Вы находитесь на странице: 1из 6

Applied Catalysis A: General 292 (2005) 223228 www.elsevier.

com/locate/apcata

Catalytic oxidation of cyclohexane to cyclohexanol and cyclohexanone over Co3O4 nanocrystals with molecular oxygen
Lipeng Zhou, Jie Xu *, Hong Miao, Feng Wang, Xiaoqiang Li
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Graduate School of the Chinese Academy of Sciences, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, PR China Received 7 April 2005; received in revised form 27 May 2005; accepted 11 June 2005 Available online 8 August 2005

Abstract Co3O4 nanocrystals with average particle sizes of 30 and 50 nm were synthesized using cobalt nitrate as precursor, and were characterized by X-ray diffraction (XRD), nitrogen adsorption, transmission electron microscopy (TEM), and Fourier transform infrared (FT-IR) spectroscopy. Catalytic oxidation of cyclohexane with molecular oxygen was studied over Co3O4 nanocrystals. These catalysts showed obviously higher activities as compared to Co3O4 prepared by the conventional methods, Co3O4/Al2O3, or homogeneous cobalt catalyst under comparable reaction conditions. The 89.1% selectivity to cyclohexanol and cyclohexanone at 7.6% conversion of cyclohexane was realized over 50 nm sized Co3O4 nanocrystals at 393 K for 6 h. # 2005 Elsevier B.V. All rights reserved.
Keywords: Nanocrystals; Co3O4; Cyclohexane; Oxidation

1. Introduction Oxidation of cyclohexane is one of the important bulk processes for the production of polyamide bers and plastics, such as nylon-6 and nylon-6,6. More than 106 tonnes of cyclohexanone and cyclohexanol (K/A oil) are produced worldwide per annum [1]. Industrially, the process for cyclohexane oxidation is carried out at 423433 K in the presence of cobalt-based homogeneous catalyst, resulting in about 4% conversion and 7085% selectivity to K/A oil [2]. This process is low in energy efciency, and generates plenty of by-products and waste. Increasing environmental concerns in recent years, as well as great demands for these products call for a more effective catalytic process using environmentally friendly oxidant, such as molecular oxygen [3,4]. Catalysts, such as TS-1, Ti-MCM-41, and metal containing aluminophosphate molecular sieve, have been employed for the cyclohexane oxidation, using molecular oxygen as oxidant [58]. MnAPO-36 was found to be the most active
* Corresponding author. Tel.: +86 411 84379245; fax: +86 411 84379245. E-mail address: xujie@dicp.ac.cn (J. Xu). 0926-860X/$ see front matter # 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.apcata.2005.06.018

one, and 13% conversion of cyclohexane and 62% selectivity to K/A oil were obtained at 403 K for 24 h under 1.5 MPa air pressure [9]. However, most of these systems show relatively low selectivity to the desired products, or exhibit low activity. As Schuchardt et al. have acknowledged, although many attempts have been made in the cyclohexane oxidation, it continues to be a challenge [10]. Nanostructured materials have attracted great interest in recent years due to their particular physical and chemical properties [11,12]. The properties of these materials mainly depend on their shape, size, and structure, which are strongly determined by the synthetic processes. Nanostructured catalysts, such as Fe2O3, Co3O4, and mixed FeCo oxide, have been employed in the cyclohexane oxidation [1315]. Unfortunately, these catalysts are unstable during the reaction. For example, employing Fe2O3 nanoparticles as catalyst in cyclohexane oxidation, the conversion in the second run was decreased to about 70% of that in the rst run [15]. The poor recyclability may attribute to the fact that the particles are amorphous and thus easy to leach into solution during reaction [15]. Although there are several reports on the preparation of cobalt oxide nanoparticles [1619],

224

L. Zhou et al. / Applied Catalysis A: General 292 (2005) 223228

synthesis of Co3O4 nanocrystals with a sharp distribution of particle size is still a difcult task, since the precipitation of cobalt in solution is very fast under basic conditions, and the particles agglomerate to form congeries of random shape in synthetic process. In this paper, we demonstrated a convenient method for synthesizing the stable Co3O4 nanocrystals from the commercially available inorganic cobalt salt. N-CetylN,N,N-trimethyl ammonium bromide (CTAB) and in situ generated triethylamine salt were applied as stabilizers to prevent the agglomeration of the particles during the preparation process, and Co3O4 nanocrystals with narrow size distribution were successfully obtained. The nanoparticles of Co3O4 could effectively catalyze the oxidation of cyclohexane to K/A oil under mild conditions with molecular oxygen as oxidant. The heterogeneity of the reaction, the stability and the recyclability of the catalyst make this system attractive for potential industry applications.

sample was designated as Catalyst 3. Co3O4 was also prepared by air-calcination of cobalt nitrate at 673 K for 2 h. Supported catalyst loading 5 wt% Co3O4 was prepared by impregnation of g-Al2O3 (BET surface area: 233 m2/g and pore volume: 0.389 cm3/g) with an aqueous solution of cobalt nitrate. After impregnation for 48 h, the materials were dried at 393 K overnight, and nally calcined at 673 K for 2 h (XRD analysis indicates cobalt oxide was in the cubic structure of Co3O4). 2.2. Characterization of catalysts X-ray diffraction (XRD) patterns were recorded on a PANalytical XPert PRO instrument with Cu Ka (0.154 nm) radiation at 40 kV and 40 mA in the range of 10708. A scan rate of 0.2088918/s with a step size of 0.0167118 was used for data collection. BET surface area of the samples was determined by nitrogen adsorption at 77 K using a Quantachrome NOWA 4000 apparatus. Fourier transform infrared (FT-IR) spectroscopy was performed on a Bruker Tensor 27 FT-IR spectrometer with 32 scans for a resolution of 4 cm1 in KBr media at room temperature. The microstructures of the catalysts were examined by transmission electron microscopy (TEM) on a JEOL JEM2000EX electron microscope at an accelerating voltage of 100 kV. For elemental analysis, the ICP analysis was carried out using a TJA Advantage ICP-AES instrument. 2.3. Catalytic tests The liquid-phase oxidation of cyclohexane was performed in a 100 mL autoclave reactor with a Teon insert inside. Typically, 15 g cyclohexane, 0.1 g (65 wt%) tert-butyl hydroperoxide (TBHP), and 0.2 g catalyst were added into the reactor. Then, the reactor was heated to the reaction temperature after sealing, while agitation was realized by means of a magnetic stirrer. Upon heating to the reaction temperature, the reactor was charged with 1.0 MPa of O2. During reaction, oxygen was fed continuously to maintain constant pressure at 1.0 MPa. When the reaction was stopped, the reaction mixture was diluted with 15 g ethanol to dissolve the by-products. The reaction products were identied by comparison with authentic products and by GCMS analysis. After the quantitative decomposition of cyclohexyl hydroperoxide (CHHP) to cyclohexanol by adding excess triphenylphosphine to the reaction mixture, the quantitative analyses of cyclohexanol and cyclohexanone were carried out by an Agilent 4890D gas chromatography, which was equipped with an OV-1701 column (30 m 0.25 mm 0.3 mm), using toluene as the internal standard. The concentrations of CHHP were determined by iodometric titration, and the acid and ester by acidbase titration. Typically, CHHP was analyzed by following procedure:

2. Experimental 2.1. Catalyst preparation All chemicals were of analytic grade without any purication. Co(NO3)26H2O (99%), N-cetyl-N,N,N-trimethyl ammonium bromide (99%), Na2CO3 (99%), and (CH3CH2)3N (TEA, 99%) were purchased from Tianjin Kermel Chemical Reagent Development Center. Co3O4 nanocrystals were prepared by precipitation method using cobalt nitrate as the precursor in the presence of CTAB. In a typical synthesis, 124.5 mmol of Co(NO3)26H2O and 25.0 mmol of CTAB were dissolved in 750 mL deionized water, in a 1000 mL round bottom ask, giving a pink solution. After stirring for 20 min, 249.0 mmol of triethylamine was rapidly added into the pink solution under vigorous agitation, giving a blue mixture. Then, the blue mixture was stirred at 348 K for 8 h, and the color of the mixture changed into black. The as-synthesized cobalt oxide was obtained by centrifugation and then dried at 353 K for 48 h. The dried material was treated by Soxhlet extraction over ethanol for 48 h to remove stabilizers. The obtained solid referred to as Catalyst 1 was calcined at 573 K under owing air for 2 h after drying at 373 K for 8 h. Catalyst 2 was prepared by the following procedure: 12.5 mmol of Co(NO3)26H2O and 2.5 mmol of CTAB were used with the same molar ratio of Co/TEA, while the other experimental conditions were same with Catalyst 1. For comparison, Co3O4 was prepared by conventional method adapted from literature [17]. A solution of Na2CO3 was added into a solution of cobalt nitrate under agitation, and then the mixture was stirred for 8 h at 348 K. The precipitate was ltrated and washed with deionized water for three times, and then the solid was calcined at 573 K under owing air for 2 h after it was dried at 373 K for 8 h. This

L. Zhou et al. / Applied Catalysis A: General 292 (2005) 223228

225

Fig. 1. XRD patterns of Catalysts 13.

20 mL acetic acid was added to 510 g reaction mixture, 10 mL saturated KI solution was added, and the mixture was stored for 20 min in darkness and then titrated with 0.1 M Na2S2O3 solution. Acids and esters were analyzed as follows: 0.2 g triphenylphosphine was added to 510 g reaction mixture to decompose CHHP, the solution was kept for 30 min before diluting with 25 mL acetone, and 25 mL deionized water and then titrated with 0.1 M NaOH. The solution was reuxed for 1 h after adding 20 mL 0.1 M NaOH and then titrated with 0.1 M HCl.

Fig. 3. FT-IR spectra of Catalyst 1: (a) after calcinations, (b) as-synthesized after, and (c) before extraction of stabilizers.

3. Results and discussion 3.1. Preparation and characterization Cobalt oxide nanoparticles was prepared from cobalt nitrate in the presence of CTAB and in situ generated triethylamine salt. The BET surface area of Catalysts 13 was 21, 57, and 79 m2/g, respectively. Fig. 1 summarizes the XRD patterns of the calcined catalysts prepared by the present method and its comparison with that prepared by the conventional method. Within these peaks, eight peaks at 36.88, 65.38, 59.48, 31.38, 44.88, 18.98, 38.58, and 55.78 in

decreasing order of peak height of the three samples can be assigned to the cubic structure of Co3O4 with Fd3m space group as compared to a data le JCPDS-43-1003, which gives eight peaks at 36.88, 65.28, 59.38, 31.38, 44.88, 19.08, 38.58, and 55.78 in decreasing order of peak height. The value of the cell constant (a = 0.8082) is very near to the published data for Co3O4 powder (a = 0.8084). This indicates that these catalysts had highly crystallized cubic structure. The particle sizes of Catalysts 13 calculated from the (3 1 1) plane using the Sherrer equation were 33, 24, and 19 nm, respectively. The microstructure of Co3O4 nanocrystals prepared by present method is given in Fig. 2. TEM analysis had revealed evidence that the particles exhibited platelet morphologies with minor particle agglomeration and a relatively narrow particle size distribution. The average particle diameters of Catalysts 1 (Fig. 1a) and 2 (Fig. 1b) were 50 and 30 nm, respectively. This result was coincident with the particle sizes calculated from the Sherrer equation. It implies that the secondary particles were not formed and the samples were highly crystalline. Fig. 3 shows the FT-IR spectra of Catalyst 1 after calcination (trace a), and those of as-synthesized Catalyst 1

Fig. 2. TEM photographs of (a) Catalyst 1 and (b) Catalyst 2.

226

L. Zhou et al. / Applied Catalysis A: General 292 (2005) 223228 Table 1 Oxidation of cyclohexane with different catalystsa Entry Catalyst Conversion (mol%) 1.2 7.6 7.3 1.3 2.8 0.2 4.7 3.5 2.3 1.6 5.8 2.7 9.6 Product distribution (mol%) A 7.3 39.6 33.9 60.4 51.1 79.8 32.6 38.3 41.0 35.0 36.5 36.9 26.7 K 35.9 49.5 41.7 33.7 33.3 20.2 28.7 31.4 35.7 21.7 41.2 56.0 59.8 CHHP 53.6 0.8 1.1 2.9 9.8 0.8 10.7 12.4 44.1 14.2 0 0 Acid 2.3 4.8 14.5 3.0 3.3 22.5 14.6 5.4 4.4 5.9 9.0 Ester 0.9 5.3 8.8 0 2.5 15.4 5.0 5.5 3.5 1.2 4.6

before (trace c) and after (trace b) extraction of the stabilizers. The bands at 2915 and 2848 cm1 in trace c were the CH stretching modes attributed to CTAB and triethylamine salt. After extraction of the stabilizers by ethanol, the peaks attribute to CH stretches almost disappeared, which suggested that CTAB and triethylamine salt were mostly removed from samples. Two bands at 668 and 579 cm1 could be assigned to the vibration of CoO due to Co3O4 [20], as indicated that Co3O4 were formed in the aging process. After addition of triethylamine, at rst the color of the mixture was blue and then it slowly turned into black, showing the color of Co3O4, in the aging process. The color change provided another evidence that Co3O4 nanoparticles were formed during the aging. CTAB and triethylamine salt stabilized the particles by capping the surface of nanoparticles with organic ligands. The formation of nanoparticles was proposed through inorganicorganic layered particles model. Because the surface of metal oxide particles was polar, and the organic cations of in situ generated triethylamine salt and CTAB were adsorbed on the surface of metal oxide particles. The organic shells can keep the nanoparticles apart from each other by presenting a strong steric hindrance to overcome van der Waals interactions between the nanoparticles [21]. It promoted the formation of nanoparticles with narrow size and shape distribution [2123]. Analogous adsorptions of surfactants on the surface of mica, silica, and metal oxide have also been reported [24,25]. 3.2. Catalytic performance Metal oxide nanoparticles can exhibit unique chemical properties due to their limited size and a high density of corner or edge surface sites [26]. Herein, the activity of the Co3O4 nanocrystals catalysts is compared with the conventionally prepared catalyst and homogeneous cobalt catalyst for the liquid phase oxidation of cyclohexane using molecular oxygen as oxidant. Since the reaction had an induction period, a small amount of TBHP that acted as a radical initiator was added to the reaction system to eliminate the induction period [27]. The products obtained were cyclohexanol, cyclohexanone, CHHP, acid (mainly adipic acid), and ester (dicyclohexyl adipate, hexanolactone, and the other esters). CHHP could decompose totally in the chromatograph to produce mainly cyclohexanol and cyclohexanone, so it could not be detected by GC [28,29]. The acid, such as adipic acid, would easily decompose in the GC analyses, hence it could not be exactly determined by GC. The components of esters were quite complicated and some of them also could not be detected by GC [30]. In this study, quantitative analyses of CHHP, acids, and esters were carried out by titration, and cyclohexanol and cyclohexanone by GC. The analytical results of a solution of cyclohexane oxidation from an industrial operation by our method were in accordance with the separated yields, which proved that our analytical method for

1 2 3 4 5 6 7 8 9 10 11 12 13
a

Blank 1 2 3 1b 1c Cod Co3O4 e Co/Al2O3 CoAPO-5f CoAPO-5g MnAPO-5 f MnAPO-5 g

Reaction time: 6 h and reaction temperature: 393 K. A: cyclohexanol, K: cyclohexanone, acid: mainly adipic acid, ester: dicyclohexyl adipate, hexanolactone, and other ester. b Reaction without TBHP. c Add 5 wt% 2,6-di-tert-butyl-4-methylphenol (BHT). CHHP, acid, and ester were not analyzed. d Co(OAc)24H2O as catalyst, [Co] = 100 ppm. e Co3O4 was prepared by air-calcination cobalt nitrate at 673 K for 2 h. XRD analysis indicated Co3O4 had highly crystallized cubic structure. f Results from Ref. [27]. Reactions were carried at 403 K for 8 h. g Results from Ref. [27]. Reactions were carried using a free-radical initiator TBHP (0.7 wt% of cyclohexane) at 403 K for 8 h.

the reaction mixture of cyclohexane oxidation was believable. 3.2.1. Oxidation of cyclohexane over different catalysts As shown in Table 1, the conversions of cyclohexane over Catalyst 3 and Co3O4 prepared by air-calcination of cobalt nitrate, which were prepared by the conventional method, were 1.3% and 3.5%, respectively. Co3O4/Al2O3 showed a cyclohexane conversion of 2.3%. Further, the reaction was carried out in the absence of catalyst (entry 1) and a low conversion was noted. On the other hand, Co3O4 nanocrystals showed obviously higher conversion as compared to Co3O4 prepared by the conventional method or Co3O4/ Al2O3 under same reaction conditions. For example, Catalyst 1 exhibited the best performance, with a cyclohexane conversion of 7.6%, 89.1% selectivity, and a yield of 6.7% to K/A oil, which was higher than those obtained using metal substituted AlPO-n, or MCM-41 as catalyst [68,27]. Addition of TBHP, which acts as a free-radical initiator, decreased the induction period and the conversion was increased (entries 2 and 5). Similar phenomenon was also observed in cyclohexane oxidation with MeAlPO-5, such as CoAPO-5 and MnAPO-5 (entries 1013), as catalysts. It should be noted that the homogeneous cobalt catalyst showed a cyclohexane conversion of only 4.7% and a selectivity of 61.3% to K/A oil under comparable reaction conditions (entry 7). The conversion over Catalysts 1 and 2 were comparable, which suggests that particle sizes range from 30 to 50 nm had slight effect on the conversion of

L. Zhou et al. / Applied Catalysis A: General 292 (2005) 223228

227

Fig. 4. Inuence of reaction temperature on Catalyst 1 performance.

Fig. 5. Inuence of reaction time on cyclohexane oxidation over Catalyst 1.

cyclohexane. However, the selectivity to K/A oil on the Catalyst 1 was higher than that of Catalyst 2. One of the possible reasons may be attributed to that smaller particles have stronger surface energy, which will lead to the overreaction of cyclohexanol and cyclohexanone to by-products. As Catalyst 1 was the most active catalyst in this work, we have chosen this catalyst for the more detailed study. 3.2.2. Effect of reaction temperature Fig. 4 shows the effect of the reaction temperature on cyclohexane oxidation at reaction time of 6 h. As expected, the conversion of cyclohexane increased with reaction temperature. On the other hand, at initial stage, the selectivity to K/A oil went up with increasing temperature and reached a maximum of 89.1% at 393 K, and then decreased remarkably with further increasing of temperature. The distribution of cyclohexanone also reached its maximum 49.5% at 393 K and the change trend was similar with that of selectivity to K/A oil. However, as reaction temperature increases, the distribution of cyclohexanol decreased all along. At rst the change was very small, and then cyclohexanol distribution decreased sharply. In addition, the molar ratio of K/A oil went up when the temperature was increased, which suggests that cyclohexanol was continuously oxidized to cyclohexanone [1]. The change trend of CHHP distribution was like that of cyclohexanol, and there was very small quantity of CHHP in the reaction mixture when the reaction temperature was above 393 K. For example, the CHHP distribution was only 0.8% at 393 K. The selectivity to K/A oil increased while the distribution of CHHP decreased at high reaction temperature, and this show that CHHP is an unstable intermediate and it could decompose to cyclohexanol, cyclohexanone [1]. Generally, the acid and ester were the products of deep oxidation of K/A oil. The distributions of these by-products increased with increasing of the reaction temperature, which indicates that, at higher temperature, the rate of oxidizing the substrate to desired products was slower than that of oxidizing K/A oil to form by-products. Thus, reaction

temperature is an important reaction parameter to maintain a high selectivity to the desired products. For this reason, 393 K was the most proper reaction temperature and thus selected as reaction temperature for the subsequent studies. 3.2.3. Effect of reaction time The effect of reaction time on cyclohexane conversion and products distributions at reaction temperature of 393 K is shown in Fig. 5. It can be seen the cyclohexane conversion steadily increased with increasing reaction time. In addition, the selectivity to K/A oil rose to its maximum 90.4% at 4 h, and decreased with further increase reaction time. At initial 4 h, the distribution of alcohol changed slightly, but it decreased with further increasing reaction time. On the other hand, the selectivity of cyclohexanone increased during the initial 8 h and reached a maximum of 55.4% at reaction time of 8 h, and then decreased with further increasing reaction time. The distribution of CHHP decreased sharply during the initial 6 h period, and thereafter the amount of CHHP became very little (<1%). The distribution of acid and ester increased with increasing reaction time. However, the highest yield of K/A oil was obtained at the reaction time of 6 h. We chose 6 h for the subsequent studies.

Fig. 6. Catalytic performance of the fresh and recycled Catalyst 1.

228

L. Zhou et al. / Applied Catalysis A: General 292 (2005) 223228

3.2.4. The stability of catalyst In order to check the recyclability of the catalyst, six reaction runs were carried out. After each reaction the catalyst was separated by ltration and washed with water and ethanol for several times, then the catalyst was dried at 393 K for 8 h to remove water and ethanol. The regenerated catalyst was used for the recycling study under the same reaction conditions. The results on the fresh and recycled catalyst are shown in Fig. 6. It can be seen that the conversion of cyclohexane on the recycled catalyst was slightly lower than that on fresh catalyst. The selectivity of K/A oil on recycled catalyst was almost equal with that over fresh catalyst. These results disclose that the catalyst is stable or could be recycled after the reaction. Analyzing the reaction mixture with ICP mass spectrometry, only a small amount of cobalt (0.1 ppm) was observed, which could be attributed to leaching of cobalt ions or the nanoparticles not removable by ltration. To study whether the catalytic reaction was via heterogeneous catalysis, the solution of reaction mixtures was ltered after reaction for 4 h, and then was reacted at 393 K for another 2 h under stirring. It was found that the conversion of the ltrate changed from 4.5% to 5.3%, which would attribute to the autoxidation or the cobalt in the ltrate. These evidences proved that the reaction was a heterogeneous process, because the amount of dissolved cobalt is much lower than in the homogeneous catalysis and the contribution from homogeneous cobalt is neglectable [7]. The oxidation of cyclohexane is believed to be a freeradical mechanism. It is evidenced by the following facts: the reaction was terminated by the addition of 5 wt% 2,6-di-tertbutyl-4-methylphenol, which was a free-radical scavenger (see Table 1, entry 6). Moreover, addition of a free radical (TBHP) decreased the induction period and increased the rate of the reaction, while the selectivity remained unchanged (see Table 1, entries 2 and 5). Another fact is that the CHHP existing in the initial stage of the reaction could decompose to cyclohexanol and cyclohexanone (see Fig. 5).

of the fresh catalyst. The oxidation of cyclohexane is considered to undergo via a free-radical mechanism.

Acknowledgments We gratefully acknowledge the National Natural Science Foundation of China (Grant: 20233040) and the National High Technology Research and Development Program of China (Grant: 2004AA32G020) for nancial supports. We also thank Prof. Ryuichiro Ohnishi for helpful discussions on preparation the paper.

References
[1] A.K. Suresh, M.M. Sharma, T. Sridhar, Ind. Eng. Chem. Res. 39 (2000) 3958. [2] K. Weissermel, H.-J. Arpe, Industrial Organic Chemistry, second ed., VCH Press, Weinheim, 1993. [3] U.R. Pillai, E. Sahle-Demessie, Chem. Commun. (2002) 2142. [4] S.-I. Murahashi, Y. Oda, T. Naota, J. Am. Chem. Soc. 114 (1992) 7913. [5] G. Lu, H. Gao, J. Suo, S. Li, J. Chem. Soc., Chem. Commun. (1994) 2423. [6] M.L. Correia, M. Wallau, U. Schuchardt, Quim. Nova 19 (1996) 43. [7] D.L. Vanoppen, D.E. De Vos, M.J. Genet, P.G. Rouxhet, P.A. Jacobs, Angew. Chem., Int. Ed. 34 (1995) 561. [8] L. Zhou, J. Xu, H. Miao, X. Li, F. Wang, Catal. Lett. 99 (2005) 231. [9] J.M. Thomas, R. Raja, G. Sankar, R.G. Bell, Stud. Surf. Sci. Catal. 130 (2000) 887. [10] U. Schuchardt, D. Cardoso, R. Sercheli, R. Pereira, R.S. da Cruz, M.C. Guerreiro, D. Mandelli, E.V. Spinace, E.L. Pires, Appl. Catal. A 211 (2001) 1. [11] A. Cho, Science 229 (2003) 1684. [12] J.M. Thomas, B.F.G. Johnson, R. Raja, G. Sankar, P.A. Midgley, Acc. Chem. Res. 36 (2003) 20. [13] V. Kesavan, P.S. Sivanand, S. Chandersekaran, Y. Koltypin, A. Gedanken, Angew. Chem., Int. Ed. 38 (1999) 3521. [14] N. Perkas, Y. Koltypin, O. Palchik, A. Gedanken, S. Chandrasekaran, Appl. Catal. A 209 (2001) 125. [15] V. Kesavan, D. Dhar, Y. Koltypin, N. Perkas, O. Palchik, A. Gedanken, S. Chandersekaran, Pure Appl. Chem. 73 (2001) 85. [16] Y. Jiang, Y. Wu, B. Xie, Y. Qian, Mater. Chem. Phys. 74 (2002) 234. [17] H. Yang, Y. Hu, X. Zhang, G. Qiu, Mater. Lett. 58 (2004) 387. [18] J. Feng, H.C. Zeng, Chem. Mater. 15 (2003) 2829. [19] F. Svegl, B. Orel, I. Grabec-Svegl, V. Kaucic, Electrochim. Acta 45 (2000) 4359. [20] B. Pejova, A. Isahi, M. Najdoski, I. Grozdanov, Mater. Res. Bull. 36 (2001) 161. [21] B.L. Cushing, V.L. Kolesnichenko, C.J. OConnor, Chem. Rev. 104 (2004) 3893. [22] D.-K. Kim, M. Mikhailova, Y. Zhang, M. Muhammed, Chem. Mater. 15 (2003) 1617. [23] W.S. Seo, H.H. Jo, K.M. Lee, J.T. Park, Adv. Mater. 15 (2003) 795. [24] J. Stiernstedt, J.C. Froberg, F. Tiberg, M.W. Rutland, Langmuir 21 (2005) 1875. [25] R.M. Pashley, B.W. Ninham, J. Phys. Chem. 91 (1987) 2902. [26] M. Fernandez-Garca, A. Martnez-Arias, J.C. Hanson, J.A. Rodriguez, Chem. Rev. 104 (2004) 4063. [27] R. Raja, G. Sankar, J.M. Thomas, J. Am. Chem. Soc. 122 (1999) 11926. [28] G.B. Shulpin, J. Mol. Catal. A 189 (2002) 39. [29] J.D. Chen, R.A. Sheldon, J. Catal. 153 (1995) 1. [30] D.L. Vanoppen, D.E. De Vos, M.J. Genet, P.G. Rouxhet, P.A. Jacobs, Angew. Chem., Int. Ed. Engl. 34 (1995) 560.

4. Conclusions A new method for preparing Co3O4 nanocrystals from inorganic cobalt salt is established. Co3O4 nanocrystals with average particle size of 30 and 50 nm were successfully synthesized using cobalt nitrate as precursor. CTAB and in situ generated triethylamine salt acted as stabilizers to prevent the nanocrystals agglomerate in the process. Co3O4 nanocrystals are effective catalysts in cyclohexane oxidation to cyclohexanol and cyclohexanone using molecular oxygen as oxidant. Co3O4 with an average particles size of 50 nm showed a better activity and selectivity to the desired products, and the conversion of cyclohexane was 7.6% while the selectivity to K/A oil was 89.1% for a reaction time of 6 h. The catalyst was recyclable and was reused ve times with only a slight lost of the activity, and exhibited practically the same selectivity as that

Вам также может понравиться