali Fathima and J. Dhaveethu Raja, J. Serb. Chem. Soc., 73(11), 1063
(2008).
81. N. Raman, S. Thalamuthu, J. Dhaveethu Raja, M.A. Neelakandan and Sharmila Banerjee.
82. N. Raman, J. Dhaveethu Raja and A. Sakthivel, J. Chem. Sci., 119(4), 303 (2007).
87. N. Raman, A. Selvan, P. Manisankar, Spectrochimica Acta Part A, 76, 161 (2010).
88. M.R. Bermejo, A.M. Gonzalez-Noya, V. Abad, M.I. Fernandez, M. Manerio, R. Pedrido
89. H. Z. Kou, Z. H. Ni, B. C. Zhou and R. J. Wang, Inorg. Chem. Comm., 7, 1150 (2004).
90. A. D. Garnovskii, I.S. Vasilchenko, D.A. Garnovskii and B.I. Kharisov, J. Coord. Chem.,
91. N. Raman, S. J. Raja and A. Sakthivel, J. Coord. Chem., 62, 691 (2009).
92. M.S. Ibrahim and S.E.H. Etaiw, Synth. React. Inorg. Met.-Org. Chem., 34, 629 (2004).
93. A. Lalehzari, J. Desper and C.J. Levy, Inorg. Chem., 47, 1120 (2008).
95. M.S. Reft, I.M. El-Deen, Z.M. Anwer and S. El-Chol, J. Coord. Chem., 62, 1709 (2009).
97. J. Ziegler, T. Schuerle, L. Pasierb, C. Kelly, A. Elamin, K.A. Cole and D.W. Wright,
98. M.S. Nair and R.S. Joseyphus, Spectrochimica Acta Part A, 70, 749 (2008).
99. R. Johari, G. Kumar, D. Kumar and S. Singh, J. Ind. Council. Chem., 26(1), 1 (2009).
100. G.G. Mohamed, M.M. Omar and A.M. Hindy, Turk. J. Chem., 30, 361 (2006).
101. S. Zhang, Y. Jiang and M. Chen, Chem. Abstr., 142, 347 (2005).
102. N. Pal Singh and A.N. Srivastava, J. Serb. Chem. Soc., 77(5) 627 (2012).
103. B. K. Singh, P. Mishra, A. Prakash and N. Bhojak, Arab. J. Chem., Artcile In Press
41
104. B.K. Keppler, C. Friesen, H.G. Moritz, H. Vongerichten and E. Vogel, Struct. Bond., 78,
97 (1991).
106. P. Nordell and P. Lincoln, J. Am. Chem. Soc., 127, 9670 (2005)
107. C.C. Cheng, W.C.H. Fu, K.C. Hung, P.J. Chen, W.J. Wang and Y.T. Chen, Nucl. Acid
108. A.A. Mokhir and R. Kraemer, Bioconjugate Chem., 14, 877 (2003).
109. W.C. Tse and D.L. Boger, Acc. Chem. Res., 37, 61 (2004).
110. P. Yang, R. Ren, M.L. Guo, A.X. Song, X.L. Meng, C.X. Yuan, Q.H. Zhou, H.L. Chen,
Z.H. Xiong and X.L. Gao, J. Biol. Inorg. Chem., 9, 495 (2004).
111. A.Th. Chaviara, P.J. Cox, K.H. Repana, A.A. Pantazaki, K.T. Papazisis, A.H. Kortsaris,
D.A. Kyriakidis, G.S. Nikolov and C.A. Bolos, J. Inorg. Biochem., 99, 467 (2005).
112. A.M. Pyle, J.P. Rehmann, R. Meshoyrer, C.V. Kumar, N.J. Turro and J.K. Barton, J. Am.
113. S. Satyanarayana, J.C. Dabrowiak and J.B. Chaires, Biochemistry, 31, 9319 (1992).
114. H. Mei and J. Barton, J. Am. Chem. Soc., 108, 7414 (1986).
115. J. Kelly, A. Tossi and D. McComel, Nucl. Acids. Res., 13, 6017 (1985)
117. L-N. Ji, X-H. Zou and J-G. Liu, Coord. Chem. Rev., 216, 513 (2001).
118. C. Metcalfe and J.A. Thomas, Chem. Soc. Rev., 32, 215 (2003).
121. L. Pecq, J.B. M.L. Bret, J. Barbet and B. Roques, Proc. Natl. Acad. Sci. U.S.A. 72, 2915
(1975).
123. W.A. Denny and L.P.G. Wakelin, Anti-Cancer Drug Des., 5, 189 (1990).
124. P.R. Reddy and A. Shilpa, Indian J. Chem., 49, 1003 (2010).
Small Molecule DNA and RNA Binders: From Synthesis to NucleicAcid Complexes; 1,
2 (2002).
126. E. Karuvilla, J. Joseph and D. Ramaiah, J. Phys. Chem. B, 109, 21997 (2005).
127. H. Morrison, Ed.; John Wiley and Sons: New York, 1, 273 (1990).
128. J.E. Rogers, T.P. Le and L.A. Kelly, Photochem. Photobiol., 73, 223 (2001).
129. J. Joseph, N.V. Eldho and D.J. Ramaiah, Phys. Chem. B, 107, 4444 (2003).
130. J. Joseph, N.V. Eldho and D.J. Ramaiah, Chem- Eur. J, 9, 5926 (2003).
131. B. Akerman and E. Tuite, Nucleic Acids Res., 24, 1080 (1996).
132. S.K. Gupta and D.D. Agarwal and D. Raina, Ind. J. Chem., 35A, 995 (1996).
133. S.K. Gupta and D. Raina, Transition Met. Chem., 22, 225 (1997).
134. S.K. Gupta and D. Raina, Transition Met. Chem., 22, 372 (1997).
135. S.K. Gupta, D.D. Agarwal and D. Raina, Ind. J. Chem., 38A, 506 (1999).
136. S.K. Gupta, P.B. Hitchcock and Y.S. Kushwah, J. Coord. Chem., 55, 1401(2002).
137. S.K. Gupta and Y.S. Kushwah, Polyhedron, 20, 2019 (2001).
138. O.B. Saif, A.A. Abdel Aziz and M.E. El-Nagger, J. Mol. Struc., 1020, 188 (2012).
139. A.T. Chaviara, E.E. Koiseoglou, A.A. Pantazaki, A.C. Tsipis, P.A. Karipidis, D.A.
140. P. Lumme, E. Honnu and J. Jubani, Inorg. Chim. Acta., 92, 241 (1984).
141. S. Dhar, D. Senapati, P.A.N. Reddy, P.K. Das and A.R. Chakravarthy, Chem. Commun.,
2452 (2003).
143. Z.H. Xu, F.J. Chen, P.X. Xi, X.H. Liu and Z.Z. Zeng, Journal of Photochemistry and
144. M.S. Nair, D. Arish and R.S. Joseyphus, Journal of Saudi Chemical Society, 12, 83
(2012).
145. R.S. Joseyphus and M.S. Nair, Arab. J. Chem., 3, 195 (2010).
146. D. Arish and M.S. Nair, Spectrochimica Acta Part A, 82, 191 (2011).
147. N. Raman and S. Johnson Raja, J. Serb. Chem. Soc., 72, 983 (2007).
148. G. Prativel, M. Pitie, J. Bernadou and B. Meunier, Angew. Chem. Int. Ed. Eng., 30, 702
(1991).
149. G. Psomas, A. Tarushi and E.K. Efthimiadou, Polyhedron, 27, 133 (2008).
150. N. Raman, S. Sobha and A. Thamaraichelvan. Spectrochimica Acta Part A, 78, 888
(2011).
151. G. Raja, R.J. Butcher and C. Jayabalakrishnan. Spectrochimica Acta Part A, 94, 210
(2012).