Вы находитесь на странице: 1из 9

Applied Catalysis B: Environmental 101 (2010) 4553

Contents lists available at ScienceDirect

Applied Catalysis B: Environmental


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

Synthesis and characterization of cobalt-substituted SBA-15 and its high activity in epoxidation of styrene with molecular oxygen
Haitao Cui a,b , Ye Zhang a, , Zegang Qiu a , Liangfu Zhao a, , Yulei Zhu a
a b

State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, PR China Graduate University of Chinese Academy of Sciences, Beijing 100049, PR China

a r t i c l e

i n f o

a b s t r a c t
Cobalt-substituted SBA-15 (Co-SBA-15x ) was successfully synthesized by a simple and effective method designated as pH-adjusting. Based on series of characterizations, the material still retained a highly ordered mesostructure of SBA-15 when a certain amount of cobalt was introduced. The cobalt species in Co-SBA-15x mainly existed in the single-site Co(II) state. Most of them were incorporated into the silica framework of SBA-15, locating at tetrahedrally coordinated sites. For epoxidation of styrene with O2 , CoSBA-15x exhibited the highest catalytic activity among the cobalt-based catalysts prepared by different method with the same cobalt content, and the single-site Co(II) in Co-SBA-15x was the most active in the reaction. The inuences of reaction conditions such as reaction temperature, catalyst amount, solvent, solvent amount, reaction time, oxidant, and oxidant amount on the performance of Co-SBA-15x were also investigated in detail. Co-SBA-15x possessed excellent stability and recyclability in the reaction. 2010 Elsevier B.V. All rights reserved.

Article history: Received 15 May 2010 Received in revised form 1 September 2010 Accepted 3 September 2010 Available online 21 September 2010 Keywords: Cobalt SBA-15 Epoxidation Styrene Molecular oxygen

1. Introduction The epoxidation of styrene has attained great attention recently because the epoxide is a key intermediate in the production of many ne chemicals and pharmaceuticals. In conventional process, the epoxide is produced by the reaction of styrene with peracid [1]. However, a number of problems from this method hinder its further application, such as formation of undesirable wastes, high cost and corrosivity of the peracid. Some environmentally benign methods have been explored in the past few years, and many attractive catalytic systems have been reported for the reaction. For instance, the epoxidation of olens with H2 O2 over TS-1 [2] or using O2 combined with a sacricial reductant (H2 , aldehyde or alcohol) catalyzed by noble-metal catalyst [36] is thought to be a green process. Nevertheless, from both environmental and economic viewpoints, the epoxidation of styrene with O2 or air without any co-reductant over effective catalysts is an attractive and promising method. Cobalt-based catalysts are widely used in a variety of catalytic reactions, such as reforming of methane [7] or ethanol [8], ethene hydroformylation [9], Fischer-Tropsch synthesis [1012], hydrogenation of aromatics [13] or aldehydes [14] and ethyl acetate total oxidation [15]. Likewise, several studies on the epoxidation of

Corresponding authors at: Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, PR China. Tel.: +86 351 4041526; fax: +86 351 4041526. E-mail addresses: yzhang@sxicc.ac.cn (Y. Zhang), lfzhao@sxicc.ac.cn (L. Zhao). 0926-3373/$ see front matter 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.apcatb.2010.09.003

alkenes with molecular oxygen or air over cobalt-based catalysts have been carried out recently. Ion-exchanged Co-X and Co-MCM41 [16], Co-ZSM-5 [17], CoOx and CoOx /SiO2 [18], etc. all exhibited better catalytic performance in the reaction. Therefore, cobaltbased catalysts have aroused great interest of the researchers for the reaction. Mesoporous silicas have attracted increasing attentions due to the large surface area and uniform mesoporous channels since the researchers at Mobil Corporation reported M41S in the early 1990s [19]. SBA-15, another type of mesoporous silica, shows broader application in catalysis because of its larger pores, thicker walls and higher hydrothermal stability compared with M41S [20]. Unfortunately, SBA-15, as a pure silica, is nearly inert for chemical reactions. The active sites in the molecular sieves are often from heteroatoms. However, it is very difcult to prepare SBA-15 containing heteroatoms in the framework because of the strong acidic synthesis condition (pH < 0). Metals always exist in cationic form rather than oxo species under acidic condition and they therefore cannot be introduced into the framework of SBA-15. Much efforts have been devoted to the effective introduction of heteroatoms. According to reported studies, postsynthesis grafting [2128] and direct-synthesis [2933] are typical processes for heteroatoms introduction. Postsynthesis grafting can introduce high-content heteroatoms into the mesoporous silica. But one drawback is that the experiment procedure is relatively complicated, and another is that metal oxides tend to appear on the external surface or in the pores, resulting in the ordering destruction of mesostructure. Direct-synthesis is a relatively simple way. However, most of the

46

H. Cui et al. / Applied Catalysis B: Environmental 101 (2010) 4553

metal ions in the initial gels cannot be effectively introduced into the mesoporous molecular sieves by this route. Wu et al. [34] reported an effective and convenient method denoted as pH-adjusting for the grafting of heteroatoms. In this way, Al and Ti could be largely introduced into the mesophase. However, the incorporation of Co is thought to be a big challenge in the eld of mesoporous material synthesis [35]. There are not many studies on synthesis of Co-SBA-15. Lou et al. synthesized a highly ordered Co-SBA-15 at pH > 2.0 using sodium hydroxide to adjust the pH of synthesis gels, and mainly discussed the effect of pH on the structure in their work [36]. In the present study, we also successfully introduced a certain amount of cobalt into the mesoporous silica SBA-15 by so-called pH-adjusting method. Ammonia was used to adjust the pH of synthesis system. The prepared Co-SBA15 was rstly used for the epoxidation of styrene with O2 and the results showed high activity due to the effective introduction of cobalt species and its suitable state of existing. 2. Experimental 2.1. Materials All chemicals used were of reagent grade, including triblock copolymer poly(ethylene oxide)poly(propylene oxide) poly(ethylene oxide) EO20 PO70 EO20 (P123, Aldrich), HCl solution (2 M), tetraethyl orthosilicate (TEOS), deionized water, cobalt(II) nitrate hexahydrate (99%, Co(NO3 )2 6H2 O), nickel(II) nitrate hexahydrate (98%, Ni(NO3 )2 6H2 O), ferric nitrate nonahydrate (98.5%, Fe(NO3 )3 9H2 O), anhydrous manganese chloride (99.5%, MnCl2 ), tin chloride dihydrate(99%, SnCl2 2H2 O), titanium chloride (99%, TiCl4 ), hydrogen peroxide (30%, H2 O2 ), tert-butyl hydroperoxide solution (65%, TBHP), N,N -dimethylformamide (99.5%, DMF), styrene (99%), bromobenzene (99%), absolute ethanol (99.7%), N,N-dimethylacetamide (99%, DMA), dimethyl sulfoxide (99%, DMSO), toluene (99.5%), pyridine (99.5%), and cyclohexanone (99.5%). 2.2. Preparation of catalysts Cobalt-substituted SBA-15 was synthesized by the pH-adjusting method according to reference [34]. In our typical run, 2 g P123 triblock copolymer surfactant was dissolved in 15 g water and 60 g 2 M HCl solution, followed by addition of 4.25 g TEOS to the homogeneous solution under stirring. The mixture was stirred at 313 K for 4 h, and then an appropriate amount of Co(NO3 )2 6H2 O was added to the mixture, followed by additional stirring at 313 K for 20 h. The mixture was then transferred into an autoclave for crystallizing at 373 K for 2 days. After the procedure above, the pH value of the system was adjusted to 7.5 by adding ammonia dropwise at room temperature and the mixture was crystallized again at 373 K for another 2 days. The resultant mixture was ltered, and the obtained solid was thoroughly washed with deionized water, and dried at room temperature. Then its surfactants were removed by Soxhlet extraction. The obtained product is denoted as Co-SBA-15x , where x is the Si/Co ratio in the initial gel. For comparison, series of the catalysts such as Fe-SBA-1520 , Ni-SBA-1520 , Sn-SBA-1520 , Ti-SBA-1520 and Mn-SBA-1520 were synthesized through the same procedure as shown above except that Co(NO3 )2 6H2 O was replaced by other corresponding heteroatom sources. In addition, cobalt-containing MCM-41 was synthesized by the template-ion exchange method (Co-MCM-41TIE) according to Tang et al. [16]. SBA-15 or silica gel (SG) was used as support to prepare Co/SBA-15 and Co/SG by the conventional impregnation method. Typically, a certain amount of SBA-15 or silica gel (surface area of 220 m2 /g) powder was immersed in the aqueous solution containing an appropriate amount of cobalt(II)

nitrate for 8 h at ambient temperature under stirring, and then the excess solvent was evaporated off at 323 K. After drying at 373 K and calcinating at 823 K for 6 h, the resultant sample is designated as Co/SBA-15 or Co/SG. Moreover, in order to investigate the inuence of preparation steps on the epoxidation of styrene with O2 , four cobalt-based catalysts were prepared by adding the same amount of identical materials in the initial gels but the preparation steps were different. (1) Co-SBA-1520 was prepared by the pH-adjusting method. (2) The preparation steps were the same as those of Co-SBA-1520 at rst. But after the rst crystallization procedure, the steps are different. The excess solvent of the produced mixture was evaporated off at 323 K. The product was dried at 373 K and calcinated at 823 K for 6 h. The obtained powder is designated as Co-SBA-1520 -1. (3) The sample was prepared without adjusting pH of synthesis system. The preparation procedures were also the same as those of Co-SBA1520 at rst. But after the rst crystallization step, the resultant product was ltered. The obtained solid was thoroughly washed with deionized water, followed by extracting the template with Soxhlets extracter, and drying at room temperature. The resultant powder is denoted as Co-SBA-1520 -2. (4) Co-SBA-1520 -3 was prepared by the same steps as Co-SBA-1520 except that the surfactants of the former were not removed by Soxhlets extracter. 2.3. Characterizations The textural properties of the samples were derived from N2 adsorption/desorption measurement at 77 K on Micromeritics Tristar 3000. In each case, the sample was outgassed under vacuum at 573 K for 3 h before N2 adsorption. Pore size distributions for mesoporous materials were calculated from the desorption isotherms by the BJH method. Cobalt content in each sample was analyzed with inductively coupled plasma-atom emission spectroscopy (ICP-AES) on TJA Atomscan 16 spectrometer. X-ray photoelectron spectroscopy (XPS) was measured with a PHI Quantum 2000 Scanning ESCA Microprobe and monochromatic Al-K radiation. The binding energies were calibrated by referencing the C 1s at 284.6 eV. High resolution transmission electron microscopy (HRTEM) measurements were carried out on JEOL JEM-2010 operating at 200 kV. X-ray diffraction (XRD) analysis was performed on RIGAKU D/maxrB X-ray diffractometer. Diffraction patterns were recorded with Cu K radiation (40 mA, 40 kV). Temperature-programmed reduction (TPR) with H2 was recorded by a ow system with a TCD detector. Generally, 0.1 g sample was rstly pretreated in a quartz reactor with a owing argon at 823 K for 1 h. After cooling to about 313 K, a H2 Ar (5% H2 ) mixture was introduced into the reactor, and the temperature was raised to 1200 K at a rate of 10 K min1 . Raman spectroscopic measurements were carried out on J.Y. LabRAM laser micro-Raman spectroscopy. A laser at 514.5 nm was used as the excitation source. UVvis spectra (UVvis) of the sample was recorded on a Shimadzu UV-2550 spectrometer. The powdered sample was loaded into a quartz cell, and the spectra were collected in the range of 200800 nm referenced to BaSO4 . 2.4. Catalytic reactions The epoxidation of styrene with O2 was performed in a 250-ml three-necked at-bottomed glass ask equipped with a liquid condenser. Typically, 8 mmol of styrene, a measured amount of solvent and a certain amount of catalyst were added to the reactor. The mixture was stirred vigorously by a magnetic stirrer and heated to the desired temperature. Then O2 at a certain stable ow rate controlled by a mass ow controller was introduced into the liquid by bubbling at atmospheric pressure. The reactant mixture was stirred vigorously for a specied period of time. After completion of the reaction, the liquid products were obtained by centrifugation and they were

H. Cui et al. / Applied Catalysis B: Environmental 101 (2010) 4553

47

quantitatively analyzed by a gas chromatograph equipped with a capillary column (DB-5, 30 m 0.25 mm 0.25 m) and a FID detector, with bromobenzene as an internal standard. In most cases, the epoxide was obtained as the main product, and benzaldehyde was main by-product. Phenylacetaldehyde selectivity was less than 10%. Other products included benzoic acid, phenylacetic acid, etc. Under the most reaction conditions, the polymerization of styrene was negligible, so polymers in the reaction products were not considered. The conversion of styrene and selectivity of the products were calculated as follows: Styrene conversion (%) = (initial moles nal moles) 100% (initial moles) (moles of the product) 100% (moles of all products)

Table 1 Textural parameters of SBA-15 and Co-SBA-15x . Materials SBA-15 Co-SBA-15100 Co-SBA-1540 Co-SBA-1520 Co-SBA-1510 Surface area (m2 /g) 537 335 305 298 292 Pore volume (cm3 /g) 1.18 1.08 1.05 1.03 1.01 Pore diameter (nm) 8.82 12.9 13.8 13.8 13.8

The product selectivity (%) =

Recycling reactions of Co-SBA-1520 were carried out at 373 K for 8 h. The catalyst recovered from the reaction mixture by centrifugation, washing with acetone, and drying at 373 K for 12 h, and then it was used in the next run under the same reaction conditions. 3. Results and discussion 3.1. Structural characteristics of Co-SBA-15x Fig. 1 shows the XRD patterns of pure silica SBA-15 and CoSBA-15x samples prepared by the pH-adjusting method with Si/Co ratios of 10100. Although the d spacing of the rst strong reection had some changes from sample to sample, all the samples exhibited three well-resolved diffraction peaks that can be indexed as the (1 0 0), (1 1 0), and (2 0 0) diffractions associated with hexagonal symmetry. The results indicate that a certain amount of Co can be grafted into SBA-15 by the pH-adjusting method without destroying the structural ordering greatly. The HRTEM images of Co-SBA-1520 also show the well-ordered hexagonal arrays of mesopores with one-dimensional channels (Fig. 1S(A) and (B) in the supplement). The result coincides with the XRD observations, and it further demonstrates that the samples prepared by the pH-adjusting method still keep the highly ordered mesostructure when a certain amount of cobalt species is introduced. N2 adsorption/desorption isotherms of pure silica SBA-15 and Co-SBA-15x with varying cobalt content indicate that all materials gave typical irreversible type IV adsorption/desorption isotherms with H1 hysteresis loop, which is characteristic of the mesoporous materials with 2D-hexagonal structure, as dened by IUPAC [24] (Fig. 2S in the supplement). The sharp steps occurred at a relative partial pressure of 0.70.9, corresponding to the capillary conden-

sation of N2 , which indicates the uniformity of the pores [36]. Pore structure parameters of SBA-15 and Co-SBA-15x calculated from the desorption branch are shown in Table 1. Compared with pure silica SBA-15, the introduction of cobalt made the surface area and pore volume signicantly decrease. For Co-SBA-15x , with the increasing of cobalt content, both the surface area and pore volume gradually reduced. The micropore area and micropore volume of the pure silica SBA-15 were 84.7 m2 /g and 0.029 cm3 /g respectively, but the value of the parameters was reduced once cobalt species was introduced into the molecular sieves. The micropores continued to decrease with an increase of cobalt amount. It is proposed that lling of the micropores leads to their reduction during the pHadjusting procedure. A similar tendency is also observed by Wu et al. [34] and Ryoo and Ko [37]. The changes for these parameters of micropores obviously bring about decrease of the total surface area and pore volume of Co-SBA-15x . Another reason for decrease of the surface area and pore volume is that some of cobalt species is incorporated into mesoporous channels [16,38]. On the other hand, pore diameter had an obvious increase when cobalt was introduced into the molecular sieves, and the parameter rst became larger and then the change was not observable with cobalt content rising (see Table 1). It shows that cobalt species is incorporated into the silica framework of SBA-15 because of the longer bond length of Co with oxygen than that of SiO [36]. As cobalt amount increases in synthesis system, more cobalt ions are introduced into the framework of the mesoporous silica, which makes pore diameter further increase. However, when cobalt amount continues to increase, more cobalt ions are incorporated into mesoporous channels rather than only framework of SBA-15, as conrmed by the change of surface area and pore volume with cobalt content [38] and the observations of UVvis. As a result, pore diameter changes little if cobalt amount is further increased. 3.2. Inuence of synthesis conditions on structure of Co-SBA-15x As Fig. 2 shows, Co-SBA-15x with an appropriate amount of cobalt could retain a highly ordered mesostructure, but higher con-

Fig. 1. XRD patterns of (A) SBA-15, (B) Co-SBA-15100 , (C) Co-SBA-1540 , (D) Co-SBA1520 , and (E) Co-SBA-1510 respectively.

Fig. 2. XRD patterns of (A) SBA-15, (B) Co-SBA-1540 -6.5, (C) Co-SBA-15100 , (D) CoSBA-1540 -7.5, (E) Co-SBA-1540 -8.5, (F) Co-SBA-1510 , (G) Co-SBA-152 , and (H) CoSBA-1540 -9.0 respectively.

48

H. Cui et al. / Applied Catalysis B: Environmental 101 (2010) 4553

tent of Co2+ in the initial gel would destroy the structural ordering of the molecular sieves. For example, almost no diffraction peaks associated with hexagonal symmetry were detected when Si/Co ratio decreased to 2. Metal ions permeate the interstitial regions between the silica and block copolymer before the pH is adjusted. When the system is adjusted from strong acid to neutral, the metal ions transform to the oxo form and condense with the adjacent silanols [34]. In the present work, cobalt species is introduced into the mesoporous walls similarly. However, more metal ions in interstitial regions probably affect the interaction of silica and organic template during the crystallization procedure, resulting in disordered structure [34,37]. Another great effect of the synthesis condition on the structure of Co-SBA-15x is the adjusted pH value. Fig. 2 also shows the XRD patterns of Co-SBA-1540 prepared at different pH values (6.5, 7.5, 8.5 and 9.0). The samples synthesized at pH value lower than 8.0 showed multiple well-resolved XRD peaks. Based on the results of Zhao et al. [39], the absence of sufciently strong electrostatic or hydrogen-bonding interactions at pH 2.07.0 led to the formation of disordered porous silica. However, in this study, highly ordered Co-SBA-15x was successfully synthesized at pH less than 8.0 by the pH-adjusting method. This is consistent with the observations of Lou et al. According to their hypothesis, because sodium hydroxide was added to adjust the pH of synthesis system, the resultant salt (NaCl) helped to improve ordered mesostructure of SBA-15 [36]. Similarly, in our study, the produced NH4 Cl is responsible for effectively improving the ordered structure since ammonia is used to adjust pH during the synthesis. However, when pH further increased, the XRD peaks were hardly observed for the sample prepared at pH 9.0 (see Fig. 2). This indicates that higher pH value (pH > 9.0) negatively inuences the mesostructure of SBA-15, which is in agreement with the results of Zhao et al. The mesoporous silica synthesized under strong acidic condition has an electrically neutral framework [39]. Under weakly alkaline conditions, hydrolysis tends to speed up due to the presence of OH [34]. On the other hand, although lower pH is favorable to highly ordered mesostructure, it results in low amount of cobalt introduction. For instance, when pH was 7.5, 1.3 wt% of Co was introduced into the mesoporous molecular sieves for Co-SBA-1540 , whereas there was only 0.005 wt% of Co content if pH decreased to 6.5, conrmed by ICP-AES analysis shown in Table 3. On the basis of our studies about synthesis of Co-SBA-15x , the optimal pH for preparation of Co-SBA-15x is nally determined to be in the range of 7.28.0. 3.3. Coordination environment and oxidation state of cobalt High-angle XRD patterns of pure silica SBA-15, Co-SBA-1520, Co/SBA-15 and Co3 O4 are shown in Fig. 3. Only a broad peak due to the amorphous feature of SBA-15 framework appeared, and no peaks of the crystalline Co3 O4 were observed for Co-SBA-1520 . While Co/SBA-15 with the same cobalt content prepared by the impregnation method, showed peaks at about 37 , 45 , 60 and 65 , attributed to the crystalline Co3 O4 . It is concluded that cobalt species grafted by the pH-adjusting method is highly dispersed in SBA-15. This is also consistent with the HRTEM observations, indicating that no large particles are located outside the mesopores of Co-SBA-15x . The state of cobalt in the samples was further investigated by Raman spectroscopic measurements (see Fig. 4). The crystalline Co3 O4 exhibited four Raman bands at 467, 510, 603 and 670 cm1 due to the vibrations of CoO bonds [16,40]. Co/SG and Co/SBA15 showed bands at 660 and 657 cm1 respectively, indicating the presence of some amount of Co3 O4 species on the surface of these materials. In addition, the Raman band at 829 cm1 was also observed for Co/SBA-15, perhaps arising from the vibrations

Fig. 3. High-angle XRD patterns of (A) SBA-15, (B) Co-SBA-1520 (Co, 2.1 wt%), (C) Co/SBA-15 (Co, 2.1 wt%), and (D) Co3 O4 .

of CoO that strongly interacts with silica. However, no Raman bands of CoO vibrations appeared for Co-SBA-1520 and Co-MCM41-TIE, indicating the highly dispersed cobalt species in the two samples. The oxidation state of cobalt was investigated by XPS. Fig. 5 exhibits Co 2p XPS spectra of the Co-containing samples. The binding energy of Co 2p3/2 of Co-MCM-41-TIE was 781.5 eV, similar to that of Co-NaX, and the cobalt species with such a binding energy can be assigned to Co(II) in an isolated state [16,41]. It is interesting to note that Co-SBA-1520 and Co-MCM-41-TIE possessed almost the same binding energy of Co 2p3/2 , suggesting that the

Fig. 4. Raman spectra of (A) Co-SBA-1520 (Co, 2.1 wt%), (B) Co-MCM-41-TIE (Co, 2.1 wt%), (C) Co/SG (Co, 2.1 wt%), (D) Co/SBA-15 (Co, 2.1 wt%), and (E) Co3 O4 respectively.

Fig. 5. Co 2p XPS spectra of (A) Co-MCM-41-TIE (Co, 2.1 wt%), (B) Co-SBA-1520 (Co, 2.1 wt%), (C) Co/SBA-15 (Co, 2.1 wt%), and (D) Co3 O4 respectively.

H. Cui et al. / Applied Catalysis B: Environmental 101 (2010) 4553

49

3.4. Studies of styrene epoxidation with O2 catalyzed by various metal species based on SBA-15 Though epoxidation of styrene with molecular oxygen is considered as a promising process for production of the epoxide, it is one of the most challenging subjects. This is because triplet ground state of O2 disfavors reactions with singlet organic compounds [17,18]. Therefore, many researchers have pursued new effective catalytic systems for the reaction. In this study, in order to compare which metal species was highly active for the reaction, some metal species that can activate oxygen in the oxidation [4752] were introduced into SBA-15 with their precursors by the pH-adjusting method. Table 2 summarizes catalytic performances of the catalysts in the epoxidation of styrene with O2 . Co-SBA-1520 achieved the highest styrene conversion (94.1%) and epoxide selectivity (65.5%) under the same reaction conditions. Other metal species were all inferior to cobalt for the reaction. According to previous studies, the coordination of DMF to cobalt(II) affected the ability of O2 to bind to the Co(II) and their redox potential, and this resulted in the activation of O2 in the epoxidation [16,53,54]. In conclusion, cobalt is an effective component for the epoxidation of styrene with O2 . 3.5. Investigations of styrene epoxidation with O2 over various cobalt-based catalysts Table 3 shows the effect of preparation steps for the cobaltbased catalysts on the epoxidation of styrene with O2 . Co-SBA-1520 prepared by the pH-adjusting method showed the best catalytic performance, with 94.1% of styrene conversion and 65.5% of epoxide selectivity. The catalytic activity of Co-SBA-1520 -1 dropped sharply, with 36.7% of the conversion and 52.1% of the epoxide selectivity. The cobalt species in the sample mainly existed in the form of Co3 O4 from the result of XRD (not shown). In addition, if the pH of synthesis system was not adjusted, the sample expressed as Co-SBA-1520 -2 showed negligible catalytic activity, with only 3.4% of styrene conversion, ascribed to no cobalt sites in the sample. It is concluded that cobalt ions in the initial gel are difcult to be introduced into the mesostructure without adjusting pH, as evidenced by ICP-AES. When an cobalt source is added into the initial gels, the cobalt species is dispersed in the synthesis system in the form of Co2+ because of the strong acidic condition. Co2+ permeates the interstitial regions between the silica and block copolymer before pH adjusting. The cobalt ions are possibly coated with silicablock copolymer mesophase, but have almost no linkage with the silicates in strong acidic media. As a result, few of cobalt species can be introduced if the semi-product is ltered before pH adjusting. This phenomenon is in agreement with the mechanism for the formation of Al-SBA-15 and Ti-SBA-15 proposed by Wu et al. [34]. Moreover, if the surfactants in Co-SBA-1520 were not removed, the sample designated as Co-SBA-1520 -3 had only 8.8% of styrene conversion, much lower than that of surfactant removing. When the pores are full of template, the reactants in the reaction system cannot enter the pores to contact with a large number of cobalt sites in the particles. However, when the template is extracted from pores, almost all the active sites participate in the reaction.

Fig. 6. TPR proles of (A) Co-MCM-41-TIE (Co, 2.1 wt%), (B) Co-SBA-1520 (Co, 2.1 wt%), (C) Co/SBA-15 (Co, 2.1 wt%), and (D) Co3 O4 respectively.

cobalt species in Co-SBA-1520 is also mainly in the single-site Co(II) state. In addition, Co/SBA-15 prepared by the impregnation method exhibited a binding energy of Co 2p3/2 close to that of Co3 O4 . The result coincides with the observations of XRD, Raman and HRTEM. The state of cobalt species of the samples was also investigated by TPR. As Fig. 6 shows, two reduction peaks at 573673 K were mainly found for Co3 O4 and Co/SBA-15, probably ascribed to the stepwise reduction of Co3 O4 to CoO and then CoO to Co [42,43]. Co-SBA-1520 and Co-MCM-41-TIE seemed to show the reduction peaks at almost the same high temperature (about 1044 K), further suggesting that the cobalt species in Co-SBA-15x is similar to that in Co-MCM-41-TIE, and is likely in the single-site Co(II) state. In addition, for Co/SBA-15 prepared by the impregnation method, a relatively weak reduction peak was also observed, which is the reduction of some cobalt species that strongly interacts with the support [16]. Electronic spectroscopy in UVvis region is considered as an effective technique for studying the local coordination environment and electronic state of isolated transition metal ions as well as aggregated transition metal oxides [44]. Co-SBA-1520 was further investigated by UVvis spectroscopy (see Fig. 3S in the supplement). The absorbance in the range of 500680 nm is characteristic of Co(II) in an isolated state [45]. The peaks at 530, 580 and 650 nm for the sample show that the cobalt species in Co-SBA-1520 mainly exists in the state of single-site Co(II), coincident with the results of XPS and TPR. Moreover, the peaks in the range of 550700 nm are absorbance of tetrahedrally coordinated Co2+ [46]. The sample indicated tetrahedrally coordinated Co2+ in the silica framework at 580, 650 nm, and most of cobalt species was incorporated into the framework of SBA-15. This is similar to the result reported by Lou et al. [36]. In addition, a weak peak at 350 nm is attributed to absorbance of Co3 O4 [36], indicating that a small part of cobalt species is introduced into the mesoporous channels in the state of Co3 O4 . In this work, the amount of Co3 O4 increased with the rise of cobalt content in the samples.
Table 2 Epoxidation of styrene with O2 catalyzed by various metal species based on SBA-15a . Catalyst Fe-SBA-1520 Ti-SBA-1520 Ni-SBA-1520 Mn-SBA-1520 Sn-SBA-1520 Co-SBA-1520
a

Metal content (wt%) 4.1 3.3 1.9 1.2 5.5 2.1

Surface area (m2 /g) 299 303 298 301 298 298

Pore volume (cm3 /g) 1.04 1.06 1.02 1.04 1.06 1.03

Pore diameter (nm) 13.9 14.0 13.7 13.8 14.2 13.8

Styrene conversion (%) 12.3 5.8 5.3 9.4 3.5 94.1

Epoxide selectivity (%) 47.1 56.8 42.7 41.0 30.3 65.5

Reaction condition: styrene, 8 mmol; DMF, 25 g; catalyst, 0.3 g; temperature, 373 K; time, 8 h; ow rate of O2 , 15 ml/min.

50

H. Cui et al. / Applied Catalysis B: Environmental 101 (2010) 4553

Table 3 Epoxidation of styrene with O2 catalyzed by various cobalt-based catalystsa . Catalyst Co content (wt%) Styrene conversion (%) Product selectivity (%) Benab Co-SBA-1520 -1 Co-SBA-1520 -2 Co-SBA-1520 -3 Co-SBA-1540 -6.5 Co-SBA-1540 -7.5 Co-SBA-1540 -9.0 Co-SBA-15100 Co-SBA-1540 Co-SBA-1520 Co-SBA-1510 Co/SBA-15 Co/SG Co-MCM-41-TIE Co-MCM-41-TIE Co(NO3 )2
a b c

Epoxide 52.1 38.9 36.7 63.9 64.3 55.6 65.1 64.3 65.5 65.5 55.6 55.1 57.9 58.3 53.8

Pheac 4.3 3.1 10.9 8.0 8.6 7.4 7.3 8.6 8.2 7.8 7.4 8.1 4.9 6.2 0

Other 13.1 0.9 9.0 13.0 12.5 15.5 12.0 12.5 11.6 11.5 15.5 20.6 7.8 7.1 0

4.5 Trace 2.0 0.005 1.3 1.28 0.12 1.3 2.1 3.6 2.1 2.1 2.1 3.6 /

36.7 3.4 8.8 10.9 88.9 42.5 46.8 88.9 94.1 93.9 43.5 18.6 54.2 54.6 1.3

30.5 57.1 43.5 15.1 14.6 21.5 15.6 14.6 14.7 15.2 21.5 16.2 29.4 28.3 46.2

Reaction condition: styrene, 8 mmol; DMF, 25 g; catalyst, 0.3 g; Co(NO3 )2 , 0.11 mmol; temperature, 373 K; time, 8 h; ow rate of O2 , 15 ml/min. Benzaldehyde. Phenylacetaldehyde.

Consequently, the catalytic activity of Co-SBA-1520 is much higher than that of Co-SBA-1520 -3. In contrast, if all the cobalt species is anchored to the external surface, the catalytic activity of the sample will not be affected by the blockage of the template and the two data should be the same. But the result is not so, indicating that most of the cobalt sites are exposed on the inner face of the pores rather than only on the external surface, and the epoxidation of styrene with O2 mainly occurs on the internal surface of pores. This is consistent with the result of Wu et al. [34], in which Al in Al-SBA1510 prepared by the same method was also chiey situated in the interior of particles. In conclusion, through comparison, different processing methods lead to the catalytic materials with signicant different results although the same amounts of identical materials are added in the initial gels at the beginning of preparation process. And some valuable information was also obtained from the comparison. Table 3 also summarizes the catalytic performance of Co-SBA1540 prepared at different pH. The sample prepared at pH 7.5 exhibited the highest conversion (88.9%) and epoxide selectivity (64.3%). If the pH of synthesis system was adjusted to 6.5, the catalyst obtained only 10.9% of styrene conversion. This further conrms that lower pH leads to the introduction of less cobalt into the molecular sieves although it helps to attain well-ordered mesostructure, and low levels of cobalt sites undoubtedly bring about low activity. In addition, the ordered mesostructure of SBA15 was hardly obtained according to the XRD observation when the pH was adjusted to 9.0. The catalyst showed 42.5% of styrene conversion, much lower than that of the sample with pH 7.5, indicating the sample that possesses well-ordered mesostructure of SBA-15 is much superior to that with disordered structure for the epoxidation of styrene with O2 . Table 3 also lists the catalytic performance of various cobaltbased catalysts in the epoxidation of styrene with O2 . Catalytic activity of the cobalt-based samples in the reaction decreased in the sequence Co-SBA-1520 > Co-MCM-41-TIE > Co/SBA-15 > Co/SG with the same cobalt content (2.1 wt%). Co(NO3 )2 was almost inactive in the reaction. According to the results of XRD, HRTEM, Raman, XPS, TPR, and UVvis illustrated above, the cobalt in both Co-SBA15x and Co-MCM-41-TIE basically existed in the single-site Co(II) state. But Co-SBA-1520 achieved much higher activity than CoMCM-41-TIE, indicative of the difference in the performance for the two cobalt-functionalized mesoporous silica. Moreover, Co/SG mainly aggregated into Co3 O4 particles, whereas Co/SBA-15 contained highly dispersed Co3 O4 particles and a small part of cobalt

species strongly interacting with support. It is likely that better dispersion of cobalt species results in higher catalytic activity. By comparison, the single-site Co(II) is the most active for the epoxidation of styrene with O2 . As Table 3 shows, cobalt content in Co-SBA-15x exerted an obvious impact on the catalytic performance in the reaction. Apparently, the conversion of styrene increased with the rising of cobalt content: Co-SBA-15200 < Co-SBA-15100 < Co-SBA-1540 < CoSBA-1520 . But the styrene conversion kept nearly unchanged when the Si/Co ratio was less than 20 for Co-SBA-15x , showing that it has little effect on the catalytic activity when cobalt content is higher. As we know, raising cobalt content leads to an increase of cobalt sites. However, based on the characterizations above, this also resulted in the formation of less active Co3 O4 which is introduced into the mesoporous channels. The amount of such cobalt species tended to increase if cobalt content in the sample rose. So the increase of the cobalt species has unconspicuous effect to improve the activity of Co-SBA-15x . In addition, Co content in Co-SBA-15x made no difference to the epoxide selectivity (see Table 3). 3.6. Studies of reaction conditions in epoxidation of styrene with O2 catalyzed by Co-SBA-1520 Generally, the catalytic activity and product selectivity in the epoxidation of styrene are strongly inuenced by reaction temperature, catalyst amount, solvent, solvent amount, reaction time, oxidant, and oxidant amount, which were systematically investigated as follows. Usually the reaction temperature has great inuence on catalytic performance. Fig. 7 presents the inuence of reaction temperature on the epoxidation of styrene with O2 over CoSBA-1520 . The conversion of styrene increased quickly with the increasing of reaction temperature at rst. For example, styrene conversion was 11.9% at 333 K, quickly rose to 58.2% at 353 K, and to 94.1% at 373 K. But styrene conversion started to decrease with further increase of temperature. The styrene conversion decreased to 89.0% at 423 K, and polymerization of styrene tended to be observably aggravated when the reaction temperature was higher than 393 K. The selectivity of epoxide followed up the trend of the styrene conversion with the temperature, which rst increased from 39.2% at 333 K to the maximum (65.5%) at 373 K, and then slightly reduced to 61.6% at 393 K and to 56.2% at 423 K. It seems that an appropriate reaction temperature is favorable to the formation of epoxide for the reaction.

H. Cui et al. / Applied Catalysis B: Environmental 101 (2010) 4553

51

Fig. 7. Effect of reaction temperature on epoxidation of styrene with O2 over Co-SBA-1520 , ( ) styrene conversion, ( ) epoxide selectivity, ( ) benzaldehyde selectivity, ( ) phenylacetaldehyde selectivity, ( ) other selectivity. Reaction condition: styrene, 8 mmol; DMF, 25 g; catalyst, 0.3 g; time, 8 h; ow rate of O2 , 15 ml/min.

Fig. 9. Effect of DMF amount on epoxidation of styrene with O2 over Co-SBA1520 , ( ) styrene conversion, ( ) benzaldehyde selectivity, ( ) epoxide selectivity, ( ) phenylacetaldehyde selectivity, ( ) other selectivity. Reaction condition: styrene, 8 mmol; catalyst, 0.3 g; temperature, 373 K; time, 8 h; oxidant amount, O2 , 15 ml/min.

Fig. 8. Effect of Co-SBA-1520 amount on epoxidation of styrene with O2 , ( ) styrene conversion, ( ) epoxide selectivity, ( ) benzaldehyde selectivity, ( ) phenylacetaldehyde selectivity, ( ) other selectivity. Reaction condition: styrene, 8 mmol; DMF, 25 g; temperature, 373 K; time, 8 h; ow rate of O2 , 15 ml/min.

The effect of Co-SBA-1520 amount on the epoxidation of styrene is depicted in Fig. 8. The conversion of styrene increased from 69.9% over 0.05 g of Co-SBA-1520 to the maximum (94.1%) with 0.3 g of the catalyst, and then decreased slowly to 90.1% using 0.7 g of the sample. But the epoxide selectivity kept more or less a constant about 65.0% in the range of catalyst amount from 0.05 g to 0.7 g. According to previous research work, increasing the catalyst amount, more cobalt sites would be obtained to coordinate with the solvent (DMF), resulting in the true active sites during the reaction or acting as a co-reductant [16,18,55]. However, the observations of Fig. 8 indicate that the raise of styrene conversion cannot merely depend on the increase of catalyst amount. Table 4 presents the inuence of solvent on the styrene epoxidation with O2 over Co-SBA-1520 at 373 K. It is found that solvent
Table 4 Effect of reaction solvent on epoxidation of styrene with O2 over Co-SBA-1520 a . Solvent Styrene conversion (%) Product selectivity (%) Bena DMF DMA DMSO Cyclohexanone Pyridine Toluene 94.1 95.3 47.9 50.5 0 0 14.7 15.0 89.3 66.3 b Epoxide 65.5 64.5 4.3 27.2 Phea 8.2 8.9 1.2 2.8 Other 11.6 11.5 5.2 3.7

a Reaction condition: styrene, 8 mmol; solvent, 25 g; catalyst, 0.3 g; temperature, 373 K; time, 8 h; ow rate of O2 , 15 ml/min. b means none.

played a great role on the epoxidation reaction. The amides such as N,N -dimethylformamide (DMF) and N,N-dimethylacetamide (DMA) were especially efcient in obtaining both high styrene conversion and high epoxide selectivity. Although a certain styrene conversion was obtained in dimethyl sulfoxide (DMSO) and cyclohexanone, the epoxide selectivity was low, and the main product was benzaldehyde. When pyridine or toluene was used as reaction solvent, no products were detected. Such results are in agreement with the observations of Tang et al. [16]. Based on their results, the solvent (DMF) could coordinate with cobalt ions in the molecular sieves, and such coordination was an important factor for the activation of O2 in the epoxidation reaction. Recently, some studies on the effect of solvent have been reported [1618,56], but the potential causes are controversial. Rao et al. proposed that the aprotic solvents with high dielectric constant were favorable to high selectivity of epoxide, but the observations of Zhan et al. were not so. In fact, many factors including the nature and structure of the solvents can affect the results. The effect of solvent on the reaction may be quite complicated. The effect of DMF amount on the epoxidation of styrene with O2 over Co-SBA-1520 at 373 K is illustrated in Fig. 9. The styrene conversion rose with the increase of DMF amount when DMF amount was less than 24 g. If DMF amount continued to increase, the styrene conversion decreased slowly. The change of the epoxide selectivity with DMF amount was similar to the trend of the styrene conversion. The epoxide selectivity rose from 45.9% with 8 g of DMF to the maximal value (65.5%) using 24 g of DMF, and then reduced to 50.2% when 48 g of DMF was used. It should be noted that the selectivity of other products (including benzoic acid, phenylacetic acid, etc.) changed in the opposite direction to that of the epoxide with DMF amount. These results suggest that an appropriate amount of solvent should be benecial for the formation of epoxide and meanwhile inhibits overoxidation of styrene to acids. Fig. 9 indicates the optimal DMF amount is 24 g. When such amount of DMF is used, both styrene conversion and epoxide selectivity reach the highest (94.1% and 65.5% respectively) but the selectivity of acids is lowest. The inuence of reaction time on the epoxidation of styrene with O2 over Co-SBA-1520 at 373 K was also investigated, as shown in Fig. 10. The styrene conversion increased with the prolonging of reaction time. When reaction time was less than 8 h, the styrene conversion rose quickly from 58.3% in 2 h to 94.1% in 8 h, showing the induction period characteristic of radical reaction. Yet the styrene conversion increased slowly if reaction time was longer than 8 h. The effect of reaction time on the epoxide selectivity was not obvious. The epoxide selectivity slowly increased from 56.7%

52

H. Cui et al. / Applied Catalysis B: Environmental 101 (2010) 4553

Fig. 10. Effect of reaction time on epoxidation of styrene with O2 over CoSBA-1520 , ( ) styrene conversion, ( ) epoxide selectivity, ( ) epoxide yield (epoxide yield = styrene conversion epoxide selectivity 100%). Reaction condition: styrene, 8 mmol; DMF, 25 g; catalyst, 0.3 g; temperature, 373 K; ow rate of O2 , 15 ml/min. Table 5 Epoxidation of styrene by various oxidants over Co-SBA-1520 a . Oxidant Styrene conversion (%) Product selectivity (%) Bena None O2 H2 O2 TBHP Air 0 94.1 50.8 40.7 93.9 14.7 24.4 18.1 15.0 Epoxide 65.5 58.8 72.4 65.4 Phea 8.2 2.5 9.4 7.8 Other 11.6 14.3 0 11.8

Fig. 12. Recycling investigations of Co-SBA-1520 in epoxidation of styrene with O2 , ( ) styrene conversion, ( ) epoxide selectivity. Reaction condition: styrene, 8 mmol; DMF, 25 g; catalyst, 0.3 g; temperature, 373 K; time, 8 h; ow rate of O2 , 15 ml/min.

a Reaction condition: styrene, 8 mmol; DMF, 25 g; catalyst, 0.3 g; temperature, 373 K; time, 8 h; Oxidant amount, O2 , 15 ml/min; H2 O2 , 10 mmol; TBHP, 10 mmol; air, 40 ml/min.

in 2 h to 65.5% in 8 h, and then began to decrease when reaction time continuously increased. As we know, prolonging reaction time makes the substrate fully react. But it will also lead to overoxidation of styrene to acid and polymerization of styrene at the same time, which results in the decrease of epoxide selectivity. Therefore, the optimal reaction time should be 8 h from our studies, in which the highest epoxide yield can be achieved. Table 5 compares the effect of various oxidants on the epoxidation of styrene catalyzed by Co-SBA-1520 . No epoxidation of styrene occurred when the reaction was carried out under the inert atmosphere (N2 or Ar) without any oxidants. If O2 or air was used as the oxidant, the highest styrene conversion was obtained. The styrene conversion was 50.8% when H2 O2 was used as the oxidant, which

was much lower than that with O2 or air, mainly because of the rapid decomposition of H2 O2 into O2 at the initial stage. Though 72.4% of epoxide selectivity was obtained when TBHP was used, the styrene conversion was the lowest, and polymerization of styrene was a serious problem. Through comparison, O2 or air is a suitable oxidant for the reaction over the cobalt-based catalyst. This also indicates that Co-SBA-15x is particularly effective in the activation of O2 for the epoxidation of olens. Fig. 11 shows the inuence of ow rate of O2 on the epoxidation of styrene over Co-SBA-1520 at 373 K. The styrene conversion was 77.4% when ow rate was 5 ml/min, and then slowly increased to 94.1% with the rising of ow rate of O2 to 15 ml/min. The styrene conversion changed little when the ow rate of O2 was higher than 15 ml/min. The ow rate had little impact on the product selectivity, and the overoxidation was not basically aggravated in the range of ow rate of O2 studied. 3.7. Recycling investigations of Co-SBA-1520 in epoxidation of styrene with O2 Catalyst recycling experiments were performed with repeated use of Co-SBA-1520 at 373 K for 8 h. It can be seen from Fig. 12 that the styrene conversion and the epoxide selectivity kept unchanged by and large with the reuse of Co-SBA-1520 for seven times. Little leaching of cobalt from the catalyst was found in the reaction mixture by the result of the elemental analyzer, further indicative of excellent stability and the recyclability of Co-SBA-1520 in the epoxidation of styrene with O2 . 4. Conclusions Highly ordered cobalt-substituted SBA-15 was successfully synthesized by the pH-adjusting method. The cobalt species in Co-SBA-15 was mainly in the single-site Co(II) state. Most of them were incorporated into the silica framework of SBA-15, existing in tetrahedral coordination. Co-SBA-1520 showed excellent catalytic performance in the epoxidation of styrene with O2 . It exhibited much higher activity than Co-MCM-41-TIE and Co/SBA-15. The optimized preparation conditions for Co-SBA-15x were Si/Co ratio 1020 and pH of synthesis system 7.28.0. The material could obtain the best catalytic performance under the reaction conditions such as reaction temperature 373 K, 0.3 g of the catalyst, 24 g of DMF (DMA) as reaction solvent, reaction time 8 h and 15 ml/min of O2 (40 ml/min of air) as oxidant if 8 mmol of styrene was added to the reaction system. Co-SBA-1520 possessed excellent stability and recyclability in the reaction.

Fig. 11. Effect of ow rate of O2 on epoxidation of styrene with O2 over Co-SBA-1520 , ( ) styrene conversion, ( ) epoxide selectivity, ( ) benzaldehyde selectivity, ( ) phenylacetaldehyde selectivity, ( ) other selectivity. Reaction condition: styrene, 8 mmol; DMF, 25 g; catalyst, 0.3 g; temperature, 373 K; time, 8 h.

H. Cui et al. / Applied Catalysis B: Environmental 101 (2010) 4553

53

Acknowledgement We are grateful for the nancial support of the National Key Technology R&D Program (No. 2007BAB24B04). Appendix A. Supplementary data Supplementary data associated with this article can be found, in the online version, at doi:10.1016/j.apcatb.2010.09.003. References
[1] M. Caldarelli, J. Habermann, S.V. Ley, J. Chem. Soc., Perkin Trans. 1 (1999) 107110. [2] Q.-H. Xia, X. Chen, T. Tatsumi, J. Mol. Catal. A: Chem. 176 (2001) 179193. [3] M.P. Kapoor, A.K. Sinha, S. Seelan, S. Inagaki, S. Tsubota, H. Yoshida, M. Haruta, Chem. Commun. (2002) 29022903. [4] A.K. Sinha, S. Seelan, S. Tsubota, M. Haruta, Angew. Chem. Int. Ed. 43 (2004) 15461548. [5] Y. Liu, K. Murata, M. Inaba, Chem. Commun. (2004) 582583. [6] J. Huang, T. Takei, T. Akita, H. Ohashi, M. Haruta, Appl. Catal. B: Environ. 95 (2010) 430438. [7] E. Ruckenstein, H.Y. Wang, Appl. Catal. A 204 (2000) 257263. [8] A. Karim, Y. Su, J. Sun, C. Yang, J. Strohm, D. King, Y. Wang, Appl. Catal. B: Environ. 96 (2010) 441448. [9] T.A. Kainulainen, M.K. Niemela, A.O.I. Krause, Catal. Lett. 53 (1998) 97101. [10] K. Okabe, X. Li, M. Wei, H. Arakawa, Catal. Today 89 (2004) 431438. [11] A.Y. Khodalkov, R. Bechara, A. Griboval-Constant, Appl. Catal. A: Gen. 254 (2003) 273288. [12] J. Panpranot, S. Kaewkun, P. Praserthdam, J.G. Goodwin, Catal. Lett. 91 (2003) 95102. [13] S.W. Ho, J.M. Cruz, M. Houalla, D.M. Hercules, J. Catal. 135 (1992) 173185. [14] C. Ando, H. Kurokawa, H. Miura, Appl. Catal. A 185 (1999) 181183. [15] T. Tsoncheva, L. Ivanova, J. Rosenholm, M. Linden, Appl. Catal. B: Environ. 89 (2009) 365374. [16] Q. Tang, Q. Zhang, H. Wu, Y. Wang, J. Catal. 230 (2005) 384397. [17] G. Xu, Q.-H. Xia, X.-H. Lu, Q. Zhang, H.-J. Zhan, J. Mol. Catal. A: Chem. 266 (2007) 180187. [18] H.-J. Zhan, Q.-H. Xia, X.-H. Lu, Q. Zhang, H.-X. Yuan, K.-X. Su, X.-T. Ma, Catal. Commun. 8 (2007) 14721478. [19] J.S. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vartuli, J.S. Beck, Nature 359 (1992) 710712. [20] P. Yang, D. Zhao, D. Margolese, B. Chmelka, G. Stucky, Nature 396 (1998) 152154. [21] M. Morey, S. OBrien, G.D. Stucky, Chem. Mater. (2000) 898911.

[22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] [56]

R. Mokaya, J. William, Chem. Commun. (1997) 21852186. S. Sumiya, Y. Oumi, T. Uozumi, T. Sano, J. Mater. Chem. (2001) 11111115. Z. Luan, J.Y. Bae, C. Kevan, Chem. Mater. (2000) 32023207. R. Ryoo, S. Jun, J.M. Kim, M.J. Kim, Chem. Commun. (1997) 22252226. Z. Luan, M. Hartmann, D. Zhao, W. Zhou, L. Kevan, Chem. Mater. (1999) 16211627. R. Mokaya, Angew. Chem. 111 (1999) 30793083. Z. Luan, E.M. Maes, P.A.W. van der Heide, D. Zhao, R.S. Czernuszewicz, L. Kevan, Chem. Mater. (1999) 36803686. Y. Liu, T.J. Pinnavaia, Chem. Mater. (2002) 35. Y. Han, N. Li, L. Zhao, D. Li, X. Xu, S. Wu, Y. Di, C. Li, Y. Zou, Y. Yu, F.-S. Xiao, J. Phys. Chem. B 107 (2003) 75517556. Y. Han, F.-S. Xiao, S. Wu, Y. Sun, X. Meng, D. Li, S. Lin, J. Phys. Chem. B 105 (2001) 79637966. B. Newalkar, J. Olanrewaju, S. Komarneni, Chem. Mater. (2001) 552557. W. Zhang, Q. Lu, B. Han, M. Li, J. Xiu, P. Ying, C. Li, Chem. Mater. (2002) 34133421. S. Wu, Y. Han, Y.-C. Zou, J.-W. Song, L. Zhao, Y. Di, S.-Z. Liu, F.-S. Xiao, Chem. Mater. (2004) 486492. A. Vinu, M. Hartmann, Chem. Lett. 33 (2004) 588589. Z. Lou, R. Wang, H. Sun, Y. Chen, Y. Yang, Microporous Mesoporous Mater. 110 (2008) 347354. R. Ryoo, C. Ko, J. Phys. Chem. B 104 (2000) 1146511471. Q. Zhang, Y. Wang, S. Itsuki, T. Shisshido, K. Takehira, J. Mol. Catal. A 188 (2002) 189200. D. Zhao, Q. Huo, J. Feng, B. Chmelka, G. Stucky, J. Am. Chem. Soc. 120 (1998) 60246036. M.A. Vuurman, D.J. Stufkens, A. Oskam, G. Deo, I.E. Wachs, J. Chem. Soc., Faraday Trans. 92 (1996) 32593265. L. Guczi, D. Bazin, Appl. Catal. A 188 (1999) 163174. S. Sun, N. Tsubaki, K. Fujimoto, Appl. Catal. A 202 (2000) 121131. L.A. Boot, M.H.J.V. Kerkhoffs, A.J. van Dillen, J.W. Geus, F.R. van Buren, B.T. van der Linden, Appl. Catal. A 137 (1996) 6986. S. Bordiga, R. Buzzoni, F. Geobaldo, C. Lamberti, E. Giamello, A. Zecchina, G. Petrini, G. Vlaic, J. Catal. 158 (1996) 486501. A. Verberckmoes, M. Uytterhoeven, R. Schoonheydt, Zeolites 19 (1997) 180189. Y. Brik, M. Kacimi, M. Ziyad, F. Verduraz, J. Catal. 202 (2001) 118128. Y. wang, K. Otsuka, J. Catal. 171 (1997) 106114. E. Heracleous, A.A. Lemondou, J. Catal. 237 (2006) 162174. N.G. Valente, L.A. Arra, L.E. Cads, Appl. Catal. A: Gen. 205 (2001) 201214. Y. Liu, K. Murata, M. Inaba, N. Mimura, Appl. Catal. A 309 (2006) 91105. W. Zeng, J. Li, S. Qin, Inorg. Chem. Commun. 9 (2006) 1012. J. Li, W. Ma, Y. Huang, X. Tao, J. Zhao, Y. Xu, Appl. Catal. B: Environ. 48 (2004) 1724. V.T. Yilmaz, Y. Topcu, Thermochim. Acta 307 (1997) 143147. A. Pui, Croat. Chem. Acta 75 (2002) 165173. J. Jiang, R. Li, H. Wang, Y. Zheng, H. Chen, J. Ma, Catal. Lett. 120 (2008) 221228. S. Rao, K. Munshi, N. Rao, J. Mol. Catal. A: Chem. 156 (2000) 205211.

Вам также может понравиться