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 (
3
); 335-343
doi:
10.1016/j.arabjc.2011.11.004

Synthesis, spectroscopic characterization and in vitro antimicrobial studies of Schiff base ligand, H2L derived from glyoxalic acid and 1,8-diaminonaphthalene and its Co(II), Ni(II), Cu(II) and Zn(II) complexes

Department of Chemistry, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
Division of Inorganic Chemistry, Department of Chemistry, Aligarh Muslim University, Aligarh 202 002, India
Interdisciplinary Biotechnology Unit, Aligarh Muslim University, Aligarh 202 002, India

⁎Corresponding author. shakir078@yahoo.com (Mohammad Shakir)

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

A novel Schiff base ligand, N,N′-bis (glyoxalicacidcarboxaldiimine)-1,8-diaminonaphthalene [H2L] obtained by the condensation of glyoxalic acid and 1,8-diaminonaphthalene and its mononuclear complexes of type, [ML] [M = Co(II), Ni(II), Cu(II), Zn(II)] have been synthesized and characterized on the basis of elemental analysis, molar conductance, magnetic susceptibility measurements and spectroscopic studies viz., FT-IR, EPR, 1H NMR, FAB-Mass, UV–vis and magnetic moment data. A square planar geometry has been assigned on the basis of UV–vis and magnetic susceptibility around Co(II), Ni(II) and Cu(II) ions while conductivity data showed non electrolytic nature of all the complexes. The synthesized ligand, H2L and its complexes have been tested against Streptococcus mutans, Staphylococcus pyogenes, MRSA (Gram positive bacteria), Pseudomonas aeruginosa, Salmonella typhimurium, Escherichia coli (Gram negative bacteria), Candida albicans, Candida krusei, Candida parapsilosis and Candida neroformans and results suggested that Cu(II) complex has significant antimicrobial activity.

Keywords

Schiff base
1,8-Diaminonaphthalene
Spectroscopic studies
Antimicrobial activity
1

1 Introduction

Schiff base ligands have played an integral and important role in the development of coordination chemistry since the late 19th century. Metal complexes of these ligands are ubiquitous due to their facile synthesis, wide applications and the accessibility of diverse structural modifications (Coles et al., 1998). The chemistry of Schiff base ligands and their metal complexes has expanded enormously and encompasses a vast area of organometallic compounds and various aspects of bioinorganic chemistry (Yamada, 1999). Their chelating structures, moderate electron donation and easy tunable electronic and steric effects also make Schiff bases as versatile ligands capable of stabilizing different metals in various oxidation states with unusual structural features and controlling the performance of metals in a variety of useful catalytic transformations (Cozzi, 2004; Garnoviskii et al., 1993; Makio et al., 2002; Mcmurry, 1989; Atwood and Harrey, 2001; Gansauer and Bluhm, 2000). Characteristically, Schiff base provides geometrical cavity control for host–guest interaction and modulation of its lipophilicity offers remarkable selectivity, sensitivity and stability for a specific metal ion. It has been observed that most of the metals make 1:1 metal complexes with Schiff bases (Jacobsen et al., 1999). Among the inorganic mimics of enzymes, metal complexes containing porphyrin, salen and phthalocyanine ligands have been investigated as possible alternative catalysts in many oxidation and hydroxylation reactions (Kim et al., 2007; Raja and Ratnasamy, 1996; Jacob et al., 1998; Lu et al., 2006; Fukuda and Katsuki, 1997, 1993b, 1995c, 1997d; Li et al., 1993; Yamashita and Katsuki, 1995; Tokunaga et al., 1997). It has been suggested that the azomethine linkage is responsible for the biological activities such as in the treatment of cancer (Yu et al., 1993), as antibactericide (Wang et al., 2001) as antivirus agent (Tarasoni et al., 2000; Singh et al., 1981) and other biological properties (Charo et al., 2004). Several applications have been related for these complexes in chemical analysis, absorption and transport of oxygen, in pesticides and heterogenous and homogenous catalysis for oxidation and polymerization of organic compounds (Kanthimathi et al., 2000; Zhang et al., 2006; Bahramian et al., 2006). Transition metal complexes containing oxygen and nitrogen donor Schiff base ligands have been of research interest for many years (Krishnaraj et al., 2008). The metals in these complexes act as active sites and thereby catalyse chemical reactions (Sheldon and Kochi, 1981). Antimicrobial diseases are now more frequent than during the first half of the century, being still difficult to diagnose clinically. During the later half of the century, particularly during the past two decades, a number of different classes of antibacterial (Appelbaum and Hunter, 2000) antifungal agents (Brickner et al., 1996; Andriole et al., 1998) have been discovered. At the present time, antibacterial sulfa drugs, nitrofuranes, penicillins, cephalosporins, tetracyclines, macrolides and oxaolidinones and antifungal agents such as fluconazole, ketoconazole and miconazole, including amphotericin B exhibit their antimicrobial activity (Appelbaum and Hunter, 2000; Current et al., 1995; Snaz-Nebot et al., 1995; Vazquez et al., 1998). Although there has been much progress in antibacterial and antifungal therapies many problems remain to be solved for most antimicrobial drugs available. For example, although amphotericin B has strong antifungal activity its serious nephrotoxicity often limits its clinical application (Fanos and Cataldi, 2000). Fungal infections are not usually limited to the superficial tissues. Indeed, a significant increase in life threatening systemic fungal infections has been reported (Sundriyal et al., 2006). The fundamental reason for this is the increasing number of patients at risk, including those with advanced age, major surgery, immunosuppressive therapy, acquired immunodeficiency syndrome (AIDS), cancer treatment and solid-organ and hematopoietic stem cell transplantation (Nucci and Marr, 2005). The search and development of more effective antifungal agents are mandatory (Martins et al., 2009a,b) and some Schiff bases are known to be promising antifungal agents.

However, a few reports appeared on Schiff base complexes derived from glyoxalic acid involving thiosemicarbazide (Tian et al., 1995) and α-alanine (Nakao et al., 1966).

Herein, we report the synthesis and spectroscopic characterization of a novel Schiff base ligand, H2L derived from the condensation of 1,8-diaminonaphthalene and glyoxalic acid and its Co(II), Ni(II), Cu(II) and Zn(II) complexes followed by their in vitro antimicrobial screening against different species of bacteria and fungi.

2

2 Experimental

2.1

2.1 Material and methods

All the reagents used were of A.R. grade. The metal salts M(NO3)2·6H2O [M = Co(II), Ni(II), Zn(II); n = 6, Cu(II); n = 3] and the chemical 1, 8-diaminonaphthalene and glyoxalic acid (All E. Merck) were commerically pure samples used as received.

2.2

2.2 Synthesis of Schiff base ligand, H2L

To a stirring methanolic solution (25 ml) of 1,8-diaminonaphthalene (1 mmol; 0.158 g), methanolic solution (25 ml) of glyoxalic acid (2 mmol, 0.22 ml) was added gradually. The reaction mixture was magnetically stirred for 24 h at room temperature, leading to the isolation of yellow solid product which was filtered, washed with methanol and dried in vacuum over anhydrous calcium chloride.

2.3

2.3 Synthesis of complexes, [ML] [M = Co(II), Ni(II), Cu(II), Zn(II)]

To a solution of metal nitrate (1 mmol) dissolved in 10 ml methanol, methanolic solutions of both 1,8-diaminonaphthalene (1 mmol, 0.158 g) and glyoxalic acid (2 mmol, 0.22 ml) were added simultaneously under stirring. The stirring of reaction mixture was continued for 24 h at room temperature, leading to the isolation of coloured solid product which was filtered, washed with methanol and finally dried in vacuum over anhydrous calcium chloride.

2.4

2.4 Physical measurements

The elemental analysis was made using Perkin–Elmer 2400 CHN elemental analyser. The FT-IR spectra (4000–200 cm−1) were recorded as KBr pellets on Perkin Elmer-2400 spectrometer. 1H and 13C NMR spectra at room temperature were recorded in DMSO-d6 using Bruker Avance II 400 NMR spectrophotometer. EPR spectrum was recorded on E112 ESR spectrometer at room temperature. The FAB-Mass spectra were recorded on Joel SX-102 mass spectrometer. The electronic spectra in DMSO were recorded on Pye-unicam-8800 spectrophotometer at room temperature. Magnetic susceptibility measurements were carried out on a Faraday balance at 25 °C. The electrical conductivities of 10−3 solution in DMSO were obtained on a systronic type 302 Conductivity bridge equilibrated at 25 °C ± 0.01. Schiff base and its complexes were tested for antimicrobial activity by disk diffusion assay with slight modifications against some Gram-positive, Gram-negative strains of bacteria and some fungi. After that minimum inhibitory concentration (MIC) of the complexes was determined.

2.5

2.5 Organism culture and in- vitro screening

Antibacterial activity of ligand H2L and its Co(II), Ni(II), Cu(II) and Zn(II) complexes were completed by the disc diffusion method with slight alterations. Inoculums of Streptococcus mutans, Pseudomonas aeruginosa, Methicillin resistant Staphylococcus aureus (MRSA + Ve), Staphylococcus pyogenes, Salmonella typhimurium, and Escherichia coli were prepared in BHI medium and incubated for 18 h at 37 °C. Subsequently a suspension of about 105 cfu mL−1 was prepared in autoclaved saline solution according to the McFarland protocol. Then 10 μL of this saline suspension was mixed with10 mL of sterile antibiotic agar at 40 °C and then 9 cm diameter plates were poured in a laminar flow cabinet. Five paper discs of 6.0 mm diameter were placed on these nutrient agar plates. One milligrams of each test compound was dissolved in 100 μl DMSO to get ready stock solution and from this stock solution varied concentrations 10, 20, 25, 50, and 100 μg/μl of the trial compound were set. Subsequently the compound of varied concentrations was poured over disc plate onto it. The disc of chloramphenicol (30 μg) was used as a positive control and DMSO poured disc as a negative control. The susceptibility was assessed on the basis of diameter of zone of inhibition against Gram-positive and Gram-negative strains of bacteria and fungi. The Zone of inhibition (mm) of the ligand, H2L and its Co(II), Ni(II), Cu(II) and Zn(II) complexes against Gram-positive and Gram-negative strains of bacteria are shown in Table 1. The macro dilution test with standard inoculums of 105 cfu mL−1 was used to assess the minimum inhibitory concentration (MIC) of test ligand and its complexes. First of all, the ligand and its complexes were dissolved in dimethyl sulfoxide (DMSO) then serial dilutions of the test compounds were prepared to final concentrations of 512, 256, 128, 64, 32, 16, 8, 4, 2 and 1 mg/mL to all tube was added 100 mL of a 24 h old inoculums. The MIC is the lowest concentration of the trial compound, which can control the visible growth after 18 h incubation at 37 °C. The MIC was determined visually after incubation for 18 h, at 37 °C. After 18 h of incubation at 37 °C MIC results for bacterial strains are shown in table 2. For assaying antifungal activity, Candida albicans, Candida krusie, Candida parapsilosis and Cryptococcus neoformans were inoculated in Sabouraud Dextrose broth medium (Hi-Media Mumbai) and incubated for 24 h at 35 °C, and subsequently a suspension of about 106 cfu mL−1 was prepared in sterile saline solution according to the McFarland protocol. Autoclaved Sabouraud Dextrose Agar (SDA) was poured onto 9 cm diameter plates in laminar flow cabinet and then with the help of sterilized cotton swabs the suspension of each fungal cell was streaked onto Sabouraud Dextrose Agar (SDA) plates. Five paper discs of 6.0 mm diameter were put onto Sabouraud Dextrose agar plate. 1 mg of test compounds was dissolved in 100 μl DMSO to get ready stock solution and from this stock solution varied concentrations 10, 20, 25, 50, and 100 μg/μl of the trial compound were set. Subsequently the compound of varied concentrations was poured over disc plate onto it. The disc of fluconazole (30 μg) was use as a positive control and DMSO poured disk as a negative control. The susceptibility of different fungal strains against test compounds was assessed on the basis of diameter of zone of inhibition after 48 h of incubation at 35 °C Table 3. MIC results for fungal strains are shown in Table 4. The in vitro antifungal activity assay showed that [CuL] complex was more active antifungal than H2L and [ML] [M = Co(II), Ni(II) and Zn(II)].

Table 1 Antibacterial activity of H2L and its [ML] complexes.
Compounds Corresponding effect on microorganism
Gram positive bacteria Gram negative bacteria
S. mutans S. Pyogenes MRSA P. aeruginosa S. typhimurium E. coli
[H2L] 10.2 ± 0.3 11.9 ± 0.5 16.9 ± 0.3 11.2 ± 0.2 15.2 ± 0.4 12.3 ± 0.5
[CoL] 14.2 ± 0.3 12.9 ± 0.5 17.9 ± 0.3 13.2 ± 0.2 19.2 ± 0.4 15.2 ± 0.4
[NiL] 12 ± 0.5 10 ± 0.5 9.1 ± 0.3 13.2 ± 0.5 11.1 ± 0 15.2 ± 0.9
[CuL] 22 ± 0.5 21 ± 0.5 20 ± 0.2 17.5 ± 0.5 16.5 ± 0.5 20 ± 0.5
[ZnL] 13.1 ± 0.3 11.1 ± 0.5 14.3 ± 0.6 10.3 ± 0.6 17.8 ± 0.4 13.4 ± 0.5
Chloramp. 26.8 ± 0.5 22.4 ± 0.4 21 ± 0.5 17.1 ± 0.2 25.2 ± 0.8 20 ± 0.2
DMSO
Methicillin resistant Staphylococcus aureus.
Table 2 Minimum inhibition concentration (MIC), [H2L] and its [ML] complexes, positive control chloramphenicol.
MIC (μg/ml) Strains [H2L] [CoL] [NiL] [CuL] [ZnL] Positive control
S. mutans 100 50 100 25 100 32
S. pyogenes 100 25 100 25 50 32
MRSA 50 35 50 32 100 32
P. aeruginosa 50 25 50 25 50 32
S. typhimurium 50 25 50 25 50 32
E. coli 100 50 50 25 100 32
Methicillin resistant Staphylococcus aureus (MRSA + Ve).
Table 3 Antifungal activity of H2L and its [ML] complexes.
Compounds Corresponding effect on microorganism
CA CK CP CN
[H2L] 15.5 ± 0.4 13.3 ± 1.2 11.1 ± 0.4 10.2 ± 0.2
[CoL] 15.5 ± 0.4 14.4 ± 1.2 12.1 ± 0.4 11.2 ± 0.2
[NiL] 12.2 ± 0.3 11.5 ± 0.3 10.2 ± 1.2 8.9 ± 0.5
[CuL] 18.5 ± 0.5 18.3 ± 1.2 16.8 ± 0.2 13.7 ± 0.5
[ZnL] 16.1 ± 0.5 17.5 ± 0.2 14.9 ± 0.7 11.1 ± 0.5
Fluconazole 20 ± 0.5 20 ± 0.5 18 ± 0.5 19 ± 0.5
DMSO

CA, Candida albicans; CK, Candida krusei; CP, Candida parapsilosis; CN, Cryptococcus neoformans.

Positive control (fluconazole), and Negative control (DMSO) measured by the Halo Zone Test (Unit, mm).

Table 4 Minimum inhibition concentration (MIC), [H2L] and its [ML] complexes, positive control fluconazole.
Strains H2L [CoL] [NiL] [CuL] [ZnL] Positive Control
Candida albicans 100 25 100 25 32 1
Candida krusei 100 50 100 25 32 64
Candida parpsilosis 100 50 100 25 64 8
Cryptococcus neoformans 100 50 50 25 50 8

3

3 Results and discussion

Schiff base ligand, H2L was synthesized by the condensation of 1,8-diaminonaphthalene and glyoxalic acid in 1:2 M ratio dissolved in methanol (Scheme 1). The metal complexes were synthesized by template procedure {in the view of very less yield (33%) of ligand required for direct metallation with ligand} in 1:2:1 M ratio in methanol (Scheme 2). All complexes were stable at room temperature and soluble in DMSO. The formation of Schiff base ligand and its complexes was confirmed on the basis of results of elemental analysis, molecular ion peak in mass spectra, the characteristic bands in the FT-IR and resonance signals in the 1H NMR and 13C NMR spectra. The overall geometry of the complexes was inferred from the observed values of magnetic moments and the position of the bands in the electronic spectra. The molar conductance measurements of all the complexes recorded in DMSO, exhibited their non electrolytic nature. All efforts failed to grow single crystal suitable for X-ray crystallography. The analytical data along with some physico-chemical properties of Schiff base ligand and complexes are summarized in Table 5.

Synthesis and proposed structure of ligand.
Scheme 1
Synthesis and proposed structure of ligand.
Synthesis and proposed structure of Schiff base complexes.
Scheme 2
Synthesis and proposed structure of Schiff base complexes.
Table 5 Analytical and physical data of Schiff base ligand and its metal complexes.
Compound Empirical formula Analysis found (calc.) Yield (%) Colour MP (°C) Molar conductance (Ω−1 cm2 mol−1) m/z found(Calcd.)
C H N
H2L C14H10N2O4 65.59(65.00) 4.45(4.65) 12.9(12.14) 25 Yellow >300 270 (270.4)
[CoL] C14H8N2O4Co 51.24(51.44) 2.23(2.46) 8.11(8.50) 70 Pink >300 12 326 (326.9)
[NiL] C14H8N2O4Ni 51.33(51.43) 2.14(2.46) 8.01(8.57) 76 Green >300 13 326 (326.6)
[CuL] C14H8N2O4Cu 50.14(50.68) 2.07(2.43) 8.08(8.28) 75 Dark green >300 11 331 (331.5)
[ZnL] C14H8N2O4Zn 50.33(50.40) 2.16(2.41) 7.99(8.17) 83 Off white >300 09 333 (333.2)

3.1

3.1 Infrared spectra

The prominent bands observed in the IR spectra of Schiff base ligand, H2L and its complexes are listed in Table 6. The spectrum of free ligand shows a strong band at 1635 cm−1 assigned to ν(C⚌N), indicating the formation of desired Schiff base ligand. This band undergoes negative shift of ∼30 cm−1 on complexation indicating the involvement of azomethine nitrogen in chelation with metal ions in the complexes (Singh and Srivastava, 1988). This was further confirmed by the presence of a new band observed in 440–460 cm−1 region assigned to ν(M–N) (Ferraro, 1971). The intense band at 1664 cm−1 present in the IR spectrum of free ligand may be assigned to ν(C⚌O) of the carboxylic acid. However, this band was found to be absent in all the complexes. Instead, IR spectra of all complexes exhibit two bands at 1615–1635 cm−1 and 1370–1400 cm−1 regions attributed to νasym (CO2) and νsym (CO2) vibrations, respectively, indicating the involvement of the carboxylate oxygen atom in coordination to the metal ion. It is to be noted that Δ value [Δ = νasym (COO) − νsym(COO)] is larger than 200 cm−1 in all complexes indicating the monodentate coordination of the COO group to the metal ion (Deacon and Phillips, 1980) which is further confirmed by the presence of new band at 430–480 cm−1 assignable to ν(M–O) (Ferraro, 1971).

Table 6 IR spectral data (cm−1) for schiff base ligand and its complexes.
Compounds ν(C⚌N) ν(CO2)asym ν(CO2)sym ν(M–N) ν(M–O)
[H2L] 1635 1664
[CoL] 1607 1632 1376 553 478
[NiL] 1602 1628 1383 567 474
[CuL] 1590 1628 1371 570 473
[ZnL] 1607 1628 1378 573 447

3.2

3.2 1H NMR and 13C NMR spectra

The formation of Schiff base ligand, H2L was further supported by the 1H NMR spectral study. The 1H NMR spectrum of free ligand, H2L (Fig. 1) showed a resonance signals at 9.92 and 8.13 ppm assigned to carboxylic and azomethine protons of proposed Schiff base moiety, respectively. The spectrum of free ligand, exhibited a multiplet in 6.30–7.50 ppm region which may be assigned to naphthalene protons (Jeong et al., 2007). However the resonance signal assigned to carboxylic protons in free ligand disappeared in the 1H NMR spectrum of Zn(II) complex (Fig. 2) indicating the deprotonation of carboxylic proton and the involvement of oxygen of carboxylate ion in chelation (Garg et al., 2002). However, the position of resonance signals for azomethine proton and the aromatic protons undergo a slight downfield shift and appeared at 8.50 and 6.52–7.62 ppm, respectively on complexation with Zn(II) ion (Hage et al., 1988; Krishnapriya and Kandaswamy, 2005).

The 1H NMR spectrum of ligand.
Figure 1
The 1H NMR spectrum of ligand.
The 1H NMR spectrum of Zn(II) complex.
Figure 2
The 1H NMR spectrum of Zn(II) complex.

The 13C NMR spectrum of Zn(II) complex revealed the presence of expected number of signals corresponding to the different types of carbon atoms. A strong NMR signal appearing at 161 ppm (Shakir et al., 2009) may reasonably be to azomethine carbon and other at 194 ppm (Geeta et al., 2010) assigned to carboxylate carbon COO. The chemical shifts of naphthalene carbons appear at 138, 129, 127, 125, 123 and 120 ppm. These values were found to be downfield shifted by about 1.0–2.0 ppm as compared with free ligand (Shakir et al., 2009).

3.3

3.3 Electronic spectra and magnetic moment

The electronic spectrum of Co(II) complex showed bands at 17,064 cm−1 and 22,779 cm−1. The band at 22,779 cm−1 was assigned to π → π transitions, suggesting a square planar environment around Co(II) ion which was further supported by its magnetic moment value of 1.76 B.M. (Manhas et al., 1995; Loginova et al., 2008). The electronic spectrum of Ni(II) complex showed band at 28,260 cm−1 which may be assigned to charge transfer and other band at 22,280 cm−1 single d-d band is assigned to 1A2g ← 1A1g transition (Lever, 1984). The electronic spectrum of Cu(II) complex displays two bands at 16,000 cm−1 and 18,000 cm−1 which may reasonably be assigned to 2A1g ← 2B1g and 2Eg ← 2B1g transitions, respectively, consistent with a square planar geometry around Cu(II) ion which is further confirmed by its magnetic moment value of 1.82 B.M (Figgis, 1966).

3.4

3.4 EPR spectrum

EPR spectrum of polycrystalline Cu(II) complex (Fig. 3) has been recorded at room temperature. The absence of hyperfine signal may be attributed to the strong dipolar and exchange interaction between Cu(II) ions in the unit cell (Ahuja and Tripathi, 1991). The calculated g|| = 2.21 and g = 2.05 values reasonably support that the 2B1 is the ground state having an unpaired electron in the dx2–y2 orbital (Balhahausen, 1962). The g|| > g (2.32 > 2.05) suggests a square planar environment around Cu(II) ion (Ottaviani et al., 1994; Kivelson and Neiman, 1961). The G parameter G = (g|| − 2)/(g − 2) which measures the exchange interaction between the metal centres in polycrystalline solid has been calculated. According to Hathway if G > 4, the exchange interaction is negligible and if G < 4 considerable exchange interaction occurs in the solid complex in the complex (Hathaway et al., 1971). The G value is greater than 4 indicating the exchange interaction is negligible.

The EPR spectrum of Cu(II) complex.
Figure 3
The EPR spectrum of Cu(II) complex.

3.5

3.5 FAB mass spectroscopy

The FAB-mass spectrum of Schiff base ligand, [H2L] (Fig. 4) showed a molecular ion peak [M]+ at m/z 270 which corresponds to its proposed molecular formula. The molecular ion peaks for the schiff base complexes, [ML] [M = Co(II), Ni(II), Cu(II) and Zn(II)] appear at m/z 326, 326, 331 and 333, respectively, consistent with the proposed molecular formulae of the complexes. A reference spectrum of [NiL] has been shown in Fig. 5.

The FAB mass spectrum of ligand.
Figure 4
The FAB mass spectrum of ligand.
The FAB mass spectrum of Ni(II) complex.
Figure 5
The FAB mass spectrum of Ni(II) complex.

3.6

3.6 In vitro antimicrobial activity

The in vitro antibacterial and antifungal activity of ligand [H2L] and its complexes type, [ML], [M = Co(II), Ni(II), Cu(II) and Zn(II)], were tested using the bacterial cultures of S. mutans, P. aeruginosa, Methicillin resistant Staphylococcus aureus (MRSA + Ve), S. pyogenes, S. typhimurium, E. coli and fungal cultures of C. albicans, Candida krusei, C. parapsilosis and C. neoformans, by the disc diffusion method (Khan et al., 2007) and then the minimum inhibitory concentration (MIC) of the ligand and complexes was determined. Chloramphenicol (30 μg) was used as positive control in case of bacterial strains and in case of fungi; fluconazole was used as a positive control. While the disk poured in DMSO was used as negative control. The minimum inhibitory concentration (MIC) was assessed by the macro dilution test using standard inoculums of 105 cfu mL−1. Initially the compounds were dissolved in dimethyl sulfoxide (DMSO) after that serial dilution of the test compound was set to final concentrations of 512, 256, 128, 64, 32, 16, 8, 4, 2 and 1 mg/mL to each tube was added 100 mL of 24 h old inoculums. The MIC is the lowest concentration of the test compound, which can restrain the apparent growth after 18 h incubation at 37 °C. The MIC was determined visually after incubation for 18 h, at 37 °C. The in vitro study results verified that the Cu(II) complex was found to be more active antimicrobial agent than ligand H2L and Co(II), Ni(II) Zn(II) complexes. Zone of inhibition was visualized after 18 h of incubation at 37 °C. The susceptibility was assessed on the basis of diameter of zone of inhibition against Gram-positive and Gram-negative strains of bacteria and fungi. The zones of inhibition (mm) of ligand and its complexes against Gram-positive and Gram-negative strains of bacteria are shown in Table 3. Zones of inhibition for fungal strains are shown in Table 3. MIC results for bacterial strains are shown in Table 2 and for fungal strains in Table 4.

The higher antimicrobial activity of the metal complexes as compared to Schiff base ligand may be explained in terms of chelation which makes metal complexes to act as more powerful and potent antimicrobial agents, thus inhibiting the growth of the microorganisms (Berejo et al., 1999; Chohan et al., 2003). Moreover, coordination reduces the polarity of the metal ion mainly because of the partial sharing of its positive charge with the donor groups within the chelate ring system. This process, in turn, increases the lipophilic nature of the central metal atom, which favours its permeation more efficiently through the lipid layer of the microorganism, thus destroying them more aggressively.

The Cu(II) complex showed enhanced activity. The distinct difference in the antimicrobial property of compound further justifies the purpose of this study. The importance of such work lies in the possibility that the new compound might be more effective against bacteria for which a thorough investigation regarding the structure activity relationship, toxicity and in their biological effects which would be helpful in designing more potent antibacterial agents for therapeutic use if required.

4

4 Conclusion

A Schiff base ligand, H2L derived from the condensation of 1,8 diaminonaphthalene and glyoxalic acid was synthesized and characterized. The metal complexes with Co(II), Ni(II), Cu(II) and Zn(II) ions were prepared by the template synthesis due to very poor yield of ligand. The bonding of the ligand in the complexes and the overall geometry have been deduced on the basis of various spectroscopic techniques. The comparative in vitro antimicrobial results suggested that the Cu(II) complex shows a significant antimicrobial activity as compared to ligand, H2L and its Co(II), Ni(II) and Zn(II) complexes.

Acknowledgement

This work was supported by King Saud University, Deanship of Scientific Research, College of Science Research Centre.

References

  1. , , . Indian J. Chem. 1991;30A:1060.
  2. , , , , . Current Clinical Topics in Infectious Diseases. Blackwell Sciences; . pp. 18, 19
  3. , , . Int. J. Antimicrob. Agents. 2000;16:5.
  4. , , . Chem. Rev.. 2001;101:37.
  5. , , , , . J. Appl. Catal. A. 2006;301:169.
  6. , . Introduction to Ligand Field Theory. New York: Mcgraw Hill; .
  7. , , , , , , , . Polyhedron. 1999;18:3695.
  8. , , , , , , , , , , , . J. Med. Chem.. 1996;39:673.
  9. , , , , , . J. Virol.. 2004;78:11321.
  10. , , , . J. Enzyme Inhib. Med. Chem.. 2003;18:259.
  11. , , , , , . J. Chem. Soc. Dalton Trans. 1998:3489-3494.
  12. , . Chem. Soc. Rev.. 2004;33:410.
  13. , , , , , , , , , , , , , . Antifungal Agents: Discovery and Mode. Oxford: BIOS; . p. 143
  14. , , . Coord. Chem. Rev.. 1980;33:227.
  15. , , . Chemotherapy. 2000;12:463.
  16. , . Low frequency vibrations of inorganic and coordination compounds. NewYork: Plenum press; .
  17. , . Introduction to Ligand Field Theory. New York: Interscience, Plenum press; . p. 276
  18. , , . Tetrahedron. 1997;53:7201.
  19. , , . Chem. Rev.. 2000;100:2771.
  20. , , , , , . Spectrochim Acta A.. 2002;58:2789.
  21. , , , . Coord. Chem. Rev.. 1993;126:1.
  22. Geeta, B., Shravankumar, K., Reddy, P.M., Ravikrishna, E., Sarangapani, M., Krishna, K., 2010. Spectrochim. Acta A 77, 911.
  23. , , , , , , , . Inorg Chem.. 1988;27:2185.
  24. , , , . Essays in Chemistry. New York: Academic Press; .
  25. , , , . Appl. Catal.. 1998;168:353.
  26. , , , , , eds. Comprehensive asymmetric catalysis. Berlin: Springer; . p. 64
  27. , , , , . Bull. Kor. Chem. Soc.. 2007;28:851-854.
  28. , , , . Chem. Phys. Lett.. 2000;324:43.
  29. , , , . Eur. J. Med. Chem.. 2007;42:103.
  30. , , , , . Bull. Korean Chem. Soc.. 2007;29:538.
  31. , , . J. Chem. Phys.. 1961;35:149.
  32. , , . Polyhedron. 2005;24:113.
  33. , , , , . Trans. Met. Chem.. 2008;33:643.
  34. , , , . J. Am. Chem. Soc.. 1993;115:5326.
  35. Loginova, N.V., Koval’chak, T.V., Osipovick, N.P., Polozov, G.I., Sorokin, V.I., Chernyavskaya, Shadrurs O.I., 2008. Polyhedron 27, 985.
  36. , , , , , , , . J. Mol. Catal. A. 2006;150:62.
  37. , , , . Adv. Synth. Catal.. 2002;344:477.
  38. , , , . Polyhedron. 1995;14:3549.
  39. , , , , , , . J. Appl. Microbiol.. 2009;107:1279.
  40. , , , , , , . J. Antimicrob. Chemother.. 2009;63:337.
  41. , . Chem. Rev.. 1989;89:1513.
  42. , , , . Bull. Chem. Soc. Jpn.. 1966;39:1471.
  43. , , . Clin Infect Dis.. 2005;41:521.
  44. , , , , . J. Am. Chem. Soc.. 1994;116:661.
  45. , , . Stud. Surf. Catal.. 1996;101:184.
  46. , , , , , , . Spectrochimica Acta Part A.. 2009;72:591.
  47. , , . Metal-catalysed oxidation of organic compounds Academic press. New York; .
  48. , , . Synth .Reac. Inorg. Met. Org. Chem. 1988;18:515.
  49. , , , . J. Inorg. Nucl. Chem.. 1981;43:1701.
  50. , , , , . J. Acta Chem. Scand. A.. 1995;51:896.
  51. , , , . Curr Med Chem.. 2006;13:1321.
  52. , , , , , , , , , . Bioorg. Med.Chem.. 2000;8:157.
  53. , , , , . Synth. React. Inorg. Met. Org. Chem.. 1995;25:417.
  54. , , , , . Science.. 1997;277:936.
  55. , , , , , , . J.Infect. Dis.. 1998;168:165.
  56. , , , , , , . Trans. Met. Chem.. 2001;26:307.
  57. , . Coord. Chem. Rev.. 1999;190:537.
  58. , , . Synlett. 1995:829.
  59. , , , , . Polyhedron.. 1993;12:1097.
  60. , , , . J. Catal.. 2006;238:369.
Show Sections