10.8
CiteScore
 
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
10.8
CiteScore
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original article
9 (
1_suppl
); S731-S742
doi:
10.1016/j.arabjc.2011.07.016

Synthesis, spectroscopic characterisation, biological and DNA cleavage properties of complexes of nicotinamide

Department of Chemistry, Anna University of Technology Tiruchirappalli, Tiruchirappalli 620 024, Tamil Nadu, India
Department of Chemistry, St. Joseph’s College, Tiruchirappalli, Tamil Nadu, India

⁎Corresponding author. Tel.: +91 9865605363. cs_dilip@yahoo.co.in (C. Surendra Dilip),

Disclaimer:
This article was originally published by Elsevier and was migrated to Scientific Scholar after the change of Publisher.

Peer review under responsibility of King Saud University.

Abstract

Transition metal complexes of nicotinamide with metal precursors such as Cr(III), Mn(II), Fe(III), Co(II), Ni(II), Cu(II) and Cd(II), were synthesized and characterised by physico-chemical and spectroscopic techniques. Based on analytical, spectral and magnetic moments, all the complexes are identified as distorted octahedral in structure. All the complexes are of the ML14L22 type. The shifts of the ν (C⚌N) (azomethine) and ν (C⚌O) (amide) stretches have been monitored in order to find out the donor sites of the ligands. Antibacterial and antifungal activities of the complexes were studied and the complexes were screened against bacteria and fungi. The activity data show that the metal complexes are more potent than the parent nicotinamide.

Keywords

Nicotinamide
Transition metal complexes
Antimicrobial studies
DNA binding
1

1 Introduction

Heterocyclic compounds play a significant role in many biological systems, especially N-donor ligand systems being a component of several vitamins and drugs such as nicotinamide. Nicotinamide is known as a component of the vitamin B complex as well as a component of the coenzyme, nicotinamide adenine dinucleotide (NAD). These are more important for transfer of hydrogen in the cell breath. The presence of pyridine ring in numerous naturally abundant compounds, adducts of nicotinamide are also scientific interest. Therefore, the structure of nicotinamide has been the subject of many studies. Pyridine derivatives are associated with some important biological activities such as antitubercular, anthelmintic, fungicidal, antitumor and antibacterial activity (Cunha et al., 2005; Collen et al., 1997; Wolff, 1970; Gilman et al., 1980). Nicotinamide is found to be pharmacologically and physiologically active (Rosu et al., 2006; Filho et al., 1998; Turhan-Zitouni et al., 2001a,b; Deepa et al., 2005; Radhakrishnan et al., 1976). In this study, the preparation and structural elucidation of the transition metal complexes of nicotinamide were undertaken by using spectroscopic methods (UV, IR and EPR) and their biological activities such as antibacterial, antifungal and DNA cleavage which were also studied. Comparisons of the IR spectrum of nicotinamide with those of the metal complexes were useful in determining the atoms of the ligand that are coordinated with the metal ion. In addition, magnetic and electronic spectral measurements have been done in the case of coloured complexes. Electrochemical properties of the Cr(III)/Cr(II), Fe(III)/Fe(II) and Cu(II)/Cu(III) couple were studied extensively for understanding the factors that affect the donor environments.

2

2 Experimental

2.1

2.1 Materials and methods

All the chemicals used for the preparation of the ligands were of BDH quality, AR grade. Molar conductance of the complexes was measured using a Systronic conductivity bridge at room temperature in DMSO. Conductivity measurements (Ω−1 cm2 mol−1) were carried out in DMF using a Tacussel conductivity bridge model. Magnetic susceptibilities were determined using a Guoy balance at room temperature (25 °C) with Hg[Co(SCN)4] as standard. Perkin-Elmer PE 938 spectrophotometers were used to record the IR spectra using KBr pellets. UV–visible spectra were recorded in acetone and chloroform on a Perkin-Elmer model 550-S spectrometer. EPR spectra were recorded using Broker Ac 200 spectrophotometer. The electrochemical experiments were carried out using a Trace-Iab50 from radiometer which includes a polarographic analyser (Pol 150), a polarographic stand (MDE 150) and trace Master 5 software performed using a conventional three electrodes system. Potentials are expressed versus the Ag/AgCl (KCl mol l−1) electrode separated from the test solution by a salt bridge containing the solvent/supporting electrolyte. A pre polished glassy carbon (GC) disc of 3 mm diameter (radiometer) was used as the working electrode and a platinum wire was the auxiliary electrode.

Solutions of CT DNA (calf-thymus DNA) in 50 mM NaCl/50 mM tris–HCl (pH = 7.2) gave a ratio of UV absorbance at 260 and 280 nm, A260/A280 of ∼1.8–19, indicating that DNA was sufficiently free of protein contamination. DNA concentration was determined by UV absorbance at 260 nm after 1:100 dilutions. The molar absorption coefficient was taken as 6600 M−1 cm−1. Stock solutions were kept at 4 °C and used within 4 days. Double distilled water was used to prepare the buffer.

2.2

2.2 Preparation of the complexes

Distilled ethanol (5.0 ml) and metal sulphates (0.50 m mol) are added to a 50 ml erlenmeyer flask fitted with a micro watch glass cover containing a magnetic stirring bar. Then the metal salt was dissolved in 1.0 mol of sodium sulphate and 2.0 m mol of nicotinamide. A large excess of ethanol is used, as it helps in the reaction completion. Heating and stirring the mixture was done for ∼1 h. As the nicotinamide combines with the metal forms complexes, while cooling, the mixture crystals begin to form as a crust at the surface of the reaction mixture (Swamy and Pola, 2008; Swamy et al., 2003; Goeta et al., 2000).

2.3

2.3 General properties

Cr(III), Mn(II), Fe(III), Co(II), Ni(II) and Cu(II) complexes are coloured but the cadmium complexes are colourless (Table 1). All the complexes are soluble in dimethyl formamide (DMF). Some of them are soluble only in nitrobenzene and acetone.

Table 1 Physical, analytical data of the complexes.
Complex Empirical formula Colour Mol. M.P. Yield
Wt °C %
[Cr2(L2)3(L1)9] Cr2C54H54N18O21S3 Pale brown 1468 270 70
[Mn(L2)(L1)5] MnC30H30N10O9S Mercedes red 758 295 70
[Fe2(L2)3(L1)9] Fe2C54H54N18O21S3 Reddish brown 1480 280 65
[Co (L2) (L1)5] CoC30H30N10O9S pink 760 275 68
[Ni(L2)(L1)5] NiC30H30N10O9S Green 762 272 60
[Cu(L2)(L1)5] CuC30H30N10O9S Pale blue 764 265 65
[Cd(L2)(L1)5] CdC30H30N10O9S Colourless 846 274 75

3

3 Results and discussion

The elemental analysis (Table 2) indicates that, all the metal complexes have 1:6 stoichiometry and are pale coloured amorphous substances, soluble in DMF. The molar conductance values obtained for these complexes at the concentration of 10−3 M are in the range of 10–20 Ω−1 mol−1 cm2. These values are too low to account for any dissociation of the complexes in DMF. Hence these complexes can be regarded as non-electrolytes. The ligand (L) is soluble in common organic solvents such as THF, C2H5OH, CH2Cl2 and DMSO. The octahedral metal complexes are highly soluble in DMSO and DMF and slightly soluble in CH2Cl2 and CHCl3.

Table 2 Elemental analysis, magnetic moment, molar conductance of the complex.
Complex % M % C % H % N % O μeff. BM Molar cond. (Ω−1 cm2 mol−1)
[Cr2(L2)3(L1)9] 7.05 67.21 4.15 14.01 7.60 3.17 18.25
(7.45) (67.32) (4.55) (14.00) (7.05)
7.05 66.50 4.41 13.10 7.44 5.92 17.15
[Mn(L2)(L1)5] (7.00) (66.51) (4.77) (13.08) (7.41)
7.56 64.84 3.42 13.17 7.54 5.61 17.38
[Fe2(L2)3(L1)9] (7.51) (64.86) (3.79) (13.15) (7.53)
7.78 64.85 3.72 13.15 7.51 4.90 19.89
[Co (L2) (L1)5] (7.74) (64.88) (3.79) (13.16) (7.53)
7.48 67.52 5.92 15.73 7.25 3.8 15.13
[Ni(L2)(L1)5] (7.43) (67.54) (5.89) (15.75) (7.21)
8.02 53.41 3.56 16.35 6.49 2.11 18.26
[Cu(L2)(L1)5] (7.96) (61.56) (3.52) (16.32) (7.47)
7.54 61.90 3.53 16.41 7.52 19.00
[Cd(L2)(L1)5] (7.47) (61.89) (3.54) (16.40) (7.51)

3.1

3.1 Conductance

The observed molar conductance values of the complex in DMF are in the range of 10–20 Ω−1 cm2 mol−1 at room temperature. The value for 1:1 electrolyte in DMF is of the order of 65–90 Ω−1 cm−1 mol−1. Hence from the conductivity measurement, it is concluded that the sulphate ions are covalently bonded to metal ions, which indicates that they act as ligands and not as simple ions. Based on the metal–ligand ratio calculated by the analytical data and the nature of the electrolytes given by the conductance measurements, compositions were assigned for the prepared complexes. From the magnetic and conductometric analysis it is predicted that the complexes may have the following structures [Cr2(L2)3(L1)9], [Mn(L2)(L1)5], [Fe2(L2)3(L1)9], [Co2(L2)3(L1)9], [Ni (L2)(L1)5], [Cu(L2)(L1)5] and [Cd(L2)(L1)5].

3.2

3.2 Magnetic moments

The Cr(III) complex showed a magnetic moment of 3.17 BM which is slightly lower than the spin only value of 3.87 BM expected for the three unpaired electrons, which offer possibility of an octahedral geometry. In this high spin complexes, the magnetic moments of the Fe(III) are close to the spin only value of 5.61 BM, because the ground state (derived, from the 5S state of the free iron) has no orbital angular momentum and there is no effective mechanism for introducing any by coupling with excited states. In the low spin complexes, with t 2 g 5 , e g 0 configurations may have considerable orbital contributions to their moments at room temperature (Chandra et al., 2007). The experimental magnetic moments of the prepared Fe(III) complex indicates the high spin (S = 5) octahedral d5-system. The magnetic moment values for Cu(II), Co(II) and Ni(II) complexes are shown in Table 2. The magnetic moment of Co(II) complexes are in the range of 4.90 BM indicating that the Co(II) complexes are typically high spin complexes and having octahedral structure. The Ni(II) complexes exhibit the magnetic moment values in the range 3.8 BM, indicating octahedral co-ordination of the ligands around Ni(II) ion. The Cu(II) complexes exhibit magnetic moment in the range of 2.11 BM suggesting distorted octahedral nature for these complexes. It is obvious that the metal complexes possess antiferromagnetic properties by slight intramolecular antiferromagnetic spin exchange interaction for binuclear chromium and iron, complexes with nicotinamide ligands. The absorption bands observed for the electronic spectra of the metal complexes also support the octahedral geometry (Turhan-Zitouni et al., 2001a,b).

3.3

3.3 Infrared spectral analysis

The IR spectra of ligand (L) with its octahedral complexes have been studied in order to characterise their structures. The IR spectra of the free ligand and its metal complexes were carried out in the 4000–400 cm−1 range. The IR spectra of all metal complexes were interpreted by comparing the spectra with those of the free ligand, and the results are listed in Table 3. The comparison of the band positions of various vibrations are ascertained with good evidence. In the infrared spectrum of the ligand, the band at 1385–1400 cm−1 is due to asymmetric C⚌N stretching vibration. The presence of δ(CONH2) asymmetric stretching vibration is confirmed by the band at 1740 cm−1 and the symmetrical stretching vibration is observed at 3350 cm−1. In the infrared spectrum of metal sulphate complexes, the ν CN stretching vibration were observed at 1560–1610 cm−1 and was due to coordination of the nitrogen from CN to the metal, stretching vibration for L reduced at the complex. The free nicotinamide ligand showed a strong peak at 1460 cm–1 for L, which is characteristic of the imine ν (CN) group. The ν (C⚌C) stretching vibrations are affected upon complexation and are situated at a frequency significantly different than the free ligands. Coordination of the nicotinamide ligands to the metal centre through the nitrogen atom is expected to reduce the electron density in the methine and imine link and hence lower the ν (C⚌C) and ν (CN) absorption frequencies. The peak due to ν (C⚌C) is slightly shifted to lower frequencies and appears between 1576 and 1579 cm−1, indicating the coordination of the imine nitrogen to the metal.

Table 3 Characteristic IR bands (cm−1) of the ligand and its complexes.
Ligand/complex νN–H
amideIII
νC⚌O
amideI
νC⚌N
imine
νC–N νM–N νM–SO4
Nicotinamide 1532 1680 1612 1255
[Cr2(L2)3(L1)9] 1537 1680 1560 1258 527 340
[Mn(L2)(L1)5] 1534 1682 1594 1262 523 342
[Fe2(L2)3(L1)9] 1535 1682 1583 1267 529 346
[Co2(L2)3(L1)9] 1533 1680 1610 1260 540 350
[Ni(L2)(L1)5] 1535 1684 1570 1272 535 342
[Cu(L2)(L1)5] 1532 1682 1597 1278 542 350
[Cd(L2)(L1)5] 1536 1685 1567 1285 555 355

Similarly the other metal complexes also show that the stretching frequency for the CN and C⚌C are shifted to lower indicating that the metal is coordinated through the imine nitrogen atom. The peak at 1680 cm−1corresponds to the asymmetric C⚌O stretching vibration and the symmetric O⚌CN stretching vibration observed at 1750 cm−1 confirms that all the complexes, the ligand do not coordinate with CONH2 nitrogen.

The band at 1680 cm−1 which is assignable to nicotinamide I band arising mainly from the CN stretching vibration in free ligand is found to have no change in frequencies (1680–1685 cm−1) in the metal complexes. The nicotinamide III band at 1532 cm−1 due to a coupled C⚌O stretching mode move to higher wave numbers (1537–1532 cm−1) compared to that of free ligands indicating the coordination of nicotinamide through the pyridine nitrogen. This results in the decreased CN bond order with concomitant increase in the C–O bond order. The bands due to υ (C–N–C) and υ (C–O–C) remain almost unchanged in all the complexes indicating that the nitrogen and oxygen of carbonyl moiety are not involved in binding. The free ligand showed a medium intensity band at 3210 cm−1 assigned to ν NH vibrations, which has been observed in the 3203–3207 cm−1 region for the complexes. It can be observed that there is no considerable shift in the ν NH vibrations in the case of the complexes compared to the ligands indicating non-involvement of amide NH function in the coordination. A strong intensity band observed at 1675 cm−1 and a medium intensity band at 1630 cm−1 indicate that the ligands are assigned to ν (C⚌O) and ν (CN) functions respectively. In the case of complexes the band due to ν C⚌O was observed in the 1670–1676 cm−1 region, indicating its non-involvement in the complexation. The low frequency skeletal vibration due to M–N stretching provides direct evidence of the complexation. In the present investigation, bands are observed in the 479–417 cm−1 region for ν M–N vibrations, respectively (Chandra and Gupta, 2001; Rai et al., 2005). The occurrence of bands at ∼1100 and ∼600 cm−1 suggests the presence of free sulphate ion (Td) due to ν3 and ν4 mode of vibrations, respectively. However in complexes, an additional series of six bands appeared at ∼1110, ∼1040, ∼970, ∼645, ∼600 and ∼437 cm−1 indicating the coordination of sulphate group in monodentate manner through oxygen atom, the symmetry being lowered to C3V upon coordination.

3.4

3.4 Electronic spectral analysis

Electronic spectra of ligand L and their metal complexes in DMF solutions have been recorded in the 200–1100 nm range. The UV–Vis spectra of the ligand and metal complexes in DMF showed two to seven numbers of absorption bands between 268 and 734 nm (Table 4). The bands below 455 nm are mostly associated with intra ligand π → π and n → π transitions. In the electronic spectra of the ligand and their metal complexes, the presence of a wide range of bands is due to both π → π, n → π and dd transitions and also due to charge transfer transition arising from π electron interactions between the metal and ligand that involves either a metal-to-ligand or ligand-to-metal electron transfer. The absorption bands observed within the range of 331–397 nm in DMF are mostly due to the transition of n → π of imine group corresponding to the ligand or metal complexes.

Table 4 Electronic spectral data and ligand field parameters of complexes.
Complex ν1 ν2 ν3 Dq B β β% ν2/ν1 ν3/ν2 λ’ LFSE kcal mol−1
[Cr2(L2)3(L1)9] 12,345 18,654 24,378 1186 949 0.80 20.00 1.5 1.3 37.43
[Mn(L2)(L1)5] 11,794 21,978 25,974 1124 980 0.871 13.90 1.8 1.2 34.86
[Fe2(L2)3(L1)9] 13,122 17,932 1086 865 0.796 21.10 1.4 35.56
[Co2(L2)3(L1)9] 8892 15,439 20,085 1002 749 0.778 23.03 1.7 2.1 27.17
[Ni2(L2)(L1)5] 13,810 15,151 25,316 1040 852 0.820 18.00 1.4 1.8 30.82
[Cu(L2)(L1)5] 10,526 16,181 25,000 1020 875 0.857 14.30 1.5 1.6 33.36

3.4.1

3.4.1 Manganese(II) complexes

The Mn(II) complex shows magnetic moment of 5.92 BM at room temperature corresponding to the five unpaired electrons. The electronic spectra of the Mn(II) complexes exhibit four weak intensity absorption bands in the range 17,794 (ε = 27 l mol−1 cm−1), 21,978 (ε = 35 l mol−1 cm−1), 25,974 (ε = 62 l mol−1 cm−1) and 35,425 cm−1 (ε = 129 l mol−1 cm−1), assigned to the transitions: 6A1g → 4T1g(4G), 6A1g → 4Eg, 4A1g(4G) (10B + 5C), 6A1g → 4Eg(4D) (17B + 5C) and 6A1g → 4T1g(4P) (7B + 7C), respectively.

Electronic spectra of Mn(II) complexes, a strong band appearing at ∼29400 cm−1 may be attributed to the ππ transition of the C⚌N group. The complexes in the visible region show very weak absorptions at ∼21,700 and ∼25,000 cm−1, which may be assigned to 6A1g → 4E(G) and 6A1 → 4A1(G) transitions of a octahedral (Hathaway and Billing, 1970). The β values of the complex lie at 0.871. β Value indicates the appreciable covalent character of the metal–ligand ‘σ’ bond.

In Mn(II), the values B and C were calculated from the second and third transitions, because these transitions are free from the crystal field splitting. Slater Condon-shortly parameter F2 and F4 are related to the Racha parameter B and C are calculated as: B = F2-5F4 and C = 35F4. On the basis of the above spectral studies distorted octahedral structures are suggested for the complexes.

3.4.2

3.4.2 Ferric complexes

The electronic spectra of the Fe(III) complex showed the strong bands at 17,932 and 13,122 cm−1 which are not possible to identify the type of the dd transition, due to a strong charge-transfer (CT) band tailing from the UV-region to the visible region. Magnetic moment of 5.61 BM was observed for the Fe(III) complex that is slightly lower than the magnetic moment of a high spin octahedral complex. Thus the structure of Fe(III) complex is tentatively interpreted to possess an octahedral geometry.

3.4.3

3.4.3 Cobalt(II) complexes

The electronic spectrum of the cobalt complexes exhibited three bands at 15,674, 18,832 and 25,971 cm−1 that are assigned to the transitions 4T1g(F) → 4T2g(F), 4T1g(F) → 4A2g(F) and 4T1g(F) → 4T1g(P), respectively, suggesting an octahedral geometry around the Co(II) ion. The magnetic moments of Co(II) complexes were found to be 4.90 BM also indicating octahedral geometry. Interelectronic repulsion parameter (B′) of Co(II) complex shows 749 cm−1 which is less than the free ion (B) value of 971 cm−1 suggesting the considerable orbital overlap and delocalisation of electrons on the metal ion. The nephelauxetic ratio (β) for the Co(II) complex (0.778) is less than one, suggesting partial covalency in the metal ligand bond. The values Dq, β%, LFSE and ν2/ν1 (Table 4) suggest the distorted octahedral geometry for Co(II) complex (Silverstein et al., 2007; Singh and Pheko, 2008).

3.4.4

3.4.4 Nickel(II) complexes

The absorption spectra of Ni(II) complexes display three dd transition bands at 12,108, 19,416 and 27,322 cm−1 which correspond to 3A2g → 3T2g, 3A2g → 3T1g(F) and 3A2g → 3T1g(P). The magnetic moments of Ni(II) complexes were found to be 3.8 BM supporting the d8 high spin distorted octahedral structure. The ligand field parameter such as Dq, B′, β, β% and LFSE have been calculated by using Band-fitting equation given by Underhilland Billing (Table 4). The ratio ν2/ν1 was found to be 1.4046 which is well which the range (1.40–1.61) and is indicative of octahedral stereochemistry for this Ni(II) complex (Kivelson and Neiman, 1961). The Racha parameter B′ is 852 cm−1, which is less than the free ion (B) value of 1040 cm−1 indicating the covalent character. The ratio ν2/ν1 and β% also support further the octahedral geometry around the Ni(II) ion (Wehrli et al., 1988). These observations reveal that the nickel complexes possess an octahedral geometry and D3 symmetry.

3.4.5

3.4.5 Copper(II) complexes

The magnetic moment values of Cu(II) complexes are 2.11 BM that fall within the normal range observed for distorted octahedral complexes. From the results, Cu(II) complexes show, a single broad band in the range 18,920 cm−1 due to transition between 2Eg → 2T2g suggesting tetragonal geometry. Tetragonal or square planar Cu(II) complexes are expected to give three bands. However, these three bands usually overlapped in tetragonal complexes, to give one broad absorption band. The electronic spectra and magnetic moment data for all Cu(II) complexes coupled with the analytical, conductance data suggest the tetragonal geometry for all the complexes.

3.4.6

3.4.6 Ligand field parameters

Various ligand field parameters were calculated and their values are listed in Table 4. The results show that the ligand field parameter values of the complexes were consistent with octahedral geometry. The ligand field strength, 10 Dq, can be estimated roughly for each complex by the relationship of nonlinear for all the anisotropic ligands. The β values indicate the covalent character, which is due to the presence of σ and π bonds between the metal and ligands. Δ Values indicate the energy difference between the principle bands, which are formed due to ligand field absorption. This type of complex may have either C4v or D4h symmetry, which arises from the lifting of the degeneracy of the orbital triplet (in octahedral symmetry) in the order of increasing energy or assuming D4h symmetry. C4v symmetry is ruled out because of the higher splitting of the first band. This suggests that it possesses distorted octahedral geometry around the metal ion.

3.5

3.5 EPR spectral analysis

3.5.1

3.5.1 EPR spectrum of Cu(II) complex

The EPR spectra of the Cu(II) complexes were recorded (Fig. 1) in a polycrystalline sample as well as in solution at room temperature in different frequencies. The g-values are calculated by using the expression, G = 2.0023 ( 1 - 4 λ / 10 Dq ) , where λ is the spin–orbit coupling constant for the metal ion.

EPR spectrum of Cu(II) complex.
Figure 1
EPR spectrum of Cu(II) complex.

The analysis of spectra gives the values for g|| 2.22–2.27 and g = 2.09 - 2.10 . The observed ‘g’ values for the complexes are less than 2.3 in agreement with the covalent character of the metal ligand bond. The trend g | | > g > 2.0023 observed for the complexes indicates that the unpaired electron is localised in dx2–y2 orbital of the Cu(II) ion and spectral features are characteristic of axial symmetry. Tetragonal elongated structure is confirmed for the aforesaid complex.

3.5.2

3.5.2 Chromium complex

The spin–orbit coupling constant from the free ion value 90 cm−1 of Cr(III) complex is reduced from its metal ion and this can be employed as a measure of metal–ligand covalency (Fig. 2). The values indicate that the complexes under study have a substantial covalent character. The g-values were calculated and found in the range of 1.97–1.99, corresponding to 6-coordinated geometry. It is possible to define a covalency parameter analogous to the nephelauxetic parameter, which is the ratio of the spin–orbit coupling constant for the complex and the free Cr(III) ions.

EPR spectrum of Cr(III) complex.
Figure 2
EPR spectrum of Cr(III) complex.

3.6

3.6 Thermal decomposition

The TG curves of the metal–nicotinamide complexes suggest that modes of decomposition resemble each other closely. All these complexes follow the same pattern of thermal decomposition. The complexes remain almost unaffected up to ∼70 °C (Table 5). After this, a slight depression is observed up to ∼110 °C. The weight loss in this temperature range is equivalent to two water molecules in the complexes indicating them to be lattice water in conformity with our earlier observations from analytical and IR spectral investigations. The anhydrous complexes remain stable up to ∼300 °C and thereafter the complexes show rapid degradation presumably due to decomposition of organic constitution of the complexes as indicated by the steep fall in the percentage weight loss. The decomposition continues up to ∼800 °C and reaches to a stable product in each complex as indicated by the constituency in weight in the plateau of the thermogram. The decomposition temperature varies for different complexes as shown in Table 5.

Table 5 Thermal analysis data of the complexes.
Complexes Total wt. for TG (mg) Temp. range of water loss Water loss (%) Temp. range of melting (°C) Temp. range of decomposition (°C)
[Cr2(L2)3(L1)9] 110 80–110 5.08 170–235 420–800
[Mn(L2)(L1)5] 90 85–110 4.92 180–240 400–780
[Fe2(L2)3(L1)9] 105 75–105 4.68 160–250 410–800
[Co (L2) (L1)5] 90 78–105 4.46 150–220 385–750
[Ni(L2)(L1)5] 110 80–105 6.23 155–230 370–750
[Cu(L2)(L1)5] 105 75–110 6.02 160–245 350–760
[Cd(L2)(L1)5] 95 72–105 5.43 190–260 430–750

There is no mass loss observed up to 185 °C for the Cr complex. The sharp endothermic DTA peaks at 123–183 °C show that the complex melts before decomposition. The thermo gravimetric curves of all the complexes show single stage decomposition wherein simultaneous decomposition and oxidation of the organic moiety and the volatilisation of chromium take place after 165–185 °C. There are broad endothermic peaks at 274 °C indicating slow decomposition of the complex. Due to the oxidation of the organic moiety two broad exothermic peaks appear between 290 °C and 355 °C for the complex. No such exothermic peaks were observed in the case of manganese complex. Due to volatilisation the weight of the final residue does not conform to any particular and product.

For Fe complex, there is no mass loss up to 181 °C, two endothermic peaks appeared at 80 and 120 °C, which are due to the decomposition and loss of coordinated water molecules. There afterwards, the complex shows single stage decomposition without giving any stable intermediate compound. A broad exothermic peak at 220 °C followed by an endothermic peak at 280 °C shows the oxidative process. The weight of the final residue almost corresponds to the formation of FeO.

3.7

3.7 Electro chemical studies

Cyclic voltammogram of cobalt complex recorded in 1.0 mM electrolyte containing DMSO is shown in Fig. 3a. The samples were scanned in the range from −1.9 to +1.9 V at the scan rate of 20 mV S−1. During the cathodic scanning process, Cobalt complexes show a reduction peaks at 0.7 V due to the deposition of Co(II) on electrode surface and an oxidation peak at 1.05 V due to the stripping of cobalt from the surface. Also a small reduction peak was observed about at −0.85 V, which is mainly attributed to the adsorption of the double hydroxide at the cathode. Under these conditions, the pH value near the cathode seems to rise due to the proton consumption. Therefore, the hydroxide of metal ion may be produced in the vicinity of the electrode.

Cyclic voltammogram of (a) 1 mM Co(II) complex (b) Cd(II) complex in 0.1 M TBATFB in DMSO on Pt electrode, V = 0.1 Vs−1, (vs. Ag|Ag+ electrode).
Figure 3
Cyclic voltammogram of (a) 1 mM Co(II) complex (b) Cd(II) complex in 0.1 M TBATFB in DMSO on Pt electrode, V = 0.1 Vs−1, (vs. Ag|Ag+ electrode).

The cyclic voltammogram of Cd(II) complex recorded from 0.1 M TBATFB in DMSO showed in Fig. 3b. Potential varied from −2.0 to 1.0 V at the scan rate of 50 mV s−1. During the cathodic scan, no reducible species present from 1.0 to −0.85 V, cathodic peak observed at −1.2 V due to the dissolution of cadmium from the electrolyte solution. During the reverse process, three oxidation peaks were observed (−0.7 and −0.2 V and a peak at 0.45 V) which is corresponding to the oxidation of cadmium in three phases.

Cyclic voltammogram of chromium complex recorded in electrolyte containing water is shown in Fig. 4a. The samples were scanned in the range from −1.0 to +1.0 V at the scan rate of 20 mV S−1. During the cathodic scanning process, chromium nicotinamide complex shows a reduction peaks at −0.25 V and an oxidation peaks at 0.8 V due to the stripping of chromium from the surface. Under these conditions, the pH value near the cathode seems to rise due to the proton consumption. Therefore, the hydroxide of metal ion may be produced in the vicinity of the electrode. Cyclic voltammogram of Ferric complex is recorded in 1 mM electrolyte containing water is shown in Fig. 4b. The samples were scanned in the range from −1.0 to +1.0 V at the scan rate of 20 mV S−1. During the scanning process, a reduction peak was observed at 0.25 V due to the dissolution of hydroxide ion and the deposition of ferrous ion on the base metal. From the figure, we can conclude that the good electrochemical performance of the electrode can be attributed to its higher electrical conductivity and facile ionic transportation that are offered by the nicotinamide additive. This metal complex also undergoes transformation between +2 and +3, and the transformation between these two oxidation states is highly reversible.

Cyclic voltammogram of (a) 1 mM Cr(III) complex (b) Fe(III) complex in 1 mM Fe(III) in DMSO on Pt electrode, V = 0.1 Vs−1, (vs. Ag|Ag+ electrode).
Figure 4
Cyclic voltammogram of (a) 1 mM Cr(III) complex (b) Fe(III) complex in 1 mM Fe(III) in DMSO on Pt electrode, V = 0.1 Vs−1, (vs. Ag|Ag+ electrode).

3.8

3.8 Powder X-ray analysis

The XRD (Powder Pattern) of the complexes [Co2(L2)3(L1)9] and [Ni2(L2)(L1)5] were indexed in X-ray diffractometer and the unit cell parameters have been calculated with the help of a computer from 2θ values (Fig. 5). The direct constant parameters like A, B, C, α, β, γ and V (volume) are given in Table 6. The density of the complexes has been determined by the floatation technique in a saturated solution of NaCl, KBr and benzene separately.

The XRD (powder pattern) of the complexes (a) [Co2(L2)3(L1)9] and (b) [Ni2(L2)(L1)5].
Figure 5
The XRD (powder pattern) of the complexes (a) [Co2(L2)3(L1)9] and (b) [Ni2(L2)(L1)5].
Table 6 X-ray powder pattern reports.
Compound 2θ Values Unit cell parameters Density (gcc) n Possible geometry
[Co (L2)3(L1)9] 13.284 13.936 16.141
17.228 18.130 18.264 A = 13.300 Å
19.300 21.756 22.659
25.010 25.516 30.914 B = 20.543 Å
32.051 39.253 41.659 C = 6.985 Å
42.244 43.849 46.389
47.308 48.194 49.881 α = 90.000°
50.032 51.419 53.174 β = 102.325° 0.83 1 Monoclinic
53.541 53.976 54.076
54.277 55.680 56.148 γ = 90.000°
56.265 56.466 56.766
56.833 56.9S4 57.184 V = 1864.37 Å3
57.836 58.0S7 58.621
[Ni(L2)(L1)5] 13.066 13.417 14.804 A = 8.546 Å
21.856 22.241 23.761 B = 18.487 Å
26.084 27.956 32.668 C = 7.371 Å
32.835 35.944 39.453 α=90.000° 1.25 1 Monoclinic
41.458 44.216 46.422 β = l 06.893°
47.324 49.898 51.251 γ = 90.000°
54.527 56.081 56.365 V = 1114.34 Å

The number of formula units per unit cell (n) was calculated from the relation n = dNV / M , where d = density of the compound, N = Avogadro’s number, V = volume of the unit cell and M = molecular weight of the complex. The value of ‘n’ is found to be 10 for both complexes which agree well with the suggested monoclinic structure of the complexes. In addition, we have carried out powder X-ray diffraction studies of complex. Powder XRD pattern of [Co(L2)3(L1)9] consists of 7 reflections in the range 5–50° (2θ) the inter planar spacing (d) has been calculated from the positions of intense peaks using Bragg’s relationship. The 2θ values with maximum intensity of the peak for ligand were found to be 5.709 (2θ) which correspond to d = 15.4539. All the important peaks have been indexed and the observed values of inter planar distance were compared with the calculated ones. It was found that there is good agreement between the calculated and observed values. The (h2 + k2 + l2) values are 1, 5, 27, 32, 58, 72 and 74. The presence of forbidden number confirms the tetragonal systems. This implies that the cobalt and nickel complexes are distorted octahedral in structure.

3.9

3.9 Antimicrobial activity

The biological and medicinal potency of the coordination complexes has been established by their antitumor, antiviral and antimalarial activities. This characteristic property has been related to the ability of the metal ion to form complexes with ligand containing nitrogen donor atoms. The ligands and their complexes were screened for their antibacterial activity (Plakatouras et al., 1992) against Escherichia coli and Staphylococcus aurious and antifungal activity against (Assour, 1965) Aspergillus niger and Aspergillus flavus at the concentration of 100 μg/0.1 cm3. The standard drugs streptomycin and clotrimazole were also tested for their antibacterial and antifungal activity at the same concentration under the conditions similar to that of the test compounds concentration.

3.10

3.10 Antibacterial activity

For in vitro antimicrobial activity, the investigated compounds were tested against the bacteria such as Shigella dysenteriae, E. coli and Bacillus subtilis. The minimum inhibitory concentration (MIC) values of the compounds against the growth of microorganisms are summarised in Table 7. It is observed that the copper and cadmium complexes are more active in Pseudomonas aeruginosa and B. subtilis, respectively compared to other bacterial organisms. Nickel and cobalt complexes are moderately active in all bacterial organisms compared with standard streptomycin. From the results it is found that the copper complex is more active in Streptococcus-β-haemolyticus than the other complexes.

Table 7 Antibacterial activities of the complexes and standard.
Diameter of zone of inhibition (in mm)
C1 C2 C3 C4 C5 C6 C7 Ligand
μg disc−1 30 200 30 200 30 200 30 200 30 200 30 200 30 200 30 200
Gram positive bacteria
Bacillus subtilis 10 17 10 15 08 11 13 20 11 17 14 19 15 21 07 09
Streptococcus-β-haemolyticus 11 16 13 16 11 15 12 17 12 17 16 21 12 16 08 11
Gram negative bacteria
Shigella dysenteriae 12 18 12 17 12 12 15 10 13 15 12 14 11 18 06 08
Pseudomonas aeruginonosa 14 22 14 15 14 14 16 11 14 16 11 17 13 16 09 11
Escherichia coli 13 21 13 16 13 16 14 12 11 09 10 16 12 16 08 10

3.11

3.11 Antifungal activity studies

The results of the antifungal screening of the nicotinamide and the metal complexes with Candida albicans, A. niger and Aspergillus fumigates at concentration of 200 μg by disc method are given in the Fig. 6. Comparative studies of the ligands and their complexes indicated that metal complexes exhibit higher antifungal activity than the free ligands (Table 8). The antifungal activity results revealed that the ligands and their Cu(II), Co(II) and Ni(II), complexes have exhibited weak to good activity against A. niger and A. flavus. The ligand and its Cu(II) and Co(II) complexes show weak activity when compared to the standard drug clotrimazole. The order of the metal complexes follow Cu(II) > Cd(II) > Ni(II) > Co(III) > Mn(II) > Fe(III) > Cr(III).

Anti fungal activity of Aspergillus flavus of (A) ligand (B) [Ni(L2)(L1)5].
Figure 6
Anti fungal activity of Aspergillus flavus of (A) ligand (B) [Ni(L2)(L1)5].
Table 8 Antifungal activities of the complexes and standard.
Diameter of zone of inhibition (in mm)
C1 C2 C3 C4 C5 C6 C7 Ligand
μg disc−1 200 200 200 200 200 200 200 200
Candida albicans 15 10 17 15 19 22 17 11
Aspergillus niger 17 10 18 15 18 23 19 10
Aspergillus fumigates 18 11 15 13 20 21 18 13

The higher activity of metal complexes can be explained on the basis of overtons concept and chelation theory. According to overtons concept of cell permeability, the lipid membranes that surround the cell favour the passage of only the lipid soluble material due to which lip solubility is an important factor, which controls antimicrobial activity. On chelation, the polarity of metal ion will be reduced to a greater extent due to the overlap of the ligand orbital and partial sharing of the positive charge of the metal ion with donor group. Further it increases the delocalisation of pi-electron over the whole chelate ring, lyophilizes, enhances the penetration of the complexes into lipid membranes blacking the metal binding sites in the enzymes of microorganisms. These complexes also disturb the respiration process of the cell and thus block the synthesis of protein that restricts further growth of the organism.

3.12

3.12 DNA cleavage studies

The metal complexes were able to convert super coiled DNA into open circular DNA. The general oxidative mechanisms proposed account for DNA cleavage by hydroxyl radicals. The general oxidative mechanisms proposed account for the abstraction of a hydrogen atom from sugar units predicting the release of specific residues arising from transformed sugars, depending on the position from which the hydrogen atom is removed. The cleavage is inhibited by the free radical scavengers implying that hydroxyl radical or peroxy derivatives mediate the cleavage reaction. The reaction is modulated by a metallo complex bound hydroxyl radical or a peroxo species generated from the co-reactant H2O2.

In the present study, the CT–DNA gel electrophoresis experiment was conducted at 35 °C using our synthesized complexes in the presence of H2O2 as an oxidant. It was found that, at very low concentrations, few complexes exhibit nuclease activity in the presence of H2O2. Control experiment using DNA alone does not show any significant cleavage of CT-DNA even on longer exposure time. Hence, we conclude that the copper complex cleaves DNA as compared with control DNA, while other complexes do not cleave DNA in the presence of H2O2. Probably this may be due to the formation of redox couple of the metal ions and its behaviour.

The redox property of the metal complexes mediates oxidation of nucleic acids. In oxidative DNA cleavage mechanism, metal ions in the complexes react with H2O2 to generate the hydroxyl radical which attacks at the C4′ position of the sugar moiety and finally cleaves the DNA. Copper complex reacts with H2O2 to produce hydroxyl radical, hydroxyl ion and Cu(II) form. The formation of hydroxyl radical by the copper complex is further compared with other complexes with H2O2. Hence, copper complex can promote redox mediated cleavage of DNA reaction on sugar ring. The Cu(II) ion formation is supported by electrochemical study. Further, the presence of a smear in the gel diagram indicates the presence of radical cleavage.

As can be seen from the results (Fig. 7), at a very low concentration, few complexes exhibit nuclease activity in the presence of H2O2. Control experiment using DNA alone (lane 1) does not show any significant cleavage of CT-DNA even on longer exposure time. From the observed results, we conclude that the complexes, copper complex (lane 6), nickel complex (lane 7) and cobalt complex (lane 8) cleave DNA as compared to control DNA while other complexes (lanes 1–5) do not cleave DNA in the presence of H2O2. 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.

DNA cleavage studies of ligand and metal complexes.
Figure 7
DNA cleavage studies of ligand and metal complexes.

4

4 Conclusion

Based on analytical, conductance, magnetic, infrared, electronic spectral data, EPR, CV, TGA and X-ray powder pattern, all these complexes are assigned to be in octahedral geometry and exhibit coordination number six. Biological studies of these complexes reveal that these complexes show better activity compared to their respective ligands. It is further confirmed on the basis of considerable low value of magnetic moments in the case of Cu(II), Ni(II) and Co(II) complexes, which indicate antiferromagnetic coupling interaction between the two metal centres in the complexes. The value covalency factor (β) and orbital reduction factor (k) suggest the covalent nature of the complexes. The electrochemical properties of the metal complexes revealed the quasi-reversible one electron/two electron transfer redox process. The nicotinamide and some of the metal complexes were found to be active against some of the representative bacterial and fungal strains.

The screening of biological activities of ligand and its complexes against the fungi Alternaria brassicae, A. niger and Fusarium oxysporum and the pathogenic bacteria Xanthomonas compestris and P. aeruginosa indicates that the complexes show the enhanced activity in comparison to free ligand. The complexes, copper nickel and cobalt complexes cleave DNA as compared to control DNA while other complexes do not cleave DNA in the presence of H2O2.

References

  1. , . Electron spin resonance of tetraphenylporphine chelates. J. Chem. Phys.. 1965;43:2477-2489.
    [Google Scholar]
  2. , , . Synthesis and spectral studies on chromium (III), manganese (II), iron (III), cobalt (II), nickel (II) and copper (II) complexes of fourteen membered and sixteen membered macrocyclic ligand. Indian J. Chem.. 2001;40A:775-779.
    [Google Scholar]
  3. , , , . Modern spectroscopic and biological approach in the characterization of a novel 14-membered [N4] macrocyclic ligand and its transition metal complexes. Trans. Met. Chem.. 2007;32:240-245.
    [Google Scholar]
  4. , , , . Characterization of 60Co y-radiation induced radical products of antipyrine by means of high-performance liquid chromatography, massspectrometry, capillary zone electrophoresis, micellar electrokinetic capillarychromatography and nuclear magnetic resonance spectrometry. J. Chromatogr. Acta. 1997;755:95-103.
    [Google Scholar]
  5. , , , , , , , , , , . Structural studies of 4-aminoantipyrine derivatives. J. Mol. Struct. 2005:32-39. 752
    [Google Scholar]
  6. , , , . Cadmium(II) complexes of 1,2-di(imino-4′-antipyrinyl)ethane. Synth. React. Inorg. Met.-Org. Nano-Met. Chem.. 2005;35:883-888.
    [Google Scholar]
  7. , , , , , , , , . Further studies on analgesic activity of cyclic imides. Farmaco. 1998;53:55-57.
    [Google Scholar]
  8. , , , . The pharmacological basis of therapeutics. New York, USA: Macmillan Publishing Co; .
  9. , , , , , , . Copper(II) complexes of the isomeric tetraazamacrocyclic ligands 1, 11- and 1, 8-bis-(2-pyridylmethyl)-1, 4, 8, 11-tatraazacyclotetradecane and of the 1, 4, 8, 11-tetraazacyclotetradecane-5–12-dione analogue at neutral and basic pH. J. Chem. Soc. Dalton Trans. 2000:1873-1880.
    [Google Scholar]
  10. , , . The electronic properties and stereochemistry of mono-nuclear complexes of the Copper(II) ion. Coord. Chem. Rev.. 1970;5:143-207.
    [Google Scholar]
  11. , , . ESR studies on the bonding in copper complexes. J. Chem. Phys.. 1961;35:149-155.
    [Google Scholar]
  12. , , , , , , . Coordination chemistry of corrosion inhibitors of the benzotriazole type: Preparation and characterization of Cobalt(II) complexes with 1-Methylbenzotriazole (Mebta) and the crystal stucutures of [CoCl2(Mebta)4], Trans-[Co(NCS)2 (Mebta)4], Trans-[Co(NCS)2 (MeOH)2 (Mebta)2] and Cis-[Co(NO3)2 (Mebta)2] Polyhedron. 1992;11:2657-2672.
    [Google Scholar]
  13. , , , . Copper(II) complexes of salicylal-4-amino antipyrine and 2-hydroxy naphthal-4-amino antipyrine. J. Inorg. Nucl. Chem.. 1976;35:2211-2220.
    [Google Scholar]
  14. , , , . Lanthanum(III) and praseodymium(III) complexes with isatin thiosemicarbazones. Spectrochim. Acta A. 2005;61:2761-2765.
    [Google Scholar]
  15. , , , , , . Copper(II) complexes with ligands derived from 4-amino-2, 3-dimethyl-1-phenyl-3-pyrazoline-5-one. Molecules. 2006;11:904914.
    [Google Scholar]
  16. , , . Spectroscopic Characterization of the 1-Substituted 3,3’-Diphenyl-4-(2’-hydroxyphenyl)azetidine-2-ones: Application of 13C NMR, 1H-13C COSY NMR and Mass Spectroscopy. Spectrochim. Acta. Part A. 2008;70:595-600.
    [Google Scholar]
  17. , , , , , , , . Small cysteine-rich peptides resembling antimicrobial peptides have been under-predicted in plants. Plant J.. 2007;51:262-280.
    [Google Scholar]
  18. , , . Spectroscopic Studies on Co(II), Ni(II) Cu(II) and Zn(II) complexes with a N4-macrocyclic ligands. Spectrochim. Acta A. 2008;70:929-933.
    [Google Scholar]
  19. , , , , , . Non-template synthesis of ‘N4 ’ di- and tetra-amide macrocyclic ligands with variable ring sizes. Tetrahedron. 2003;59:10093-10096.
    [Google Scholar]
  20. , , , , . Synthesis of some triazolyl-antipyrine derivatives and investigation of analgesic activity. Eur. J. Med. Chem.. 2001;36:685-689.
    [Google Scholar]
  21. , , , , . Synthesis of some tiazolyl-antipyrine derivatives and investigation of analgesic activity. Eur. J. Med. Chem.. 2001;36:685-689.
    [Google Scholar]
  22. , , , . Biological Magnetic Resonance Imaging: Principles, Methodology, Applications. New York: VCH; .
  23. Wolff, M.E. 1970. Burger’s Medicinal Chemistry, third ed., vol. 1. Wiley, New York, USA.
Show Sections