Sie sind auf Seite 1von 7

Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 113 (2013) 408414

Contents lists available at SciVerse ScienceDirect

Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy


journal homepage: www.elsevier.com/locate/saa

Group 12 dithiocarbamate complexes: Synthesis, spectral studies and their use as precursors for metal suldes nanoparticles and nanocomposites
Peter A. Ajibade , Benjamin C. Ejelonu
Department of Chemistry, University of Fort Hare, Private Bag X1314, Alice 5700, South Africa

h i g h l i g h t s
 Three dithiocarbamate complexes of

g r a p h i c a l a b s t r a c t

group 12 metal ions were synthesized and characterized.  Complexes characterized by spectroscopic techniques and elemental analyses.  Complexes used as precursors to prepare metal sulde nanoparticles/ PMMA nanocomposites.  Nanoparticles sizes varied between 3.03 and 23.45 nm.

a r t i c l e

i n f o

a b s t r a c t
Zn(II), Cd(II) and Hg(II) dithiocarbamate complexes have been synthesized and characterized by elemental analysis, thermogravimetric analysis, UVVis, FTIR, 1H- and 13C NMR spectroscopy. The complexes were thermolysed at 180 C and used as single molecule precursors for the synthesis of HDA capped ZnS, CdS and HgS nanoparticles and polymethylmethacrylate (PMMA) nanocomposites. The optical and structural properties of the nanoparticles and nanocomposites were studied by UVVis, PL, XRD and SEM. The crystallites sizes of the nanoparticles varied between 3.03 and 23.45 nm. SEM and EDX analyses of the nanocomposites conrmed the presence of the nanoparticles in the polymer matrix. 2013 Elsevier B.V. All rights reserved.

Article history: Received 5 January 2013 Received in revised form 14 April 2013 Accepted 24 April 2013 Available online 7 May 2013 Keywords: Dithiocarbamate complexes Synthesis Spectral studies Precursors Nanoparticles Nanocomposites

Introduction Dithiocarbamate complexes are known to possess striking structural features, diversied industrial and biological applications [1]. The dithiocarbamate complexes of elements with d10 congurations represent a large and interesting group of inorganic compounds that have been extensively studied in recent years [25].
Corresponding author. Tel.: +27 40 602 2055; fax: +27 40 602 2094.
E-mail address: pajibade@ufh.ac.za (P.A. Ajibade). 1386-1425/$ - see front matter 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.saa.2013.04.113

They have been found as useful precursors for the deposition of II/VI compound semiconductor materials because of their reasonable volatility and less carbon deposition as impurity [68]. Nanoparticles sizes range between 1 and 100 nm and they have wide applications in catalysis, electronic, optical and magnetic because of their unprecedented chemical and physical properties. They are also used in light emitting devices, solar cells and bio-imaging [9]. The use of CdS nanoparticles in modern technology as light-emitting diodes, solar cells, and optical devices based on its non-linear optical properties has been reported. Also, ZnS semiconductor has

P.A. Ajibade, B.C. Ejelonu / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 113 (2013) 408414

409

band gap energy of 3.7 eV; and it is said to possess such potential applications in modern technology as in solar cells, infrared windows, sensors and displays, bio-imaging, and catalysis [9]. It has been reported that nanocomposites containing two or more different nanoscale functionalities with controlled structure and interface interactions possess novel properties than individual components [10]. The polymer matrix in metal-lled polymer nanocomposites provides such qualities as processibility, performance and solubility or stabilizes the systems thermally [11]. In this study, we present the synthesis and characterization of Zn(II), Cd(II) and Hg(II) complexes of some dithiocarbamate ligands and their use as precursors for the preparation of ZnS, CdS and HgS nanoparticles and their poly methyl methacrylate nanocomposites. Experimental Material and physical measurements Ammonia solution, butyl amine, ethyl phenyl amine; and the respective metal salts- zinc(II) chloride, cadmium(II) chloride and mercury(II) chloride used in this study were purchased from Merck and used without further purication. The thermogravimetric analysis was performed using a Perkin Elmer thermogravimetric analyzer (TGA 7) tted with a thermal analysis controller (TAC 7/DX). A ow of nitrogen was maintained with a heating rate of 10 C min1 between ambient temperature and 900 C. About 1016 mg mass of each of the samples was loaded into an alumina cup and weight changes were recorded as a function of temperature. The FTIR determination of the metal complexes was done as KBr discs on a Perkin Elmer Paragon 2000 FTIR spectrophotometer in the range 4000370 cm1. The 1H and 13C NMR determination of the metal complexes was done using 400 MHz and 101 MHz Bruker NMR spectrophotometers respectively. UVVis spectra of the metal complexes and the metal sulde nanoparticles were recorded using Perkin Elmer Lambda 25 spectrophotometer in chloroform from 800 to 200 cm1. Samples for scanning electron microscope (SEM) were prepared as follows: the respective nanoparticles were mounted on stub using carbon double-sided tape. They were coated with Au/Pd using the Eiko IB. 3 Ion Coater. They were observed using JOEL JSM-6390 LVSEM at a rating voltage of 1520 kV at different magnications as indicated on the SEM images. The photoluminescence measurement of the nanoparticles was done using a Perkin Elmer LS 45 Fluorimeter while powder X-ray diffraction patterns of the nanoparticles were recorded on a Bruker D8 Advanced, equipped with a proportional counter using Cu Ka radiation (k = 1.5405 A, nickel lter). Synthesis of ammonium butyl dithiocarbamate (L1) 6.00 mL (0.10 mol) of carbon disulde was added into a mixture of 9.88 mL of butylamine (0.10 mol) and 25 mL concentrated aqueous ammonia and stirred for 1 h. The cream colored precipitate (which turned white when dried) was washed with diethyl ether and dried in a vacuum over sodium hydroxide and phosphorus pentoxide. The white ammonium butyl dithiocarbamate (L1) obtained is air and temperature sensitive. Yield: 4.92 g (59.28%). Synthesis of ammonium N-alkyl-N-phenyl dithiocarbamate (L2)

three times with ice cold ethanol three times. The yellowish white ligand is temperature and air sensitive. Yield: 7.133 g (66.70%). Synthesis of the metal complexes All the metal complexes were prepared following the same method: a mixture of 1.25 mmol (0.268 g) of N-ethyl-N-phenyl dithiocarbamate and 1.25 mmol (0.208 g) of butyl dithiocarbamate, each dissolved in 20.00 mL of distilled water was added into a 20.00 mL distilled water solution of the respective metal ions [ZnCl2 (0.170 g); CdCl2 (0.229 g); HgCl2 (0.339 g]. The resulting solutions were then stirred for 1 h; and the solid products obtained for the individual metal complexes were rinsed with distilled water and dried over CaCl2. The yields of the metal complexes are as follow: [(C4H9)NHCS2HgS2CN(C2H5)C6H5] [0.31 g (65.98%)]; [(C4H9)NHCS2CdS2CN(C2H5)C6H5) [0.42 g (79.22%)]; and [(C4H9)NHCS2HgS2CN(C2H5) C6H5] [0.38 g (58.19%)]. Synthesis of the HDA capped metal suldes nanoparticles The metal sulde nanoparticles were prepared from the respective metal complexes. In a typical synthesis 0.70 g of each metal complex was dissolved in TOP (10 mL) and injected into hot HDA (7 g) at 180 C. An initial decrease in temperature (about 20 30 C) was observed. The solution was stabilized and the reaction was continued for 1 h at 180 C. After completion, the reaction mixture was allowed to cool to 70 C and methanol was added to precipitate the nanoparticles. The solid was separated by centrifugation and washed three times with methanol. The resulting solid precipitates of HDA-capped metal sulde (MS) nanoparticles were dispersed in toluene for further analysis. Synthesis of the metal suldes/PMMA nanocomposites The respective nanocomposites were synthesized by dissolving about 2.0 g of the polymer in 10.0 mL of toluene in a beaker and adding 10.0 mL toluene solution of each metal sulde nanoparticles into it. This was done slowly with heating and vigorous stirring. The solution became turbid after approximately 20 min; yellow solution was obtained for CdS/PMMA nanocomposites while colorless solutions were obtained for ZnS/PMMA and HgS/ PMMA nanocomposites. Results and discussions Synthesis The metal complexes are all air-stable and soluble in dichloromethane, chloroform, but insoluble in methanol, hexane etc. The poor solubility of the mercury(II) complex in both polar and nonpolar solvents may be attributed to the possibility of the complex existing as polymers. The polymer formation by the complex might be due to the fact that dithioacid complexes of the ML2 type are coordinative unsaturated, which leads to the formation of higher coordination number either by adding one or two molecules of Lewis base or by polymerization [12,13]. The elemental analyses of the complexes are presented in Table 1. Thermal analysis of the complexes

Ammonium N-ethyl-N-phenyl dithiocarbamate (L2) was prepared according to the procedure given in the literature [12]. In a typical synthetic procedure, a mixture of 6.44 mL of ethylaniline and 15.00 mL of concentrated aqueous ammonia in ice was added into 3.00 mL of ice-cold carbon disulde and the resultant solution stirred for 67 h. The solid product obtained was ltered and rinsed

The thermal properties of the complexes were determined by TGA in the temperature range of 20800 C under nitrogen atmosphere. The content of a given compound changes when it is subjected to a temperature change in a controlled atmosphere. The mass loss curve of such a compound gives its mass losses as a

410

P.A. Ajibade, B.C. Ejelonu / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 113 (2013) 408414

Table 1 Elemental analysis of the metal complexes. Metal complexes Formula weight Analytical data calculated (found) (%) Carbon [(C4H9)NHCS2ZnS2CN(C2H5)C6H5] [(C4H9)NH CS2CdS2CN(C2H5)C6H5] [(C4H9)NH CS2HgS2CN (C2H5)C6H5] 407.98 456.98 546.02 41.18 (40.88) 36.80 (35.39) 30.77 (30.99) Hydrogen 4.94 (5.20) 6.13 (6.26) 3.69 (3.89) Nitrogen 6.86 (7.10) 4.41 (4.06) 5.13 (5.45)

Fig. 1. TGA curves showing the degradation pattern of the metal complexes.

ligands, the m(CAN) bond that appears in the range 1509 1488 cm1 shifted to 15091488 cm1 in the metal complexes. This suggests an increase in the CAN double bond characters as a result of the delocalization of electrons towards the metal centre after due to coordination to the dithiocarbamate ligands [16]. In the ligands, the m(CAS) stretching vibrations were observed as two bands in the region 1060940 cm1. These bands occurred as single sharp bands in the complexes at about 990 cm1. This conrms the coordination of the ligands to the metal ions as bidentate chelating ligands through the sulfur atoms. The presence of the m(CN) and m(CSS) absorption bands in the metal complexes conrm the presence of dithiocarbamate ligands in the complexes [19]. It has been observed in previous studies that the coordination modes of alkyl-aryl dithiocarbamate ligands with group 12 metals are mostly through the sulfur atoms [2022]. This is further supported by the singular absorptions which appeared at about 1000 70 cm1 region (Zn complex at1003 cm1; Cd complex at 993 cm1; and Hg complex at 992 cm1) in the FTIR spectra of the metal complexes [16]. The characteristic frequencies of the aromatic groups in the complexes, resulting from the out-of-plane bending vibrations of the aromatic CAH are said to appear in the range 600900 cm1 [23]. The (@CAH) bending modes of the aromatic ring was seen to absorb at around 700 cm1 in the ethyl phenyl dithiocarbamate free ligand; and at about 695, 695, and 696 cm1 respectively in the Zn, Cd, and Hg complexes. The characteristic absorption peak of NAH of the butyl dithiocarbamate ligand is seen to appear at 3281 cm1 in the free ligand and between 3208 and 3263 cm1 in the metal complexes, which is slightly below the literature value of 33003500 cm1 [24]. Electronic spectra of the metal complexes The electronic spectra of the complexes are shown in Fig. 2. Three characteristic absorption bands are known for dithiocarbamate complexes due to the chromospheres NS2 caused by p ? p transition of NCS and SCS moieties; and g ? p transition arising from transition of an electron of the lone pair electrons on the sulfur atom to an antibonding p-orbital [1]. In the metal complexes an intense band around 291294 nm was observed due to the p ? p transition of the phenyl ring [25]. The aniline group is also said to show a secondary band at 285 nm [26]; this band was observed between 281 and 286 in the metal complexes. The d10 electronic conguration of the pseudo-transition (d10) metals is suggestive why no dd transition over the visible region was observed [1]. NMR spectra of the metal complexes In the 1H NMR spectra of the metal complexes two signals were observed as multiplets around 7.447.30 ppm, which corresponds to the aromatic protons of the phenyl ring of the ethyl phenyl moiety [27]. The methyl protons of the butyl group experienced up eld signals due to more shielding effect of the nuclei of the carbon chain, and it appeared as triplets about 0.99 ppm; while the CH3 protons of the ethyl group appeared around 1.30 ppm. The peaks

function of temperature or time [14]. Initial decomposition for the metal complexes starts around 130 C, 125 C, and 89 C for the Zn, Cd, and Hg complexes respectively (Fig. 1). At a temperature of about 223334 C about 41% of the mass of the complexes lost. Heating the metal complexes further results into much more gradual loss of mass, with about 22.85% of the mass left for the Zn complex; 30.85% for the Cd complex; and 0.5% for the Hg complex- all at 800 C. The products of the thermal decomposition of the metal complexes correspond to their respective oxides, but for Hg where volatilization results as observed in the thermogram. The Hg complex initial decomposition at 30 C (0.78% mass loss) suggests the escape of entrapped solvent molecule within its matrix [15]. The calculated and experimental masses could be said to agree favourably (Table 2). The mercury complex is seen to have the least stability [16,17], the stability trend of the complexes is in the order: Hg < Cd < Zn. Infrared spectra studies of the metal complexes The infrared spectra of the ligands and complexes were compared and assigned on careful comparison. In dithiocarbamate complexes, three regions of great importance are: 1580 1450 cm1 due to m(CAN) of NCS; 1060940 cm1 due to m(CAS) of CSS and 420250 cm1 due to the MAS bond [16,18]. In the free
Table 2 Thermal decomposition data of the metal complexes. Compounds Decomposition ranges (C) DTG max value (C) Weight loss (%)

Decomposition reaction

Product expected

Mass changes (mg) Cal. Found 3.63 4.20 0.61

Zn Cd Hg

130752 125340 89469

181 177 144

76.98 68.24 98.89

[(C4H9)NHCS2HgS2CN(C2H5) C6H5] ? ZnS ? ZnO [(C4H9)NHCS2CdS2CN(C2H5) C6H5] ? CdS ? CdO [(C4H9)NHCS2HgS2CN(C2H5) C6H5] ? HgS volatilization

ZnO CdO HgS

3.63 4.20 0.61

P.A. Ajibade, B.C. Ejelonu / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 113 (2013) 408414

411

attributed to the signal of the NA(CH2) of the ethyl group which is more deshielded than the NA(CH2) of the butyl ring. The CS2 for the ZnL1L2, CdL1L2 complexes were observed at 207.87, 205.19 and 207.55, 205.44 ppm respectively. Optical properties of the metal suldes nanoparticles The absorption and emission of ZnS, CdS and HgS nanoparticles are given in Fig. 3. The peak of the spectra of the nanoparticles is 288 nm (ZnS); 291 nm (CdS) and 285 nm (HgS), which are at shorter wavelengths than that of bulk ZnS, CdS and HgS (334, 516 and 620 nm for ZnS, CdS and HgS bulks respectively) as a result of quantum connement effects. This observation suggests a decrease in the particle size of the respective nanoparticles compared to the bulk materials. The band gap in a nanomaterial has been obtained from the absorption maxima of their spectra. Using the fundamental absorption edges in the samples (absorption maxima), their band gaps are calculated thus; ZnS = 4.31 eV (288 nm), CdS = 4.26 eV (291 nm) and HgS = 4.35 eV (285 nm). The band gap energies of the bulk samples are also calculated to be: ZnS = 3.71 eV (334 nm), CdS = 2.40 eV (516 nm) and HgS = 2.0 eV (620 nm) [28,29]. Two emission peaks due to the exciton and the trapped luminescence are said to exist for semiconductor nanocrystals. The exciton emission is usually observed as sharp band while the trapped emission appears as broad band [29]. The ZnS, CdS and HgS nanoparticles were found to exhibit emission maxima at

Fig. 2. Electronic spectra of the metal complexes.

at about 4.20 (quartets) and 3.50 ppm (triplets) are assigned to the NACH2 protons of the ethyl and butyl groups respectively. The methylene protons NA(CH2)2 peaks were seen at around 1.70 and 1.40 ppm in the complexes. The NH proton peak appeared around 1.50 ppm. In the 13C NMR spectra, two different signals were observed for the phenyl ring attached to the N-ethyl group. The rst signal appeared around 145.91 ppm; while the second signal with higher intensity was observed as three peaks between 129.14 and 126.86 ppm. This is assigned to the N-butyl group. The methylene carbon peaks of the complexes were observed as four signals between 54.69 and 19.50 ppm. The peak observed at 78.92 ppm is

Fig. 3. Absorption and emission spectra of the nanoparticles: (A) absorption (a) and emission (b) spectra of ZnS nanoparticle; (B) absorption (a) and emission (b) spectra of CdS nanoparticle and (C) absorption (a) and emission (b) spectra of HgS nanoparticle.

412

P.A. Ajibade, B.C. Ejelonu / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 113 (2013) 408414

Fig. 4. SEM/TEM images of ZnS (A), CdS (B) and HgS (C) nanoparticles.

Fig. 5. Powder XRD pattern of ZnS nanoparticle (A); powder XRD pattern of CdS nanoparticle (B) and powder XRD pattern of HgS nanoparticle (C).

P.A. Ajibade, B.C. Ejelonu / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 113 (2013) 408414 Table 3 Values of crystallite sizes for ZnS, CdS and HgS nanoparticles. Metal sulde nanoparticles ZnS Angle (degree) observed maxima 28.86 47.94 56.94 26.86 44.17 52.29 20.18 23.40 33.93 39.90 Full weight half maxima (FWHM) 1.41 1.57 1.32 0.72 1.76 1.84 0.38 0.59 0.45 0.49

413

Grain size from XRD (nm) 6.45 6.15 3.03 12.65 5.42 5.35 23.45 15.20 20.41 19.27

CdS

HgS

393, 391 and 389 nm respectively. As shown in Fig. 3, a shift was observed for all the nanoparticles with respect to absorption edge. Fig. 4 shows the SEM/TEM pictures of samples ZnS (A); CdS (B) and HgS (C) nanoparticles. They all gave a somewhat spherical morphology, which agglomerated to form larger particles; thus giving them a dense surface morphology [2931]. The nanoparticle

gave different shapes and sizes under the TEM analysis. Dots-, ricelike- and spherical- shapes were observed for the ZnS, CdS and HgS nanoparticles respectively. Their sizes ranged between 2.78 and 14.54 nm [29]. The XRD patterns of the metal sulde nanoparticles are presented in Fig. 5. The XRD patterns reveal their nanocrystalline

Fig. 6. SEM/TEM micrographs of the respective nanocomposites: PMMA (A); ZnS/PMMA (B/B); CdS/PMMA (C/C) and HgS/PMMA (D/D).

414

P.A. Ajibade, B.C. Ejelonu / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 113 (2013) 408414

natures, which indicate the broadening of the diffraction peaks due to the higher surface to volume ratio [28]. ZnS nanoparticle showed broad peaks at 2h = 28.86, 47.95 and 56.94, which can be indexed to (1 1 1), (2 2 0) and (3 1 1) planes of cubic zinc blende structure (JCPDS 05-0566) Fig. 3 [28,29]. The three peaks observed in the XRD patterns of CdS nanoparticle correspond to (1 1 1), (2 2 0) and (3 1 1) of the cubic crystal phase of CdS (JCPDS 5-0566) [32]. The XRD patterns (1 1 1), (2 0 0), (2 2 0) and (3 1 1) obtained for HgS nanoparticle are in agreement with the pattern JCPDS 00-006-0261, which correspond to HgS metacinnabar, syn [29]. The ZnS and CdS samples appeared to be of much more smaller particle sizes than that of HgS as indicated by the observed relatively broader diffraction peaks for the ZnS and CdS nanoparticle XRD patterns compared to that of HgS nanoparticle (Fig. 5AC). Table 3 contains the values of crystallite sizes for ZnS, CdS and HgS nanoparticles. Structural properties of the metal sulde/PMMA nanocomposites SEM was used to study the morphology and the extent of interaction of the ZnS, CdS and HgS nanoparticles with PMMA. Fig. 6A shows the SEM image of PMMA, while Fig. 6BD shows the SEM images of ZnS/PMMA, CdS/PMMA and HgS/PMMA respectively. The SEM monograph obtained for the polymer (PMMA) shows a well-dened shape of spherical aggregates [33,34]. The pronounced spherical aggregates observed in PMMA micrograph were somehow lost (dissolved) in the PMMA interaction with different nanoparticles; thus suggesting an interfacial interaction between the polymer and the metal sulde nanoparticles. The TEM images revealed that some interaction occurred between the inorganic nanoparticle materials and the polymer, which has been observed, could help to increase the stability of the pure PMMA polymer and hence, enhance its usability [33]. The chemical composition of the metal sulde/polymer nanocomposites were studied by EDX, which conrmed the presence of Zn, S; Cd, S and Hg, S from ZnS, CdS and HgS nanoparticles respectively. Some broadening of the diffraction peaks and the decrease of the intensity of the spectra were observed in the XRD patterns of the nanocomposites. This may be attributed to the pore lling effects, which is said to reduce the scattering contrast between the pores and the frame work PMMA materials [3537]. Infrared spectral studies of the nanocomposites The FTIR data revealed that the absorption band patterns of the ZnS/PMMA, CdS/PMMA and HgS/PMMA nanocomposites are similar to those of the pure PMMA. However, a slight shift was observed in their stretching frequencies suggesting the possibility of an interaction between the metal sulde nanoparticles and the polymer. This possible interaction between the inorganic and organic phases could inuence the thermal stability of the pure PMMA [32,33]. Conclusion Zn(II), Cd(II) and Hg(II) complexes of some alkyl-phenyl dithiocarbamates were synthesized and characterized by elemental analysis, TGA, UVVis, FTIR, 1H- and 13C NMR spectroscopy. Spectroscopic analyses conrmed the coordination of the metal ions to the dithiocarbamate anions acting as bidentate chelating ligands. The complexes were used as single molecular precursors to synthesize ZnS, CdS and HgS nanoparticles and PMMA nanocom-

posites. The crystallites sizes of the nanocomposites varied between 3.03 nm and 23.45 nm. EDX analysis from the SEM of the nanocomposites conrmed the presence of the nanoparticles in the polymer matrix and the broadening observed in the XRD of the nanocomposites of the nanocomposites also conrmed the formation of the nanocomposites. Acknowledgements The author acknowledged with thanks nancial support from Govan Mbeki Research and Development Centre, University of Fort Hare. BCE acknowledged Adekunle Ajasin University, Akungba Akoko, Nigeria for study leave to undertake this study. References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] D.C. Onwudiwe, P.A. Ajibade, B. Omondi, J. Mol. Struct. 987 (2011) 5866. N. Srinivasan, P.J. Rani, S. Thirumaran, J. Coord. Chem. 62 (2009) 12711277. H. Nabipour, Int. J. Nano Dim. 1 (2011) 225232. H. Nabipour, S. Ghammamy, S. Ashuri, Z.S. Aghbolagh, Org. Chem. J. 2 (2010) 7580. S.R. Trifunovi, Z. Markovi, D. Sladi, K. Andjelkovi, T. Sabo, D. Mini, J. Serb. Shem. Soc. 67 (2002) 115122. P. O Brien, R. Nomura, J. Mater. Chem. 5 (1995) 17611773. O.C. Montteiro, T. Trindale, J.H. Park, P. O Brien, Chem. Vap. Depos. 6 (2000) 230232. I.W. Lenggoro, H. Moon Lee, K. Okuyama, J. Colloid Interface Sci. 303 (2006) 124130. M.L. Hassan, A.F. Ali, J. Cryst Growth 310 (2008) 52285252. M.S. Bakshi, P. Thakur, S. Sachar, T.S. Banipal, Mater. Lett. 61 (2007) 3762 3767. Nanocomposites: Metal & Ceramic Filled Polymers for Dielectrics, <www.ngimat.com>. S. Thirumaran, K. Ramalinam, G. Bocelli, A. Cantoni, Polyhedron 19 (2000) 12791282. S. Thirumaran, K. Ramalinam, G. Bocelli, A. Cantoni, Polyhedron 18 (1999) 925930. Thermographic Analysis A Beginners Guide, <http://www.perkinelmer.com/ CMSResources/Images/>. S. Khan, S.A.A. Nami, K.S. Siddiqi, J. Organomet. Chem. 693 (2008) 10491057. D.C. Onwudiwe, P.A. Ajibade, Int. J. Mol. Sci. 12 (2011) 19641978. S. Kumar, N.K. Kaushik, J. Therm. Anal. 21 (1981) 37. M. Shahid, T. Ruffer, H. Lang, S.A. Awan, S. Ahmad, J. Coord. Chem. 62 (2009) 440445. N. Adwang, N.S.A.M. Yousof, I. Bada, N.F. Rajab, A. Hamid, B.M. Yamin, et al., World Appl. Sci. J. 9 (2010) 804810. A. Gossage, H.A. Jenkins, Acta Chim. Slov. 56 (2009) 329333. A. Howie, E.R.T. Tiekink, J.L. Wardell, S.M.S.V. Wardell, J. Chem. Crystallogr. 39 (2009) 293298. D. Ondudruova, M. Pajtova, E. Jna, M. Koman, Solid State Phenom. 9091 (2003) 383388. S. Sunitha, K.K. Aravindakshan, Int. J. Pharm. Biomed. Sci. 2 (2011) 108113. J. McMurry, Organic Chemisrty, seventh ed., Thomson Books/Cole, United States, 2008. C.Su.N. Tang, M. Tan, X. Gan, L. Cai, Synth. React. Inorg. Met.-Org. Chem. 27 (1997) 291300. S. Kumar, Organic ChemistrySpectroscopy of Organic Compounds, Amritsar, <http://nsdl.niscair.res,in/bitsream/123456789/793/1/ spectroscopy+of+Organic+compounds.pdf>, 2006, p. 16. D.C. Onwudiwe, P.A. Ajibade, Synth. React. Inorg. Met. -Org. Nano-Met. Chem. 40 (2010) 279284. B. Barman, K.C. Sarma, Opto. Adv. Mater. 4 (2010) 15941597. D.C. Onwudiwe, P.A. Ajibade, Int. J. Mol. Sci. 12 (2011) 55385551. M. Salavati-Niasari, F. Davar, M. Mazaheri, Mater. Res. Bull. 44 (2009) 2246 2251. A.U. Ubale, S.G. Ibrahim, Inter, J. Mater. Chem. 2 (2012) 5764. , Colloid Polym. T. Radhakrishnan, M.K. Georges, P.S. Nair, A.S. Luyt, V. Djokovic Sci. 286 (2008) 683689. N. Moloto, N.J. Coville, S.S. Ray, M.J. Moloto, Physica B 404 (2009) 44614465. Sh. Sohrabnezhad, A. Pourahmad, J. Alloys Compd. 505 (2010) 324327. A. Pourahmad, Sh. Sohrabnezhad, M.S. Sadjadi, K. Zare, Mater. Lett. 62 (2008) 655658. vez, F.F. Castillo n-Barrazad, M. Flores-Acosta, M. Sotelo-Lermac, H. Arizpe-Cha rez-Bona, Solid State Commun. 128 (2003) 407411. R. Ram M. Sathish, B. Viswanathan, R.P. Viswanath, Int. J. Hydrogen Energy 31 (2006) 891898.

[27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37]

Das könnte Ihnen auch gefallen