27
CHAPTER I General Introduction The area of inorganic chemistry, which most widely developed in the last few decades is mainly due to coordination chemistry and applies very particularly to the coordination compounds of transition metals. The chemistry of coordination compounds has always been a challenge to the inorganic chemists as it has more branches now-a- days. Coordination compounds play a very significant role in our lives, the study of them has contributed to the highest degree of understanding the chemical bond in inorganic chemistry. As a whole classical coordination chemistry deals with the formation of adducts by metal in their higher oxidation states bonded to inorganic or organic ions or molecules. Interest in both basic and hi-tech research with these materials continues at a rapid pace. Metals play a vital role in an immense number of extensively differing biological processes. Some of these processes are quite specific in their metal ion requirements, in that only certain metal ions in specified oxidation states can accomplish the necessary catalytic structural requirement. Metal ion dependent processes are found through out the life science and vary tremendously in their function and complexity. It is now appreciated that metal ions control a vast range of processes in biology. Many new and exciting developments in the field of biochemistry create interest out of inorganic chemists to court in the new area called “Bioinorganic Chemistry”. 1

CHAPTER I General Introductionshodhganga.inflibnet.ac.in/bitstream/10603/39/3/chapter 1_puthilibai.pdf · Figure 1.1 Structure of the bidentate Schiff bases. L-Amino acid Schiff bases

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • CHAPTER I

    General Introduction

    The area of inorganic chemistry, which most widely developed in the last few

    decades is mainly due to coordination chemistry and applies very particularly to the

    coordination compounds of transition metals. The chemistry of coordination compounds

    has always been a challenge to the inorganic chemists as it has more branches now-a-

    days.

    Coordination compounds play a very significant role in our lives, the study of

    them has contributed to the highest degree of understanding the chemical bond in

    inorganic chemistry. As a whole classical coordination chemistry deals with the

    formation of adducts by metal in their higher oxidation states bonded to inorganic or

    organic ions or molecules. Interest in both basic and hi-tech research with these materials

    continues at a rapid pace.

    Metals play a vital role in an immense number of extensively differing biological

    processes. Some of these processes are quite specific in their metal ion requirements, in

    that only certain metal ions in specified oxidation states can accomplish the necessary

    catalytic structural requirement. Metal ion dependent processes are found through out the

    life science and vary tremendously in their function and complexity.

    It is now appreciated that metal ions control a vast range of processes in biology.

    Many new and exciting developments in the field of biochemistry create interest out of

    inorganic chemists to court in the new area called “Bioinorganic Chemistry”.

    1

  • One of the principal themes of bioinorganic chemistry is the synthesis of metal

    complexes that have the ability to mimic the functional properties of natural

    metalloproteins [1,2]. Proteins, some vitamins and enzymes contain metal ions in their

    structure involving macromolecular ligands. The chemistry of metal complexes with

    multidentate ligands having delocalized π-orbitals, such as Schiff bases or porphyrins has

    recently gained more attention because of their use as models in biological systems.

    Schiff bases

    The condensation of primary amines with aldehydes and ketones give imines.

    Imines that contain an aryl group bound to the nitrogen or to the carbon atom are called

    Schiff bases, since their synthesis was first reported by Schiff [3].

    Schiff bases are capable of forming coordinate bonds with many of metal ions

    through both azomethine group and phenolic group or via its azomethine or phenolic

    groups [4-19]. A large number of Schiff bases and their complexes are significant interest

    and attention because of their biological activity including anti-tumor, antibacterial,

    fungicidal and anti-carcinogenic properties [7-12] and catalytic activity [12-19].

    Naphthylideneimine Schiff base complexes (Figure 1.1) possessing luminescence

    property, catalyze oxidation of primary and secondary alcohols into their corresponding

    carbonyl compounds in the presence of N-methylmorpholine-N-oxide (NMO) as the

    source of oxygen have been reported recently [14]. The formation of high valent RuIV= O

    species as a catalytic intermediate is proposed for the catalytic process.

    2

  • Figure 1.1 Structure of the bidentate Schiff bases.

    L-Amino acid Schiff bases with N,O donor system have been reported by Taqui

    Khan et al [20]. and are used as catalyst of enantio selective epoxide of 1,2-di hydro-

    naphthalene. Nitro substituted benzaldehyde Schiff bases were used in organic catalytic

    reactions [21]. Schiff bases of N-methyl and N-acetyl isatin derivatives with different

    arylamines have been prepared and screened for anti convulsant activities [22].

    Antibacterial screening of monobasic bidentate Schiff base complexes

    (Figure 1.2) with N,O donor have been reported [23].

    C N

    H R

    OH

    OCH3

    (R = -CH3, -C5H4N, -C6H12)

    Figure 1.2 Structure of monobasic bidentate Schiff base complexes with N,O donor.

    3

  • Schiff bases of ethylenediamine/triethylenetetramine (salen) with benzaldehyde/

    cinnamic aldehyde/salicylaldehyde as corrosion inhibitors of zinc in sulphuric acid have

    been reported by Desai et al [24].

    Recent report says, Schiff bases (Figure 1.3) are also employed as fluorescent

    indicators by spectrofluorimetric monitoring of small changes of pH [25].

    Figure 1.3 Structure of fluorescent Schiff bases.

    Transition Metal complexes

    The phenomenon of complex formation is really a very general one, but is

    especially noted among the transition metal ions. For bonding, the metal must posses

    vacant orbitals and these orbitals symmetrically must be correct, sterically available and

    4

  • of reasonably low energy. Since transition metal ions generally meet these requirements

    best, it is not surprising that they form complexes readily.

    The transition elements play vital role in coordination complexes mainly because

    of the following characteristics [26-29]:

    variable oxidation state (electron transfer properties),

    coordination geometries (octahedral, tetrahedral, square planar, pyramidal, etc.),

    spectral and magnetic features, ligand field effects, unpaired d-electrons,

    formation of chelated complexes,

    most M2+ and higher oxidation states are borderline or hard acids and generally

    prefer borderline or hard base such as O and N-donor groups; lower oxidation states

    e.g. Cu(I) are softer acid will bind the soft bases such as O2, CO, N2, and S. and

    formation of polynuclear metal species e.g. dimers, tetramers with bridging.

    Metal coordination complexes have a wide diversity of technological and

    industrial applications ranging from catalysis to anticancer drugs [30].

    Ruthenium complexes

    Ruthenium Schiff base complexes, particularly those containing oxygen and

    nitrogen as donor atoms were found to be very efficient catalysts in the oxidation of

    alcohols and alkenes. Electron transfer reactions are fundamental and play important role

    in chemical and biological processes. As the coordination environment around the central

    metal ion directs properties of the complexes, complexation of ruthenium by ligands of

    different types has been of significant importance.

    5

  • Ruthenium(II) complexes

    Ruthenium(II) compounds display long luminescence life time and are extremely

    photo stable [31,32].

    There is a continuing interest in the development of probes for nucleic acid

    structure determination. In recent years, a number of metal chelates have been used as

    DNA structural probes and as chemotherapeutic agents. In particular, ruthenium(II)

    complexes of the type [Ru(LL)3]n+, where LL is a bidentate ligands of various nature,

    have been extensively studied as probes for the determination of nucleic acid

    structure.[33-38] The application of these complexes as DNA structural probes is due to

    their water solubility, coordinatively saturated nature and substitution inertness.

    Catlytic activity and antibacterial screening of Ru(II) Schiff base complexes of

    the type [Ru(CO)(EPh3)(B)(L)] (E = P or As; B= PPh3, AsPh3, py or pip; L = Schiff

    bases have been reported by Balasubramanian et al [39].

    DNA binding and antibacterial screening of dehydroacetic acid complexes

    (Figure 1.4 A&B) of Ru(II) and Ru(III) containing PPh3/AsPh3 have been recently

    reported by Chitrapriya et al [40].

    Figure 1.4(A) Structure of Ru(III) dehydroacetic acid complexes.

    6

  • Figure 1.4(B) Structure of Ru(II) dehydroacetic acid complexes.

    Recent studies indicate that ruthenium complexes (Figure 1.5) are promising

    candidates for NLO materials because of their rich photochemical properties and varied

    coordination form [41-46].

    7

  • Figure 1.5 Structures of the ruthenium(II) complexes possess NLO property.

    Ruthenium(II) complexes containing triphenylphosphine/triphenylarsine and tetra

    dentate Schiff bases are found to be effective catalysts in the oxidation of primary and

    secondary alcohols using N-methylmorpholine-N-oxide as oxidant [47]. The catalytic

    activity of these triphenylarsine complexes have been compared with that of

    triphenylphosphine complexes and with similar ruthenium(III) complexes.

    Ruthenium(III) complexes

    Prabhakaran et al [11] synthesized ruthenium(III) complexes (Figure 1.6)

    containing tetradentate Schiff base to test its antibacterial activity.

    Figure 1.6 Structure of tetradentate Schiff base ligand.

    8

  • Ramesh et al [18] reported catalytic oxidation of primary alcohols by some of the

    ruthenium(III) Schiff base complexes (Figure 1.7) was carried out in CH2Cl2 in the

    presence of N-methylmorpholine-N-oxide.

    Figure 1.7 Ruthenium(III) Schiff base complexes.

    Kannan et al [48] synthesized a series of new ruthenium(III) Schiff base

    complexes (Figure 1.8) incorporating triphenylphosphine/triphenylarsine and chloride/

    bromide ligands, and have been reported as efficient catalyst for the oxidation of both

    primary and secondary alcohols to the corresponding carbonyl compounds with excellent

    yields in the presence of NMO. Further, the possible explanations for the mode of action

    of these complexes against two different microbes S. aureus and E. coli are described.

    Figure 1.8 Structure of Ru(III) Schiff base complexes.

    9

  • Copper(II)/Cobalt(II)/Nickel(II) complexes

    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., contains copper atoms bound to protein molecules.

    Copper(II) complexes with amino acids are cited as having potent anti-

    inflammatory and anti- ulcer activity [49]. Copper ions are found to present in the active

    sties of large number of metalloproteins, which involved in important biological electron

    transfer reactions as well as in the molecular oxygen redox reactions [50,51].

    There has been a substantial interest in the rational design of novel transition metal

    complexes, which bind and cleave duplex DNA with high sequence or structure

    selectivity. The characterization of DNA recognition by small transition-metal complexes

    has been substantially aided by the DNA cleavage chemistry that is associated with

    redox-active or photoactivated metal complexes.

    10

  • Non–covalent interactions between positively charged metallointercalators and

    the base–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 favorable for application in industries and medicine [52,53]. 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 1.9(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 al [54] reported the nuclease activity (DNA cleavage) of Cobalt(II)

    complexes with pBR 322 DNA in the absence of any external agents.

    11

  • (A)

    (B)

    (C)

    Figure 1.9(A-C) Co(II) complexes of porphyrins, Schiff bases and tetraaza systems.

    12

  • The intercalative DNA-Binding studies of a Nickel(II) coordination compound

    Ni(bpy)2dppz2+ have been reported [55]. The calculated intrinsic binding constant for the

    same is 1.5 × 104 M−1. These features are equivalent to those observed with

    Ru(bpy)2dppz2+ and suggest that nickel complex binds by intercalation in a manner that

    parallels Ru(bpy)2dppz2+ [56].

    Square planar nickel(II) complexes were studied owing to their known catalytic

    activity towards olefin epoxidation.

    Biological studies

    Nucleic acids viz. ribonucleic acid (RNA) and deoxyribonucleic acid (DNA),

    contain three types of basic structural nucleotide units: (a) pentose sugar,

    (b) nitrogenous base (pyrimidins/purines) and (c) phosphate residue. They are long,

    thread-like polymers, consisting of a linear array of monomers called nucleotides.

    Nucleotides are the phosphate ester of nucleosides, which are the basic components of

    DNA. The stacks of DNA contain A, G, C, T while RNA contains A, G, C, U. The types

    of pentose also distinguish the nucleic acids 2-Deoxyribose is found in DNA while it is

    ribose in RNA.

    DNA usually consists of two complementary polymeric chains twisted about each

    other in the form of a right-handed helix, making a complete turn every 34 0A (3.4 nm),

    with a diameter of 20 0A (2 nm). Since the distance between adjacent nucleotides is

    3.4 0A, there must be 10 nucleotides per turn. The constant diameter of the helix can be

    explained if the bases in each chain face inward and are restricted so that a purine is

    13

  • always opposite to a pyrimidine avoiding organization of purine-purine (too thick) or

    pyrimidine-pyrimidine (too thin). Irrespective of the actual amounts of each base, the

    proportion of G is always the same as the proportion of C in DNA, and the proportion of

    A is always the same as that of T. The composition of any DNA can be described by the

    proportion of its bases ic., G+C, which ranges from 26% to 74% for different species.

    The two chains of DNA (Figure 1.10) are complementary to each other.

    Figure 1.10 Structure of right handed double helical DNA (B-Type).

    14

  • Since DNA is the basis for the storage, transmission and expression of genetic

    information, any reaction or damage caused to it will have important consequences.

    There has been substantial interest in exploring the factors that determine kinship and

    selectivity in binding of small molecules with DNA [57-59]. A quantitative

    understanding of such factors that determine recognition of DNA sites would be valuable

    in designing small molecules that binds to specific sites in DNA and finds application in

    chemotherapy. A number of metal chelates mostly polypyridyls have been used as

    probes of DNA structure in solution [60], as agents for mediation of strand scission of

    duplex DNA [61] and as chemotherapeutic agents [62].

    The interaction of small molecules like metal complexes with DNA has been an

    active area of research at the interface of chemistry and biology [63-68]. 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 helix [69], and the electrostatic

    interaction of the cation with phosphate group of DNA [70]. Studies directed toward the

    design of site- and conformation- specific reagents provide rationales for new drug design

    as well as a means to develop sensitive chemical probes of nucleic acid structure.

    Small molecules (metal complexes) bind to DNA double helix by three

    distinguished binding modes Figure 1.11. They are,

    A. Electrostatic binding / External binding

    15

  • B. Groove binding

    C. Intercalative binding / Intercalation

    A. External binding / Electrostatic binding

    Complexes are positively charged and the DNA phosphate sugar backbone is

    negatively charged and their interaction is known as electrostatic. This association mode

    was proposed for [Ru(bpy)3]2+, due to the luminescence enhancement of this complex

    upon binding to DNA. Cations such as Mg2+ usually interact in this way [71].

    B. Groove binding

    The molecules approach within van der Waals contact and reside in the DNA

    groove. Hydrophobic and/or hydrogen-bonding are usually important components of this

    binding process, and provide stabilization. The antibiotic netropsin is a model groove-

    binder where the methyl groups prevent intercalation [72].

    16

  • C. Intercalation

    This association involves the insertion of a planar fused aromatic ring system

    between the DNA base pairs, leading to significant �-electron overlap. Stacking

    interactions stabilizes this mode of binding and is thus less sensitive to ionic strength

    relative to the two other binding modes. This mode of binding is usually favored by the

    presence of an extended fused aromatic ligand like DPPZ [73]. Indeed with less extended

    aromatic systems, the intercalation is usually prevented through clashing of the ancillary

    ligands with the phosphodiester backbone, so that only partial intercalation can occur as

    in the case for [Ru(phen)3]2+ [74].

    17

  • Figure 1.11 Binding modes of small molecules with DNA.

    Many metal complexes can bind to DNA in noncovalent modes such as

    electrostatic, intercalative and groove binding [75,76]. 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 behaviors of the complexes and will be helpful to more clearly understand the

    binding mechanism of transition metal complexes to DNA [77,78].

    18

  • Scope of the present work

    Schiff bases, a class of chelators, considered as privileged ligands, attractive due

    to their stability, the ease by which modified variations can be obtained, a diverse range

    of applications and are flexible both in terms of size and charge.

    Schiff base complexes of transition metals having O and N donor atoms

    containing phosphine/arsine especially ruthenium complexes, find application in classical

    catalytic processes such as hydrogenation, isomerisation, decarbonylation, reductive

    elimination, oxidative addition and in making C–C bonds. Transition metal carbonyl

    complexes are reactive species in homogeneous catalytic reactions such as

    hydrogenation, hydroformylation and carbonylation.

    Literature says, though there is a considerable growth of Schiff base complexes

    of first row transition metals, the chemistry of ruthenium complexes is less well

    developed. But, there has been considerable interest in ruthenium complexes now-a-days

    because of their redox stability, excited state life time, excited state reactivities their

    ability to act as probes in investigating the structure of DNA and its antimicrobial

    activity. The interaction of ruthenium complexes with DNA has received a great deal of

    attention during the past decade. Also metal complexes of ruthenium containing nitrogen

    and oxygen donor ligands are found to be very effective catalysts for oxidation,

    reduction, hydrolysis and other organic and inorganic transformations especially

    regioselective oxidation. The redox property of the central ruthenium can be tuned by

    changing the substituents in the ligand. The oxidation states of ruthenium complexes can

    vary from –II to +VIII. The chemotherapeutic Schiff bases are now attracting the

    19

  • attention of inorganic chemists. Reports show that some drugs show increased activity

    when it is administered as metal complexes rather than as organic compounds. Yet the

    mechanism of antitumour activity of ruthenium compounds is not fully understood, it is

    believed that, similar to platinium drugs, the chloride complexes can hydrolyze in vivo,

    allowing the Ru to bind to the nucleobases of the DNA. However a deep survey of

    literature on Schiff base transition metal complexes of N and O donors reveal that

    antimicrobial and DNA binding studies have been largely ignored.

    In the present project, which has originated from our interest in the chemistry of

    transition metals especially ruthenium in different coordination environments. We have

    chosen different types of phenolic ligands to synthesis, characterize and to study their

    antibacterial activity and DNA binding properties.

    20

  • References

    [1] Z. Lu, L. Yang, J. Inorg. Biochem. 95 (2003) 31.

    [2] J.Z. Wu, H.Li, J.G. Zhang, H. Xu Ju, Inorg. Chem. Commun. 5 (2002) 71.

    [3] H. Schiff, Ann. Chim.(Paris) 131 (1864) 118.

    [4] S.A. Sadeek, M.S. Refat, J. Korean Chem. Soc. 50 (2) (2006) 107.

    [5] A. Bigotto, V. Galasso, G. Dealti, Spectrochim. Acta 28A (1972) 1581.

    [6] S. Ren, R. Wang, K. Komatsu, P. Bonaz-Krause, Y. Zyrianov, C.E. Mckenna,

    C. Csipke, Z.A. Tokes, E.J. Lien, J. Med. Chem. 45 (2002) 410.

    [7] N. Chitrapriya, V. Mahalingam, L.C. Channels, M. Zeller, F.R. Fronczek,

    K. Natarajan, Inorg. Chim. Acta 361 (2008) 2841.

    [7] M.S. Refat, S.A. El-Korashy, D.N. Kumar, A.S. Ahmed, Spectrochim. Acta

    Part A 70 (2007) 898.

    [9] R. Prabhakaran, R. Huang, K. Natarajan, Inorga. Chim. Acta 359 (2006) 3359.

    [10] K.P. Balasubramanian, K. Parameswari, V. Chinnusamy, R. Prabhakaran,

    K. Natarajan, Spectrochim. Acta Part A 65 (2006) 678.

    [11] R. Prabhakaran, A. Geetha, M. Thilagavathi, R. Karvembu, V. Krishnan,

    H. Bertagnolli, K. Natarajan, J. Inorg. Biochem. 98 (2004) 2131.

    21

  • [12] S. Kannan, R. Ramesh, Polyhedron 25 (2006) 3095.

    [13] S. Kannan, K. Naresh Kumar, R. Ramesh, Polyhedron 27 (2008) 701.

    [14] M. Sivagamasundari, R. Ramesh, Spectrochim. Acta Part A 67 (2007) 256.

    [15] S. Kannan, R. Ramesh, Y. Liu, J. Organomet. Chem. 692 (2007) 3380.

    [16] V. Mahalingam, R. Karvembu, V. Chinnusamy, K. Natarajan, Spectrochim. Acta

    Part A 64 (2006) 886.

    [17] G. Venkatachalam, R. Ramesh, Inorg. Chem. Commun. 9 (2006) 703.

    [18] R. Ramesh, Inorg. Chem. Commun. 7 (2004) 274.

    [19] R. Karvembu, S. Hemalatha, R. Prabhakaran, K. Natarajan, Inorga. Chem.

    Commun. 6 (2003) 486.

    [20] M.M. Taqui Khan, R.I. Kureshy, N.H. Khan, Tetrahedron: Asymmetry, 2 (1991)

    101.

    [21] W.T. Gao, Z. Zhing, Molecules 8 (2003) 788.

    [22] M. Verma, S.N Pandeya, K. Singh, T.P. Stables, Acta Pharma. 54 (2004) 49.

    [23] K. Naresh Kumar, R. Ramesh, Spcetrochim. Acta Part A 60 (2004) 2913.

    [24] M.N. Desai, J.D. Talati, N.K. Shah, Anti-Corrosion Methods and Materials 55

    (2008) 27.

    22

  • [25] M.N. Ibrahim, S.E. A. Sharif, E-Journal of Chemistry 4 (4) (2007) 531.

    [26] J.C. Wu, N. Tang, W.S. Liu, M.Y. Tanchan, A.S. Chin, Chem. Lett. 12 (2001)

    757.

    [27] L. He, S.H.Q.F. Gou, J. Chem. Crystallogr. 29 (1999) 207.

    [28] P. Bhattacharya, J. Parr, A.J. Ross, J. Chem. Soc. Dalton (1998) 3149.

    [29] C.M. Liu, R.G. Xiong, X.Z. You, Y.J. Liu, K.K Chiung, Polyhedron 15 (1996)

    4564.

    [30] A.D. Burrows, Science Progress 85 (3) (2002) 199.

    [31] F.N. Castellano, J.D. Dattelbaum, J.R. Lakowicz, Anal. Biochem. 255 (1998) 165.

    [32] H.Wolpher, O. Johansson, M. Abrahamson, M. Kritikos, L. Sun, B. Akermark,

    Inorg. Chem. Commun. 7 (2004) 337.

    [33] T. Urathamkul, J.L. Beck, M.M. Sheil, J.R. Aldrich-Wright, S.F. Ralph, Dalton

    Trans (2004) 2683.

    [34] H. Chao, J.-G. Liu, L.-N. Ji, X.-Y. Li, J. Biol. Inorg. Chem. 6 (2001) 143.

    [35] J.-G. Liu, B.-H. Ye, Q.-L. Zhang, X.-H. Zou, Q.-X. Zhen, X. Tian, L.-N. Ji,

    J. Biol. Inorg. Chem. 5 (2000) 119.

    [36] A. Anagnostopoulou, E. Moldrheim, N. Katsaros, E. Sletten, J. Biol. Inorg. Chem.

    4 (1999) 199.

    23

  • [37] R. Hartmann, A. Robert, V. Duarte, B.K. Keppler, B. Meunier, J. Biol. Inorg.

    Chem. 2 (1997) 143.

    [38] D.B. Hall, R.E. Holmin, J.K. Barton, Nature 382 (1996) 73.

    [39] K.P. Balasubramanian, R. Karvembu, R. Prabhakaran, V. Chinnusamy,

    K. Natarajan, Spectrochim. Acta Part A 68 (2007) 50.

    [40] N. Chitrapriya, V. Mahalingam, M. Zeller, R. Jayabalan, K. Swaminathan,

    K. Natarajan, Polyhedron 27 (2008) 939.

    [41] C. Dhenaut, I. Ledoux, I.D.W. Samuel, J. Zyss, M. Bourgault, H. Le Bozec,

    Nature (London) 374 (1995) 339.

    [42] B.J. Coe, G. Chadwick, S. Houbrechts, A. Persoons, J. Chem. Soc., Dalton Trans.

    (1997) 1705.

    [43] B.J. Coe, S. Houbrechts, I. Asselberghs, A. Persoons, Angew. Chem., Int. Ed.

    Engl. 38 (1999) 366.

    [44] I.R. Whittall, M.P. Cifuentes, M.G. Mark, B. Luther-Davies, M. Samoc,

    S. Houbrechts, A. Persoons, G.A. Heath, D.C.R. Hockless, J. Organomet. Chem.

    549 (1997) 127.

    [45] A.M. McDonagh, M.G. Humphrey, M. Samoc, B. Luther- Davies, S. Houbrechts,

    T. Wadw, H. Sasabe, A. Persoons, J. Am. Chem. Soc. 121 (1999) 1405.

    24

  • [46] H. Chao, R.H. Li, B.H. Ye, H. Li, X.L. Feng, J.W. Cai, J.Y. Zhou, L.N. Ji,

    J. Chem. Soc., Dalton Trans. (1999) 3711.

    [47] R. Karvembu, S. Hemalatha, R. Prabhakaran, K. Natarajan, Inorg. Chem.

    Commun. 6 (2003) 486.

    [48] S. Kannan, R. Ramesh, Polyhedron 25 (2006) 3095.

    [49] J.R.J. Sorenson, Progress in Medicinal Chemistry, G.P. Ellis, and G.B. West,

    Ed., Elsevier, New York, Vol. 26 (1989) 437 and references cited therein.

    [50] R.H. Holm, P. Kennepohl, Solomon, Chem. Rev. 96 (1996) 2239.

    [51] K.D. Karlin, A.D. Zuber buhler, Bioinorganic Catalysis, 2nd Ed., J. Reediijk,

    E. Bouwman, Ed., MarcelDekker, NewYork (1999) 469.

    [52] S. Srinivasan, J. Annaraj, PR. Athappan, J. Inorg. Biochem. 99 (2005) 876.

    [53] D.H. Busch, N.W. Alcock, Chem. Rev. 94 (1994) 585.

    [54] J.-L. Tian, L. Feng, W. Gu, G.-J. Xu, S.-P. Yan, D.-Z. Liao, Z.-H. Jiang,

    P. Cheng, J. Inorg. Biochem. 101 (2007) 196.

    [55] L. Jin, P. Yang, Microchemical Journal 58 (1998) 144.

    [56] A.E. Friedman, J.C. Chambron, J.P. Sauvage, J. Am. Chem. Soc. 112 (1990)

    4960.

    [57] W.S. Wade, P.B. Dervan, J. Am. Chem. Soc. 109 (1987) 1574.

    25

  • [58] J.K. Barton, Science 233 (1986) 727.

    [59] K. Kissinger, J.b. Dabrowiak, J.W. Lown, Biochemisty 26 (1987) 5590.

    [60] J.K. Barton, Inorg. Chem. 7 (1985) 321.

    [61] D.S. Sigman, Acc. Chem. Res. 19 (1986) 180.

    [62] S.J. Lippand, Acc. Chem. Res. 11 (1978) 211.

    [63] P. Nordell, P. Lincoln, J. Am. Chem. Soc. 127 (2005) 9670.

    [64] C.C. Cheng, W.C.H. Fu, K.C. Hung, P.J. Chen, W.J. Wang, Y.T. Chen, Nucl.

    Acid Res. 31 (2003) 2227.

    [65] A.A. Mokhir, R. Kraemer, Bioconjugate Chem. 14 (2003) 877.

    [66] W.C. Tse, D.L. Boger, Acc. Chem. Res. 37 (2004) 61.

    [67] 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, X.L. Gao, J. Biol. Inorg. Chem. 9 (2004) 495.

    [68] A.Th. Chaviara, P.J. Cox, K.H. Repana, A.A. Pantazaki, K.T. Papazisis,

    A.H. Kortsaris, D.A. Kyriakidis, G.S. Nikolov, C.A. Bolos, J. Inorg. Biochem. 99

    (2005) 467.

    [69] A.M. Pyle, J.P. Rehmann, R. Meshoyrer, C.V. Kumar, N.J. Turro, J.K. Barton,

    J. Am. Chem. Soc. 111 (1989) 3051.

    26

  • [70] S. Satyanarayana, J.C. Dabrowiak, J.B. Chaires, Biochemistry 31 (1992) 9319.

    [71] J. Kelly, A.Tossi, D. McComell, Nucl. Acids. Res. 13 (1985) 6017.

    [72] H. Mei, J. Barton, J. Am. Chem. Soc. 108 (1986) 7414.

    [73] I. Haq, P. Lincoln, D. Suh, B. Norden, B. Chowdhry, J. Chaires, J. Am. Chem.

    Soc. 117 (1995) 4788.

    [74] P. Lincoln, B. Norden, J. Phys. Chem. B. 102 (1998) 9583.

    [75] J.A. Cowan, Curr. Opin. Chem. Biol. 5 (2001) 634.

    [76] L.-N. Ji, X.-H. Zou, J.-G. Liu, Coord. Chem. Rev. 216 (2001) 513.

    [77] C. Metcalfe, J.A. Thomas, Chem. Soc. Rev. 32 (2003) 215.

    [78] H. Chao, L.N. Ji, Bioinorg. Chem. Appl. 3 (2005) 15 and references cited therein.

    27