Journal of Molecular Catalysis A: Chemical

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Molecular Catalysis A: Chemical 380 (2013) 94103

Contents lists available at ScienceDirect

Journal of Molecular Catalysis A: Chemical


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

Synthesis, crystal structure and spectroscopic studies of a cobalt(III) Schiff base complex and its use as a heterogeneous catalyst for the oxidation reaction under mild condition
Sk. Manirul Islam a, , Anupam Singha Roy a , Sasanka Dalapati b,1 , Rajat Saha c,1 , Paramita Mondal a , Kajari Ghosh a , Saptarshi Chatterjee d , Keka Sarkar d , Nikhil Guchhait b , Partho Mitra e
a

Department of Chemistry, University of Kalyani, Kalyani, Nadia 741235, W.B., India Department of Chemistry, University of Calcutta, Kolkata 700009, W.B., India c Department of Physics, Jadavpur University, Jadavpur, Kolkata 700032, W.B., India d Department of Microbiology, University of Kalyani, Kalyani, Nadia 741235, W.B., India e Department of Inorganic Chemistry, IACS, Kolkata 700032, W.B., India
b

a r t i c l e

i n f o

a b s t r a c t
A new Schiff base oriented Co(III) complex, (CoL)Cl 4H2 O (L = Schiff base), has been synthesized by careful design of the Schiff base ligand thorough oxidation of Co(II) to Co(III) during the reaction process. The complex has been characterized by single-crystal X-ray structure analysis and various spectral analyses. Structure analysis reveals that this monomeric complex crystallizes in triclinic P1 space group. Supramolecular hydrogen bonding interactions among the guest water molecules and counter anion chloride leads to the formation of 1D water-chloride chain. A polymer anchored heterogeneous cobalt complex has been synthesized, characterized by various physicochemical techniques and successfully used for the oxidation of alkenes and suldes using H2 O2 as oxygen source. The inuence of the various reaction parameters has been studied. The heterogeneous cobalt complex can be reused seven times without any signicant loss in its catalytic activity. 2013 Elsevier B.V. All rights reserved.

Article history: Received 18 March 2013 Received in revised form 15 September 2013 Accepted 16 September 2013 Available online 25 September 2013 Keywords: Cobalt complex Polymer anchored Heterogeneous catalyst Oxidation Sulfoxide H2 O2

1. Introduction Schiff base ligands have attracted enormous attention for the synthesis of several homo- and hetero-metallic complexes. The use of such Schiff base ligands for the synthesis of 3d-metal complexes has been receiving considerable attention due to their important catalytic, magnetic and biological properties. Among 3d metal ions, Co(II) has signicant inuence in catalytic, unusual magnetic and biological properties. Between its two oxidation states +2 and +3, later is more important in these respects [114]. Oxidation of olens is especially interesting because it gives oxygen containing value added products like alcohols, aldehydes, ketones, acids, epoxides, etc. which are extremely important and useful in both chemical and pharmaceutical industries [15]. Sulfoxides and sulfones are useful synthetic intermediates for the

Corresponding author. Tel.: +91 33 2582 8750; fax: +91 33 2582 8282. E-mail address: manir65@rediffmail.com (Sk.M. Islam). 1 These authors have equal contribution. 1381-1169/$ see front matter 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.molcata.2013.09.022

synthesis of several important organic compounds [16,17]. Sulfoxides and sulfones are generally prepared via oxidation of the corresponding suldes. Cobalt(II) salts and its complexes have been used as homogeneous catalysts in the oxidation reaction. Cobalt(III)-peroxo/alkylperoxo species are considered as intermediates in these catalytic processes [18,19]. Thus, the direct use of Co(III) may enhance the rate of reaction. With this in mind, several Co(III) catalytic systems have been developed though the research on Co(III) system is still in infant stage [2023]. Immobilization of metal complexes onto the surface of solid supports is highly desirable in the development of reusable catalysts [2433]. Surface functionalization and incorporation of metal complexes is one of the general methods for the preparation of supported catalysts. In recent years, polymer especially chloromethyl polystyrene has opened a new opportunities for application in the eld of supported catalysts [3439]. Among the variety of catalytically active metal complexes the transition metals bearing N, O-donor ligands are of great interest and several works have been devoted to investigate their catalytic activity in various catalytic organic reactions. The chemical modication of a polymer

Sk.M. Islam et al. / Journal of Molecular Catalysis A: Chemical 380 (2013) 94103

95

Scheme 1. General procedure for the synthesis of (CoL)Cl 4H2 O.

incorporating bifunctional ligand such as N, O donor, has however received less attention. In the present work, we describe the synthesis of a new Co(III) complex by using a Schiff base ligand (L) through oxidation of a Co(II) salt to Co(III). Structural studies indicate that the central Co(III) ion adopts octahedral geometry in the complex with formation of 1D water-chloride chain. A polymer anchored cobalt complex has been synthesized and characterized by various physicochemical techniques. Its catalytic activity was evaluated for the oxidation of alkenes and suldes with H2 O2 under mild reaction conditions. The condition of maximum conversion and selectivity for the desired product was optimized. The catalytic property for a recycled catalyst was also evaluated. 2. Experimental 2.1. Materials and instruments All reagents and solvents were purchased from commercial sources and were further puried by standard procedures. Chloromethylated polystyrene, alkenes and suldes were obtained from SigmaAldrich chemicals Company, USA. Triethylenetetramine, Cobalt(II) chloride hexahydrate, 30% H2 O2 and 2hydroxyacetophenone were obtained from Merck and used as received without further purication. The synthesized materials were characterized by 1 H NMR, single-crystal X-ray diffraction (XRD), thermogravimetric analysis (TGA), Fourier transform infrared (FT-IR) spectroscopy and scanning electron microscopy (SEM). 1 H NMR spectra of the Schiff base ligand was recorded with a Bruker AMX-400 NMR spectrophotometer. A mass spectrum of the products was recorded on a Trace GC ultra DSQ II. The elemental analyses (C, H and N) were performed with a PerkinElmer model 240C elemental analyzer. A simultaneous TGA-DTA of the complex and polymer anchored heterogeneous complex was carried out on a Mettler Toledo TGA/SDTA 851 instrument. The experiments were performed in N2 at a heating rate of 10 C min1 in the temperature range 30600 C. The FT-IR spectra of the samples were recorded on a PerkinElmer FT-IR 783 spectrophotometer using KBr pellets. Electronic spectra were registered on a Shimadzu UV/3101 PC spectrophotometer. Cobalt content of the sample was measured by Varian AA240 atomic absorption spectrophotometer (AAS). Morphology of

functionalized polystyrene and complex was analyzed using a scanning electron microscope (SEM) (ZEISS EVO40, England) equipped with EDX facility. 2.2. Synthesis of Schiff base ligand (L) The outline for the preparation of Co(III) Schiff base complex is given in Scheme 1. A 20 mL solution of N,N -bis-(2aminoethyl)ethylenediamine (10 mmol) in methanol was added to a solution of 2-hydroxyacetophenone (20 mmol) in 20 mL methanol. The resulting solution was reuxed for 4 h on a water bath to give a dark yellow solution. Then the resulting mixture was poured in crushed ice whereby a yellow precipitate was obtained. It was ltered off, washed several times with water, recrystalized from ethanol and nally dried. C22 H30 N4 O2 (3.99 g, 88%): calcd. C 69.11, H 7.85, N 14.66 (%); found C 69.01, H 7.89, N 14.39 (%). Mass spectrum (EI): m/z 382 (M+ = L+ ). 1 H NMR (CDCl3 , ppm): = 16.11 (s, 2 H, phenolic OH), 7.526.90 (m, 8 H), 3.673.71 (m, 4 H), 2.78 (t, 4 H), 2.61 (t, 4 H), 2.34 (s, 6 H, methyl), 2.21(s, 2H, amine). 13 C NMR (CDCl3 , ppm): = 163.3, 160.2, 131.5, 130.7, 123.5, 120.8, 116.8, 49.9, 46.3, 43.5, 14.5. 2.3. Synthesis of cobalt(III) complex (CoL)Cl 4H2 O To a pale yellow solution of Schiff base ligand (1 mmol) in ethanol (30 mL), an ethanolic solution of NEt3 (2 mmol) was slowly added. An ethanolic solution of CoCl2 6H2 O (1 mmol) was added dropwise with magnetic stirring. The reaction mixture was then reuxed for another 2 h. The resulting solution was then cooled down to room temperature and kept unperturbed for the slow evaporation of the solvent. Pink colored single crystal suitable for X-ray analysis was obtained after several weeks from the mother liquor which were isolated by ltration. C22 H36 ClCoN4 O6 (0.4 g, 70%): calcd. C 48.32, H 6.60, N 10.25, Co 10.78 (%); found C 48.17, H 6.49, N 10.19, Co 10.62 (%). 2.4. Preparation of polymer anchored cobalt complex To a 250 mL round bottom ask equipped with a magnetic stirrer bar and containing DMF (100 mL), were added chloromethylated polystyrene (2 g, 5.5 mmol/g of Cl) and Schiff base (5.5 mmol), and the reaction mixture was stirred for 24 h at 100 C. The reaction

96

Sk.M. Islam et al. / Journal of Molecular Catalysis A: Chemical 380 (2013) 94103

H3C

N OH

N H

N H

N HO

CH3

Polymer support P CH2Cl DMF 100 0C H 3C N N N N CH3

OH HO (2) EtOH CoCl2 6H2O NEt3 Polymer anchored cobalt complex (3)

Scheme 2. General procedure for the synthesis of polymer anchored cobalt complex.

mixture was ltered and washed thoroughly with DMF, and dried in vacuo for 12 h. The Schiff base-functionalized polymer 2 (2 g) was treated with ethanol (50 mL) for 30 min. An ethanolic solution of CoCl2 6H2 O (1% w/v) was added, and the resulting mixture was heated to 70 C for 6 h. The resulting polymer, impregnated with the metal complex, was ltered and washed with ethanol to obtain PS-TETA-Co 3 (Scheme 2). PS-TETA (2): C 77.87, H 7.32, N 8.46 (%), PS-TEAT-Co (3): C 73.59, H 6.71, N 7.93, Co 5.86 (%). 2.5. General procedure for oxidation reaction The liquid phase oxidation reactions were carried out in a twonecked round bottom ask tted with a water condenser and placed in an oil bath at different temperatures under vigorous stirring for a certain period of time. Substrates (5 mmol) were taken in CH3 CN solvent (5 mL) for different sets of reactions together with 2.5 102 mmol catalyst in which 10 mmol of H2 O2 (30% in aq.) was added. Product analysis was performed using Varian 3400 gas chromatograph equipped with a 30 m CP-SIL8CB capillary column and a Flame Ionization Detector. All reaction products were identied by using Trace DSQ II GCMS. 2.6. Single-crystal X-ray diffraction study Suitable single crystal of (CoL)Cl 4H2 O has been mounted on a BrukerAXS SMART APEX II diffractometer equipped with a radiation. The graphite monochromator and Mo K ( = 0.71073 A) crystal was positioned at 60 mm from the CCD. 360 frames were measured with a counting time of 5 s. The structure was solved using Patterson method by using the SHELXS 97. Subsequent difference Fourier synthesis and least-square renement revealed the positions of the remaining non-hydrogen atoms. Non-hydrogen atoms were rened with independent anisotropic displacement parameters. Hydrogen atoms were placed in idealized positions and their displacement parameters were xed to be 1.2 times larger than those of the attached non-hydrogen atom. Successful convergence was indicated by the maximum shift/error of 0.001 for the last cycle of the least squares renement. Absorption corrections were carried out using the SADABS program [40]. All calculations were carried out using SHELXS 97, SHELXL 97, PLATON 99, ORTEP32 and WinGX system Ver-1.64 [4145]. 3. Results and discussion The synthesized Schiff base ligand (L) was characterized by FTIR, and 13 C NMR and elemental analysis. All spectral data agreed with the ligand structure. The cobalt(III) complex (CoL)Cl 4H2 O has been synthesized by a general procedure based on the mixing of an ethanolic solution of CoCl2 6H2 O with an ethanolic solution of the

Schiff base ligand in a 1:1 molar ratio (Scheme 1). The complex is stable to air and moisture without any kind of decomposition. Atomic absorption spectroscopy suggested 5.86% (0.994 mmol/g) of cobalt in the polymer anchored complex (Scheme 2). Elemental analyses data show that the loading of ligand on polymer support is 1.42 mmol/g. Therefore, metal to ligand loading in polymer complex is close to 1:1. 3.1. Crystal structure of (CoL)Cl 4H2 O X-ray diffraction studies reveal that the complex is comprised of a mononuclear cationic complex unit, (CoL)+ along with chloride anion for electroneutrality (ORTEP view, Fig. 1). In the complex, the central Co(III) ion adopts an approximate octahedral geometry. The distortions in the coordination spheres of metal ions from the ideal geometries are obvious due to the structural constraints imposed by the polydentate ligand framework. The imine nitrogen atoms bind in trans fashion with each other in the complex. It is evident from the cisoid and transoid angles around the Co(III) ion in complex being in the range of 82.97(8)-96.46 (9) and 175.45(8)178.04 (8), respectively. The trans bonds are close to the ideal 180 (ideal octahedral geometry). The CoN and CoO bond distances agree well with similar other Schiff ranging from 1.876 to 1.9488 A base complexes of cobalt(III) [4648]. Data collection, structure renement parameters and crystallographic data for the complex are given in Table 1. The selected bond lengths, bond angles and non-covalent interaction parameters are summarized in Tables S1 and S2 (Supporting information). 3.2. Supramolecular structures of (CoL)Cl 4H2 O The asymmetric unit contains four guest water molecules and one charge neutralizing chloride ions (Fig. 1). Among these three guest water molecules (O2W, O3W and O4W) and the chloride anion (Cl ) take part in the formation of 1D water-chloride chain through hydrogen bonding interactions (Fig. 2). During formation of such 1D supramolecular chain structure, two alternating cyclic tetrameric assembly are formed (Fig. 2). Between these two, in one case only water molecules (O3W and O4W) take part and in other case chloride ions and water molecules (O2W) takes part, which are further connected by hydrogen bonding interactions leading to the 1D supramolecular water-chloride channels. The one rest water molecule (O1W) is hanged from the 1D supramolecular chains toward the metal-organic moieties and connects them through hydrogen bonding interactions. The metal-organic moieties are also connected to the water-chloride chains through other hydrogen bonding interactions and thus form 3D supramolecular architecture (Fig. 3).

1H

Sk.M. Islam et al. / Journal of Molecular Catalysis A: Chemical 380 (2013) 94103

97

Fig. 1. ORTEP diagram of complex (guest water and hydrogen atoms are removed for clarity).

Fig. 2. 1D water-chloride H-bonded channel.

3.3. IR spectral study FT-IR spectra of Schiff base ligand (L) and the cobalt complex (CoL)Cl 4H2 O are given in Fig. S1 (supporting information). The sharp peak at 1614 cm1 in the spectrum of Schiff base is characteristic of the C N functionality of the ligand [49]. This band is shifted to lower region in complex and appeared at 1596 cm1 . It suggested the coordination of the Schiff base to the central Co(III) ion through the azomethine nitrogen. A strong band in the region 33003400 cm1 in Schiff base is observed due to NH (secondary amine) and (OH) stretching vibration. The phenolic (C O) stretching frequency is observed at 1282 cm1 (in ligand L) which shifted to higher frequency at 1322 cm1 in cobalt complex, indicating coordination of metal through phenolic oxygen [50]. Bands at 3467 and 1542 cm1 are observed corresponding to (OH) stretching and bending vibrations from lattice water molecule in the cobalt complex [49]. The N H (secondary amine) bending vibration is observed at 1491 cm1 in L but in the complex this band is absent. In the case of immobilized complex, the FT-IR spectra (Fig. S2 in supporting information) of polymer, polymer supported Schiff base ligand and polymer supported cobalt complex were studied. The intensity of the IR spectrum of polymer supported metal complex was weak due to the low concentration of the complex. The sharp C Cl peak due to CH2 Cl group in polymer at 1266 cm1 was weak after loading of Schiff base on the support [51]. A band at 1662 cm1 in polymeric support is observed due to C N bond [52]. Above two IR data conrm the loading of Schiff base (L) on the polymer matrix. Here, the N H (secondary amine) bending vibration, which

Fig. 3. 3D supramolecular architecture of the complex (CoL)Cl 4H2 O.

98

Sk.M. Islam et al. / Journal of Molecular Catalysis A: Chemical 380 (2013) 94103

Table 1 Crystal data and details of the structure determination. Crystal data Formula Formula weight Crystal system Space group a () b () c () ( ) ( ) ( ) V (3 ) Z D (calc) (g/cm3 ) (Mo K) (/mm) F(0 0 0) Crystal size (mm) Data collection Temperature (K) Radiation () minmax ( ) Dataset Tot. Uniq. data R (int) Observed data [I > 2.0 sigma (I)] Renement Nref Npar R wR2 S Max. and av. shift/error Min. and max. resd. dens. (e/3 ) C22 H28 CoN4 O2 ,4(H2 O), Cl 546.93 Triclinic P1 (no. 2) 10.0877 (13) 10.3763 (13) 13.0627 (17) 71.238 (2) 70.839 (2) 72.802 (2) 1195.0(3) 2 1.520 0.876 576 0.08 0.10 0.12 293 Mo K 0.71073 1.7, 26.5 12: 12; 12: 12; 16: 16 12,524 4889 0.030 4464 4889 307 0.0395 0.1158 1.06 0.00, 0.00 0.65, 0.50

1.5

Absorbance

1.0

0.5
(a) (b)

0.0 200 300 400 500 600 700 800

Wavelength (nm)
Fig. 4. DRS-UVvis absorption spectra of (a) polymer anchored Schiff base ligand and (b) polymer anchored cobalt complex.

+ 0.7871P] where P = (Fo2 + 2Fc2 w = 1/[\s2 (Fo2 ) + (0.0645P)2 )/3.

in acetonitrile solution. This result indicates that the complex maintains its geometry in solution phase. The reectance spectra of the polymer anchored Schiff base ligand and heterogeneous cobalt complex are given in Fig. 4. Polymer anchored Schiff base shows several absorption bands at ca 230, 260, 320 and 415 nm, respectively due to * transition and the intraligand n* transition. These absorptions also present in the spectrum of the cobalt complex but shifted which indicates the coordination behavior of the polymer anchored Schiff base ligand to the metal ion. Polymer anchored cobalt complex shows an additional band at 625700 nm which can be assigned to dd transition [54]. In the recycled catalyst all types of transition bands were present which indicates its heterogeneous nature (Fig. S5). 3.6. Scanning electron micrographs (SEM) and energy dispersive X-ray analyses (EDAX) Field emission-scanning electron micrographs of single bead of pure chloromethylated polystyrene, polymer anchored ligand (PS-TETA) and its complex PS-TETA-Co were recorded and it was observed that the morphological changes occurred on the polystyrene beads at various stages of the synthesis. The SEM images of polymer anchored ligand (A) and the immobilized cobalt complex (B) on functionalized polymer are shown in Fig. 5. The pure chloromethylated polystyrene bead has a smooth surface (Fig. S6, supporting information) while polymer anchored ligand and complex show slight roughening on the top layer of polymer beads. This roughening is relatively more in complex. Also the presence of cobalt metal can be further proved by energy dispersive spectroscopy analysis of X-rays (EDAX) (Fig. S7, supporting information) which suggests the formation of metal complex with the polymer anchored ligand. 3.7. TGA studies To conrm further the presence of water molecules in the structure, thermal analyses of the complex (CoL)Cl 4H2 O was investigated using TGA-DTA at a heating rate of 10 C min1 in N2 atmosphere over a temperature range of 30600 C. TG curve (Fig. 6) shows that mass loss started at 50 C. This mass loss is observed from 50 to 140 C. It is clear from the curve that the mass loss of the complex up to 140 C is 12.34% (calcd. 13.16%) corresponds to four lattice water molecules. This complex is stable up to 245 C and after that it decomposed. Results of DTA analysis of the complex show a broad endothermic peak at 115 C. This

is observed at 1491 cm1 in L, is absent. This indicates that the polymer is attached to the Schiff base through secondary amine N atom [53]. In polymer anchored ligand, the phenolic (C O) stretching frequency is observed at 1369 cm1 . After complexation, the bands due to C N and phenolic (C O) are shifted to lower and higher frequency region, respectively. In Fig. S2 (d), the FT-IR spectrum of the recycled PS-TETA-Co catalyst is shown which suggests the existence of all the properties in the recycled catalyst. 3.4. Electronic spectroscopy The electronic spectra of both the Schiff base ligand (L) and its cobalt complex have been measured in acetonitrile and are shown in Fig. S3 (Supporting information). Ligand absorption bands are observed at 215, 250 and 320 nm, respectively which are assigned to * transition and the intraligand n* transition. Similar kind of absorption bands are also observed in cobalt complex. [(CoL)Cl 4H2 O] also shows dd transition band at 583 nm with charge transfer band at 363 nm [49]. The shifting of the spectrum and appearance of new bands clearly indicate the complexation behavior of the ligand to the central cobalt(III) ion. 3.5. DRS-UV spectroscopy The solid state electronic spectra of the complex (CoL)Cl 4H2 O and polymer anchored cobalt complex were recorded in diffuse reectance spectrum mode as MgCO3 /BaSO4 disc. The UVvis spectrum (DRS) of the cobalt complex, (CoL)Cl 4H2 O is added in Fig. S4 (Supporting information). Solid state electronic spectrum of the complex (CoL)Cl 4H2 O is almost identical as spectrum

Sk.M. Islam et al. / Journal of Molecular Catalysis A: Chemical 380 (2013) 94103

99

Fig. 5. FE-SEM images of polymer anchored Schiff base ligand (A) and polymer anchored cobalt complex (B).

3.8. Photoluminescence spectra of complex These cobalt complexes are potential photoluminescence materials because their extended structure and ligand dependent emissions. Here, the emission spectra of the neat cobalt complex (A), polymer anchored Schiff base (B) and polymer anchored cobalt complex (C) were studied in solid state at room temperature, as shown in Fig. 7. These metal complexes exhibit an emission band at around 420480 nm upon excitation at 363 nm. This emission band can be attributed to the ligand-to-metal charge transfer (LMCT) emission [55,56]. When excited at 363 nm, the polymer anchored ligand showed an emission maximum at 480 nm (B), which shifted to lower wavelength region and showed maximum at 455 nm upon binding to the cobalt metal ions (C) (quantum yield = 0.038). Thus, it is evident that the uorescence emission intensity of the ligand decreases upon complex formation. The interaction between the metal and ligand due to complexation results in a reduction of photoluminescence intensities.

Fig. 6. TGA-DTA plot for (CoL)Cl 4H2 O complex.

endothermic dip in the DTA curve corresponds to the removal of the lattice water molecules from the crystal. TGA curve of polymer anchored cobalt complex is shown in Fig. S8 (Supporting information). Here mass loss started at 250270 C and this is observed up to 420 C. Polymer anchored cobalt complex decomposed at 420430 C. So from the thermal stability, it concludes that polymer anchored metal complex degraded at considerably higher temperature.

3.9. Catalytic activity Since polymer anchored metal systems exhibit catalytic activity in a wide range of industrially important processes and have been extensively studied, we have decided to investigate the catalytic activity of polymer anchored cobalt complex in the oxidation of alkenes and suldes with H2 O2 as oxygen source in acetonitrile.

Fig. 7. The photoluminescence spectra of complex (CoL)Cl 4H2 O (A), polymer anchored Schiff base ligand (B) and polymer anchored cobalt complex (C).

100

Sk.M. Islam et al. / Journal of Molecular Catalysis A: Chemical 380 (2013) 94103

CHO Polymer anchored cobalt catalyst H2O2 / acetonitrile 50 0C Styrene

100 80 Conversion (%) Selectivity (%)

+
60

%
40 20 0 0

Benzaldehyde

Styrene oxide

Scheme 3. Oxidation product of styrene.

3.9.1. Oxidation of alkenes with H2 O2 catalyzed by PS-TETA-Co For optimization of reaction conditions, we choose oxidation of styrene as a probe reaction. Generally, oxidation of styrene gives styrene oxide, benzaldehyde, acetophenone and benzoic acid. However, in this case benzaldehyde is formed as the major product (Scheme 3). 3.9.1.1. Effect of the mole ratio of styrene to H2 O2 In order to determine the effect of H2 O2 on the oxidation of styrene to benzaldehyde, the mole ratio of styrene to H2 O2 was varied, keeping other parameters are xed: catalyst (2.5 102 mmol), temperature (50 C) and reaction time (8 h). The results are shown in Fig. 8. Styrene to H2 O2 molar ratios of 1:0.5 and 1:1 resulted in 33% and 64% conversion, respectively and when styrene to H2 O2 molar ratio was changed to 1:2, conversion increased to 97% keeping all other conditions similar. However, conversion and selectivity were found to decrease for styrene to H2 O2 molar ratios of 2:1 and 1:3. Therefore, a 1:2 molar ratio of styrene to H2 O2 was found to be the optimum. 3.9.1.2. Effect of the amount of catalyst The amount of catalyst has a signicant effect on the oxidation of styrene. Five different amounts viz., 1 102 , 1.5 102 , 2 102 , 2.5 102 and 3 102 mmol, were used keeping all other reaction parameters are xed: styrene (5 mmol), 30% H2 O2 (10 mmol), temperature (50 C) and reaction time (8 h). The results are shown in Fig. S10; 74, 83, 94, 97, 97% conversion corresponding to 1 102 , 1.5 102 , 2 102 , 2.5 102 and 3 102 mmol of catalyst, respectively. The lower conversion of styrene into benzaldehyde with 1 102 and 1.5 102 mmol of catalyst may be due to fewer catalytic sites. The maximum conversion was observed with 2.5 102 mmol of catalyst, but there was no noticeable difference in the conversion when 3 102 mmol of catalyst was employed. Therefore, 2.5 102 mmol of catalyst was taken as optimal.

10

Reaction time (h)


Fig. 9. Effect of the reaction time on oxidation reaction of styrene catalyzed by polymer anchored cobalt complex. Reaction condition: [styrene] = 5 mmol, H2 O2 (10 mmol), acetonitrile (5 mL), temperature (50 C).

3.9.1.3. Effect of reaction temperature The effect of temperature on the performance of the catalyst was studied at ve different temperatures, viz. 30 (room temperature), 40, 50 and 60 C at xed amount of styrene (5 mmol), H2 O2 (10 mmol), catalyst (2.5 102 mmol) in 5 mL CH3 CN (Fig. S11). A maximum of 97% conversion was achieved on carrying out the reaction at 50 C. At lower temperature, the conversion was low. At 60 C, the initial conversion of styrene was high than at 50 C but when reaction continues the conversion of styrene was almost same with conversion at 50 C. Therefore 50 C was the optimum temperature and all the catalytic oxidation reactions of olens were carried out at this temperature. 3.9.1.4. Effect of the reaction time The effect of time on the oxidation of styrene was studied by performing the oxidation reaction of styrene (5 mmol) with 30% H2 O2 (10 mmol) in the presence of 2.5 102 mmol of catalyst at 50 C. The results are given in Fig. 9. It is seen from the gure that the initial conversion of styrene was low, with an increase in the reaction time, conversion also increases. This is because some time is required for the formation of an intermediate which is converted to products. It is seen that maximum conversion was observed within 8 h. The optimum conditions for maximum conversion of styrene to benzaldehyde are: mole ratio of styrene to H2 O2 of 1:2, with 2.5 102 mmol of catalyst and 8 h reaction time at 50 C. The above optimized reaction conditions can be applied to the oxidation reaction of other olens using polymer anchored cobalt catalyst and the results are shown in Table 2. This catalyst efciently converts olens to their corresponding allylic products with H2 O2 . In the oxidation of cyclohexene, allylic products 2-cyclohexene1-one and 2-cyclohexene-1-ol were produced. Substituted styrene produced selectively corresponding aldehydes. Acetophenone was detected in the oxidation of -methyl styrene as major product. trans-stilbene was also oxidized by this heterogeneous catalyst in high yields. trans-stilbene gives benzaldehyde as a major product with benzil. Oxidation of olens was also carried out with (CoL)Cl 4H2 O complex under the above optimized reaction conditions and the results are given in Table 2. From the results it is seen that the present polymer anchored cobalt catalyst is more active than the homogeneous cobalt complex in the oxidation of olens with H2 O2 . 3.9.2. Oxidation of suldes with H2 O2 catalyzed by PS-TETA-Co In continuation of our research interest in oxidation, we have investigated the catalytic activity of the polymer anchored cobalt catalyst in sulde oxidation (Scheme 4). Thus when methyl phenyl

100 80 60

Conversion (%) Selectivity (%)

%
40 20 0

1:0.5

1:1

1:2

1:3

2:1

Mole ratio (styrene:H2O2)


Fig. 8. Effect of molar ratio of styrene to H2 O2 on oxidation reaction of styrene catalyzed by polymer anchored cobalt complex. Reaction condition: amount of catalyst = 2.5 102 mmol, acetonitrile (5 mL), temperature (50 C) and time (8 h).

Sk.M. Islam et al. / Journal of Molecular Catalysis A: Chemical 380 (2013) 94103 Table 2 Oxidation of alkenes using H2 O2 a . Entry Substrate Conversion (%)b PS-TETA-Co (CoL)Cl 4H2 O 97 81 88 73 62 44 89 75 86 68 77 64 79 67 Product (selectivity %)b PS-TETA-Co/(CoL)Cl 4H2 O Benzaldehyde (93/87) Styrene oxide (7/13) 2-cyclohexene-1-one (68/61) 2-cyclohexene-1-ol (28/30) Cyclohexene epoxide (4/9) Benzaldehyde (74/70) Benzil (17/18) trans-stilbene oxide (9/12) 4-Methyl benzaldehyde (88/83) 4-methylstyrene epoxide (12/17) 4-Chloro benzaldehyde (86/82) 4-chlorostyrene epoxide (14/18) 4-Nitro benzaldehyde (80/77) 4-Nitro styrene oxide (20/23) Acetophenone (100/100)

101

TON PS-TETA-Co (CoL)Cl 4H2 O 194 162 176 146 124 88 178 150 172 136 154 128 158 134

1 2

Styrene Cyclohexene

3 4 5 6 7

trans-stilbene 4-Methylstyrene 4-Chlorostyrene 4-Nitrostyrene -Methyl styrene

TON = moles of substrate converted per mole of catalyst (Co loading 0.994 mmol/g of catalyst). a Reaction conditions: catalyst (2.5 102 mmol), substrate (5 mmol), acetonitrile (5 mL), H2 O2 (10 mmol), temperature (50 C), time (8 h). b Determined by GC. The conversion and TON data of both the immobilized PS-TETA-Co catalyst and its homogeneous (CoL)Cl 4H2 O complex are given in two parallel rows, rst row for immobilized catalyst and the second one for homogeneous cobalt complex.

sulde was treated with polymer anchored cobalt catalyst and 30% H2 O2 in acetonitrile at room temperature, methyl phenyl sulfoxide was obtained as major product. Considering methyl phenyl sulde as a model substrate, it was subjected to different reaction conditions as given in Table S3. The effect of temperature was investigated for the oxidation of methyl phenyl sulde. While reaction temperature increased from room (Table S3, run 1) to 50 C (Table S3, run 2) there was a slight increase in conversion but conversion kept constant on further increase in temperature (Table S3, run 3). Amount of H2 O2 has a signicant effect on the conversion and selectivity of methyl phenyl sulde oxidation reaction. Conversion increased when amount of H2 O2 increased from 5 mmol (Table S3, run 4) to 10 mmol (Table S3, run 1) but there was no signicant change in conversion when amount changed to 15 mmol (Table S3, run 5) at the same reaction conditions. The selectivity of sulfoxide decreased with the increase of the amount of H2 O2 . This is due to further oxidation of sulfoxide. The reaction was also monitored with different catalyst amount. At a lower catalyst amount (1.5 102 mmol, Table S3, run 6) the reaction took longer time and was not complete. While with 2.5 102 mmol of catalyst (Table S3, run 1) the reaction was complete in 6 h, yielding 94% sulfoxide with a small amount of sulfone (6%). It was also observed that by increasing the amount of catalyst to 4.0 102 mmol (Table S3, run 7), the reaction led to more sulfone along with sulfoxide. Therefore, an optimum amount of catalyst and oxidant with a suitable solvent are necessary for the above oxidation and based on the above observation we come to a conclusion that 2.5 102 mmol of polymer anchored catalyst, 10 mmol 30% H2 O2 and acetonitrile as a solvent make for the best condition for the oxidation of suldes to sulfoxides with good selectivity. It was also found that without H2 O2 , the catalyst alone cannot oxidize suldes (Table S3, run 8). To examine the reactivity of the catalyst further, oxidation of other suldes also has been investigated (Table 3). Substrate scope

was extended to diphenyl sulde, ethyl phenyl sulde and dibutyl sulde etc. The sulfoxides were selectively obtained in all cases. A series of substrates, aryl alkyl, aryl allyl and dialkyl suldes, can be oxidized to their corresponding sulfoxides. The reactivity and conversion were dependent on the nature of the substituent. In the case of allyl suldes, small oxidation was observed at the carboncarbon double bond. Similarly, benzylic suldes can be oxidized to the corresponding sulfoxides with small oxidation products of the benzylic C H bond. Oxidation of suldes was also carried out with (CoL)Cl 4H2 O complex under optimized reaction conditions and the results are given in Table 3 for comparison with the polymer anchored cobalt complex. From the results it is seen that the present polymer anchored cobalt catalyst is more active than the homogeneous cobalt complex in the oxidation of suldes with H2 O2 . 3.10. Test for heterogeneity A heterogeneity test was carried out for the oxidation of styrene. For the rigorous proof of heterogeneity, a test was carried out by ltering the catalyst from the reaction mixture at 50 C after 3 h, and then the ltrate was allowed to react up to 6 h. The reaction mixture after 3 h and the ltrate were analyzed by gas chromatography. No change in percent conversion as well as percent selectivity was found indicating that the present catalyst acts as a heterogeneous catalyst. 3.11. Recycling of catalyst The catalyst remains insoluble in the present reaction conditions and hence can be easily separated by simple ltration followed by washing. The oxidation of styrene and methyl phenyl sulde was carried out with the recycled catalyst under optimized reaction conditions. The catalyst was removed from the reaction mixture

O S CH3 PS-TETA-Co H2O2/acetonitrile RT S CH3

O S CH3

O methyl phenyl sulfone

methyl phenyl sulfide

methyl phenyl sulfoxide

Scheme 4. Oxidation product of methyl phenyl sulde.

102 Table 3 Oxidation of suldes using 30% H2 O2 a . Substrates

Sk.M. Islam et al. / Journal of Molecular Catalysis A: Chemical 380 (2013) 94103

Conversion (%)b PS-TETA-Co (CoL)Cl 4H2 O 86 71 90 76 89 73 83 69 80 66 73 62 87 72 88 76

Selectivity of sulfoxide/sulfone (%)b PS-TETA-Co (CoL)Cl 4H2 O 79/21 93/7 91/9 89/11 84/14 83/14 77/23 90/10 71/29 82/18 79/21 77/23 73/22 67/25 62/38 78/22

TON PS-TETA-Co (CoL)Cl 4H2 O 172 142 180 152 178 146 166 138 160 132 146 124 174 144 176 152

C4 H9 SC4 H9 PhSPh PhSC2 H5 PhSC6 H13 PhSCH2 Ph PhSCH2 CH CH2 p-ClC6 H4 SCH3 p-CH3 C6 H4 SCH3

TON = moles of substrate converted per mole of catalyst (Co loading 0.994 mmol/g of catalyst). a Conditions: sulde (5 mmol); 30% aq. H2 O2 (10 mmol); acetonitrile (5 mL); catalyst (2.5 102 mmol), time (6 h) at room temperature. b Conversion and selectivity were determined by GC. The conversion and TON data of both the immobilized PS-TETA-Co catalyst and its homogeneous (CoL)Cl 4H2 O complex are given in two parallel rows, rst row for immobilized catalyst and the second one for homogeneous cobalt complex.

Conversion (%)

after completion of the reaction by simple ltration, washed with acetonitrile and dried at 100 C. The catalyst was recycled in order to test its activity as well as stability. The obtained results are presented in Fig. 10. As seen from Fig. 10, there was no appreciable change was observed in activity however a small decrease in conversion was observed which shows that the catalyst is stable and can be regenerated for repeated use.

100

95

3.12. Reaction mechanism and comparison with other reported systems It is proposed that the catalytic oxidation proceeds through hydroperoxocobalt(III) species. The FT-IR spectra (Fig. S12, supporting information) of the residue obtained after stirring polymer anchored complex and H2 O2 at 50 C for about 12 h showed a strong band at 3394 cm1 and a medium intensity band at 856 cm1 which might be attributed to O H and O O groups respectively present in Co(III)-peroxo intermediate [57,58]. The hydroperoxocobalt(III) species is obtained from the solvated intermediated which may be formed rst in the interaction of complex with H2 O2 . When the polymer anchored cobalt complex was stirred with H2 O2 and cyclohexene at 50 C, medium intensity band at 856 cm1 was disappeared and strong band at around 3394 cm1 is shifted and
Table 4 Oxidation of alkenes and suldes with H2 O2 catalyzed by a variety of catalysts. Substrates Cyclohexene Catalyst PS-TETA-Co (CoL)Cl 4H2 O CoMCM-41 Co/SBA-15 l-glu-Co/MCM-41 V-MCM-41 PS-TETA-Co (CoL)Cl 4H2 O [Mn(pydx-en)Cl(H2 O)]-Y [Mn(pydx-1,2-pn)Cl(H2 O)]-Y V-MCM-41 PS-TETA-Co (CoL)Cl 4H2 O [Mn(pydx-en)Cl(H2 O)]-Y [Mn(pydx-1,2-pn)Cl(H2 O)]-Y

90 Oxidation of methyl phenyl sulfide Oxidation of styrene 85 1 2 3 4 5 6 7

No of reaction cycle
Fig. 10. Recycling efciency for the oxidation of styrene and methyl phenyl sulde with PS-TETA-Co complex.

reduced in intensity. Similar mechanisms via the metal peroxo intermediates have been proposed in the literatures [5962]. The catalytic activity of the polymer anchored cobalt catalyst was also superior with other reported catalysts for the oxidation

Conversion (%) 88 73 41 54 31 84 97 81 76.9 72.3 83 92 78 87.5 84.9

Reference This work 63 64 65 This work 66 65 This work 66

Styrene

Methyl phenyl sulde

Sk.M. Islam et al. / Journal of Molecular Catalysis A: Chemical 380 (2013) 94103

103

reaction of alkenes and suldes with H2 O2 [6366]. The comparative results are given in Table 4. 4. Conclusions In summary, we have synthesized an octahedral Co(III) complex and characterized structurally and spectroscopically. The polymer anchored cobalt complex has proven to be successful in the oxidation of alkenes and suldes under mild reaction conditions. Alkenes and suldes oxidized to their corresponding allylic and sulfoxides products by both homogeneous and heterogeneous cobalt complexes. Comparison between catalytic activities of homogeneous cobalt complex with heterogeneous complex was done and it was shown that polymer anchored cobalt complex is more active than the homogeneous complex. The active sites do not leach out from the polymer support and thus polymer anchored cobalt catalyst can be reused without appreciable loss of activity, indicating that the anchoring procedure was effective. The reusability of this catalyst is high and can be reused seven times without signicant decrease from its initial activity. Acknowledgements We thank the Indian Association for the Cultivation of Science, Kolkata for providing the Single-Crystal X-ray instrument support. S.M.I. acknowledges Department of Science and Technology (DST) and Council of Scientic and Industrial Research (CSIR), New Delhi, India for funding. ASR acknowledges CSIR, New Delhi, for providing his senior research fellowship. P.M. acknowledges UGC, New Delhi, for providing her research fellowship. We acknowledge DST, New Delhi, India for providing support to the Department of Chemistry, University of Kalyani, under FIST and PURSE programme. Appendix A. Supplementary data Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.molcata. 2013.09.022. References
[1] K. Ghosh, V. Mohan, P. Kumar, U.P. Singh, Polyhedron 49 (2013) 167. [2] P.F. Rapheal, E. Manoj, M.R. Prathapachandra Kurup, E. Suresh, Polyhedron 26 (2007) 607. [3] E. Vinuelas-Zahnos, F. Luna-Giles, P. Torres-Garca, M.C. Fernndez-Caldern, Eur. J. Med. Chem. 46 (2011) 150. [4] M. Fleck, D. Karmakar, M. Ghosh, A. Ghosh, R. Saha, D. Bandyopadhyay, Polyhedron 34 (2012) 157. [5] L.I. Rodionova, A.V. Smirnov, N.E. Borisova, V.N. Khrustalev, A.A. Moiseeva, W. Grnert, Inorg. Chim. Acta 392 (2012) 221. [6] R. Chakrabarty, S.J. Bora, B.K. Das, Inorg. Chem. 46 (2007) 9450. [7] D.S. Nesterov, E.N. Chygorin, V.N. Kokozay, V.V. Bon, R. Bo ca, Y.N. Kozlov, L.S. Shulpina, J. Jezierska, A. Ozarowski, A.J.L. Pombeiro, G.B. Shulpin, Inorg. Chem. 51 (2012) 9110. [8] N. Gunasekaran, P. Jerome, S.W. Ng, E.R.T. Tiekink, R. Karvembu, J. Mol. Catal. A: Chem. 353 (2012) 156. [9] M.X. Li, C.L. Chen, D. Zhang, J.Y. Niu, B.S. Ji, Eur. J. Med. Chem. 45 (2010) 3169. [10] A. Panja, Polyhedron 43 (2012) 22. [11] S. Chattopadhyay, G. Bocelli, A. Musatti, A. Ghosh, Inorg. Chem. Commun. 9 (2006) 1053. [12] Z. Chen, M. Furutachi, Y. Kato, S. Matsunaga, M. Shibasaki, Angew. Chem. Int. Ed. 48 (2009) 2218. [13] E.L. Chang, C. Simmers, D.A. Knight, Pharmaceuticals 3 (2010) 17111728. [14] A. Bottcher, T. Takeuchi, K.I. Hardcastle, T.J. Meade, H.B. Gray, Inorg. Chem. 36 (1997) 2498. [15] B. Fredrich, W. Gerhartz (Eds.), Ullmanns Encyclopedia of Industrial Chemistry, A3, Weinheim, New York, 1985, p. 470.

[16] J. Drabowicz, P. Kielbsinski, M. Mikolajczyk, S. Patai, Z. Rappoport, C. Stirling (Eds.), The Chemistry of Sulphone and Sulphoxide, John Wiley and Sons, NY, 1988. [17] I. Frenandez, N. Khiar, Chem. Rev. 103 (2003) 3651. [18] V.B. Sharma, S.L. Jain, B. Sain, J. Mol. Catal. A: Chem. 227 (2005) 47. [19] T. Punniyamurthy, S. Velusamy, J. Iqbal, Chem. Rev. 105 (2005) 2329. [20] R. Chakrabarty, P. Sarmah, B. Saha, S. Chakravorty, B.K. Das, Inorg. Chem. 48 (2009) 6371. [21] I. Fernandez, J.R. Pedro, A.L. Rosello, R. Ruiz, I. Castro, X. Ottenwaelder, Y. Journaux, Eur. J. Org. Chem. (2001) 1235. [22] P. Sarmah, R. Chakrabarty, P. Phukan, B.K. Das, J. Mol. Catal. A: Chem. 268 (2007) 36. [23] G. Blay, L. Cardona, I. Fernandez, J.R. Pedro, Synthesis (2007) 3329. [24] J. Mondal, A. Modak, A. Dutta, S. Basu, S.N. Jha, D. Bhattacharyya, A. Bhaumik, Chem. Commun. 48 (2012) 8000. [25] M.R. Maurya, P. Saini, A. Kumar, J.C. Pessoa, Eur. J. Inorg. Chem. (2011) 4846. [26] B.S. Rana, B. Singh, R. Kumar, D. Verma, M.K. Bhunia, A. Bhaumik, A.K. Sinha, J. Mater. Chem. 20 (2010) 8575. [27] S. Parihar, S. Pathan, R.N. Jadeja, A. Patel, V.K. Gupta, Inorg. Chem. 51 (2012) 1152. [28] S. Bhunia, D. Saha, S. Koner, Langmuir 27 (2011) 15322. [29] P. Mukherjee, A. Bhaumik, R. Kumar, Ind. Eng. Chem. Res. 46 (2007) 8657. [30] S. Jana, B. Dutta, R. Bera, S. Koner, Inorg. Chem. 47 (2008) 5512. [31] B. Tamami, S. Ghasemi, Appl. Catal. A: Gen. 393 (2011) 242. [32] K.C. Gupta, H.K. Abdulkadir, S. Chand, J. Macromol. Sci. Pure Appl. Chem. 40 (2003) 475. [33] Q. Tang, Y. Wang, J. Liang, P. Wang, Q. Zhang, H. Wan, Chem. Commun. (2004) 440. [34] S.M. Islam, A.S. Roy, P. Mondal, K. Tuhina, M. Mobarak, J. Mondal, Tetrahedron Lett. 53 (2012) 127. [35] S.M. Islam, A.S. Roy, P. Mondal, N. Salam, J. Mol. Catal. A: Chem. 358 (2012) 38. [36] L. Li, Z. Liu, Q. Ling, X. Xing, J. Mol. Catal. A: Chem. 353 (2012) 178. [37] D.-H. Lee, J.-H. Kim, B.-H. Jun, H. Kang, J. Park, Y.-S. Lee, Org. Lett. 10 (2008) 1609. [38] M. Islam, P. Mondal, K. Tuhina, A.S. Roy, S. Mondal, D. Hossain, J. Organomet. Chem. 695 (2010) 2284. [39] S. Sharma, S. Sinha, S. Chand, Ind. Eng. Chem. Res. 51 (2012) 8806. [40] SAINT, version 6.02; SADABS, version 2.03, Bruker AXS, Inc., Madison, WI, 2002. [41] G.M. Sheldrick, SHELXS 97 Program for Structure Solution, University of Gttingen, Germany, 1997. [42] G.M. Sheldrick, SHELXL 97 Program for Crystal Structure Renement, University of Gttingen, Germany, 1997. [43] A.L. Spek, PLATON, Molecular Geometry Program, J. Appl. Crystallogr. 36 (2003) 7. [44] L.J. Farrugia, J. Appl. Crystallogr. 30 (1997) 565. [45] L.J. Farrugia, J. Appl. Crystallogr. 32 (1999) 837. [46] C. Floriani, M. Fiallo, A.C. Villa, C. Guastini, J. Chem. Soc. Dalton Trans. (1987) 1367. [47] L.-J. Liu, Synth. React. Inorg. Met. Org. Chem. 41 (2011) 531. [48] S. Sasi, M.R.P. Kurup, E. Suresh, J. Chem. Crystallogr. 37 (2007) 31. [49] S. Paul, W-T. Wong, D. Ray, Inorg. Chim. Acta 372 (2011) 160. [50] B.V. Patel, K. Desai, T. Thaker, Synth. React. Inorg. Met. Org. Chem. 19 (1989) 391. [51] V.B. Valodkar, G.L. Tembe, M. Ravindranathan, R.N. Ram, H.S. Rama, J. Mol. Catal. A: Chem. 208 (2004) 21. [52] K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination Compounds, 4th ed., Wiley, New York, 1986, pp. 313. [53] S.M. Islam, M. Mobarok, P. Mondal, A.S. Roy, N. Salam, D. Hossain, S. Mondal, Trans. Met. Chem. 37 (2012) 97. [54] A.B.P. Lever, Inorganic Electronic Spectroscopy, 2nd ed., Elsevier Science Publication, New York, 1986. [55] Z. Onal, H. Zengin, M. Sonmez, Turk. J. Chem. 35 (2011) 905. [56] S. Konar, A. Jana, K. Das, S. Ray, S. Chatterjee, J.A. Golen, A.L. Rheingold, S.K. Kar, Polyhedron 30 (2011) 2801. [57] W. Kemp, Organic Spectroscopy, Palgrave Publishers, New York, 2008. [58] L. Saussine, E. Brazi, A. Robine, H. Mimoun, J. Fischer, R. Weiss, J. Am. Chem. Soc. 107 (1985) 3534. [59] T. Pru, D.J. Macquarrie, J.H. Clark, App. Catal. A: Gen. 276 (2004) 29. [60] Y.H. Lin, I.D. Williams, P. Li, Appl. Catal. A: Gen. 150 (1997) 221. [61] R.A. Sheldon, J. Van Door, J. Catal. 31 (1973) 427. [62] N. Gunasekaran, P. Jerome, S.W. Ng, R.T. Edward, R. Tiekink, Karvembu, J. Mol. Catal. A: Chem. 353 (2012) 156. [63] N. Malumbazo, S.F. Mapolie, J. Mol. Catal. A: Chem. 312 (2009) 70. [64] R.-M. Wang, Z.-L. Zhan, J.-F. Lou, Y.-P. Wang, P.-F. Song, C.-G. Xia, Polym. Adv. Technol. 15 (2004) 48. [65] K.M. Parida, S. Singha, P.C. Sahoo, J. Mol. Catal. A: Chem. 325 (2010) 40. [66] M.R. Maurya, P. Saini, C. Haldar, F. Avecilla, Polyhedron 31 (2012) 710.

You might also like