Vous êtes sur la page 1sur 5

Catalysis Communications 26 (2012) 122126

Contents lists available at SciVerse ScienceDirect

Catalysis Communications
journal homepage: www.elsevier.com/locate/catcom

Short Communication

Site-selective ethanol conversion over supported copper catalysts


Andr G. Sato a, Diogo P. Volanti b, Isabel C. de Freitas a, Elson Longo b, Jos Maria C. Bueno a,
a b

Departamento de Engenharia Qumica, Universidade Federal de So Carlos, C.P. 676, 13565-905, So Carlos, SP, Brazil Departamento de Fsico-Qumica, Instituto de Qumica, Universidade Estadual Paulista, 14800-900, Araraquara, SP, Brazil

a r t i c l e

i n f o

a b s t r a c t
The origin and inuence of active sites on supported copper catalysts, and their catalytic properties for ethanol conversion, were investigated using Cu/SiO2, Cu2O/SiO2, and Cu/ZrO2. Diffuse reectance infrared Fourier transform spectroscopy of adsorbed CO revealed that Cu 1 + was more prevalent than Cu0 species at the Cu/ SiO2 surface, and that Cu 0 was the main species on Cu/ZrO2. The Cu 1 +/Cu 0 pair provided by Cu/SiO2 generates sites that are highly selective for the dehydrogenation of ethanol. Ethyl acetate can be formed from ethanol and acetaldehyde at ZrO2 or Cu surfaces. A combination of the Cu 1 +/Cu 0 pair and Cu 0 interfaced to ZrO2 is needed to efciently transform ethanol to ethyl acetate. 2012 Elsevier B.V. All rights reserved.

Article history: Received 20 April 2012 Received in revised form 11 May 2012 Accepted 14 May 2012 Available online 18 May 2012 Keywords: DRIFTS XANES Zirconia polymorphs Hydrogen Copper catalysts Ethanol chemistry

1. Introduction There is increased interest in ethanol chemistry as a way to increase the production of chemicals such as ethyl acetate, acetaldehyde, hydrogen, and others. The approach, based on green chemistry, employs renewable resources [13]. Copper-based catalysts have been successfully employed for the selective conversion of ethanol to ethyl acetate or acetaldehyde [4,5]. The best results for selective conversion to ethyl acetate have been achieved with ZrO2-supported copper catalysts [613]. In the case of the Cu/ZnO/Al2O3/ZrO2 catalyst, the active site for the coupling of ethanol and aldehyde is at the mixed metal-oxide surface, not at the Cu metal surface [9]. In addition, the co-existence of Cu 0 and Cu 1 + over Cu/ZrO2 might provide a synergistic interaction for the conversion of ethanol to ethyl acetate [10]. Although the presence of a certain amount of Cu 1 + species over a Cu/ZrO2 catalyst can be important for improving the selectivity to ethyl acetate, the contribution of support oxide phases is not clear. A full understanding of the properties of the active sites for conversion of ethanol to ethyl acetate is crucial for the development of a selective catalyst. In this paper, we investigate the properties of Cu species in the Cu/ ZrO2 and Cu/SiO2 catalysts, in order to understand the nature of the active sites that contribute to the special abilities of Cu/ZrO2 for ethanol conversion. The prepared catalysts were characterized by X-ray diffraction (XRD) and temperature-programmed reduction (TPR). In-situ X-ray absorption spectroscopy (XANES) measurements

supported the proposition that there are changes in Cu speciation in different supports during TPR in H2. Finally, diffuse reectance infrared Fourier transform spectroscopy (DRIFTS) was used to conrm the metalsupport interactions. The ndings were supported by experimental results for CO adsorption on Cu-supported catalysts. 2. Experimental 2.1. Catalyst preparation Commercial grade monoclinic ZrO2 (Saint-Gobain NorPro) and amorphous SiO2 (AEROSIL 380, Degussa) were used as supports. Cu/ ZrO2 and Cu/SiO2 samples were prepared by incipient wetness impregnation of the support with a solution of Cu(NO3)22.5H2O (98%, Aldrich) in methanol. The solids were dried at 373 K overnight, and calcined at 673 K for 5 h under a ow of synthetic air. The theoretical amount of Cu was about 10 wt.%. Cu supported on ZrO2 and SiO2 is hereafter referred to as Cu/ZrO2 and Cu/SiO2, respectively. In an attempt to evaluate the performance of the Cu 1 + species in the ethanol reaction, SiO2 was impregnated with Cu2O (reagent grade, Aldrich), resulting in a Cu2O/SiO2 sample containing 10 wt.% of copper. 2.2. Methods The crystallographic phase of the supports and the Cu catalysts was analyzed by ex-situ XRD. XRD patterns were collected using a Rigaku DMax 2500PC instrument, operated at 40 kV and 150 mA, with Cu K ( = 1.5406 ) radiation. The data were collected in a 2 range from 20 to 45, with a 0.5 divergence slit and a 0.3 mm receiving slit, in xed-time mode with a 0.02 step size.

Corresponding author. Tel.: + 55 16 33518439; fax: + 55 16 33518466. E-mail address: jmcb@ufscar.br (J.M.C. Bueno). 1566-7367/$ see front matter 2012 Elsevier B.V. All rights reserved. doi:10.1016/j.catcom.2012.05.008

A.G. Sato et al. / Catalysis Communications 26 (2012) 122126

123

Temperature-programmed reduction (TPR) proles of the samples were performed with a Micromeritics Pulse Chemisorb 2705 instrument. A ow of H2/He (30 mL min 1) was passed through the sample, and the temperature was raised to 523 K at a rate of 10 K min 1 while recording the thermal conductivity detector (TCD) signal. In-situ XANES spectra at the Cu K-edge (8979 eV) were measured at the D06A-DXAS beam line of the Brazilian National Synchrotron Light Laboratory (LNLS), in Campinas, Brazil. The D06A-DXAS beam line is equipped with a curved Si (111) focusing monochromator, operated in Bragg mode, for the selection of the desired range of X-ray wavelengths (89009400 eV). The samples were prepared as selfsupporting pellets containing ~25 mg of catalyst, and placed into a tubular quartz furnace (di = 20 mm, X-ray path length = 440 mm), sealed with refrigerated kapton windows, for the transmission measurements. Time-resolved XANES spectra at the Cu K-edge were acquired during TPR under a ow of H2/He (5:95), with heating from room temperature to 800 K at a rate of 10 K min 1 (XANES-TPR). Information about the proportions of CuO, Cu2O, and Cu metal evolved in the H2-TPR experiments using Cu/SiO2 and Cu/ZrO2 was achieved by the linear combination tting (LCF) method. In the LCF method, the X-ray absorption spectrum is modeled by least-squares tting using a linear combination of known species (standards) to t an unknown sample spectrum. The diffuse reectance FT-IR (DRIFT) spectra of adsorbed CO were recorded using a Thermo Nicolet 4700 Nexus FTIR spectrophotometer, tted with an MCT detector and a Spectra Tech reaction cell with CaF2 windows (DRIFT HTHV cell). IR spectra were collected using 64 scans at 4 cm 1 resolution. Before reduction, the catalyst samples were dried at 473 K for 1 h in a ow of He. The catalysts were then reduced at 503 K for 2 h in a ow of 34% H2/He. CO was introduced for 1 s at a pressure of 20 Torr, and adsorbed at 323 K. Temperature-programmed desorption (TPD) of adsorbed CO was carried out in a ow of He (30 mL min 1), using a heating rate of 10 K min 1. 2.3. Catalytic test Activity and selectivity measurements of the ethanol reaction were carried out in a continuous-ow, tubular xed-bed glass reactor (10 mm i.d.), over the temperature range 473548 K, at atmospheric pressure. The measurements, at each temperature, were performed in steady state during 2 h. Samples were reduced in-situ in a ow of pure H2, with heating from room temperature to 503 K at a rate of 10 K min 1, and holding at the nal temperature for 180 min. Reactions were carried out with 0.100 g of samples (Cu/SiO2, Cu2O/SiO2, or Cu/ ZrO2). Ethanol (99.9%) in the gas phase was generated by passing a ow of He (2.4 ml min 1) through a saturator at 328 K, enabling an

ethanol feed partial pressure of 37.42 kPa at W/F =29.1 gcat h mol 1EtOH, where W and F are the catalyst weight (g) and the ethanol ow rate (molEtOH h 1), respectively. The reaction products were periodically collected and analyzed by on-line gas chromatography (Varian GC-3400 CX tted with a Chromasorb 102 column). 3. Results and discussion XRD patterns of the ZrO2 support and calcined Cu/SiO2, Cu/ZrO2, and Cu2O/SiO2 samples are shown in Fig. 1. Intensive XRD pattern peaks were obtained for the Cu/SiO2 sample at 35.5 and 38.7, corresponding to a crystalline monoclinic CuO phase with a tenorite structure (JCPDS 48-1548). For the Cu/ZrO2 sample, peaks at 24.2, 28.2, 31.4, and 34.3 were attributed to a monoclinic ZrO2 phase (JCPDS 37-1484). The peaks at 35.5 and 38.7, corresponding to CuO, were also detected for the Cu/ZrO2 sample. Peaks at 36.4 and 42.3 for the Cu2O/SiO2 sample were assigned to the Cu2O cubic phase (JCPDS 05-0667). Fig. 2 depicts the TPR-H2 proles for the CuO/ZrO2, Cu/SiO2, and Cu2O/SiO2 samples. Cu/ZrO2 exhibited two peaks, at 400 and 480 K. TPR peaks below 450 K are related to the reduction of welldispersed Cu oxide species, indicating a strong CuOZrO2 interaction. The peak at 480 K could be assigned to the reduction of CuO due to weak interaction with the ZrO2 surface or bulk CuO, as conrmed by the XRD analyses. Cu2O/SiO2 exhibited one peak between 625 and 725 K, indicating the reduction of Cu2O crystals. The higher temperature of reduction for Cu2O than CuO is expected from the higher apparent activation energy of 27.4 kcal/mol for Cu2O, compared to a value of 14.5 kcal/mol for CuO [14]. The Cu/SiO2 sample showed overlapping peaks at 500, 550, and 600 K. The species at around 500 K was assigned to bulk CuO, while the species reducible at higher temperatures were indicative of CuOSiO2 interaction. Fig. 3A and B presents the evolution of the XANES spectra at the Cu K-edge, collected during TPR-H2 experiments using Cu/SiO2 and Cu/ ZrO2, respectively. The proles of oxidation states of Cu as a function of temperature were obtained by applying a linear combination of spectra for the CuO, Cu2O, and Cu foil standards. The reduction of copper oxide species supported on SiO2 and ZrO2 initiates at around 450 K. The content of Cu 1 + species remains low in Cu/ZrO2 catalysts, even when they are reduced at high temperatures. Although Cu/SiO2 is totally reduced at high temperatures (up to 650 K, Fig. 3C), a high amount of Cu 1 + remained at temperatures at which the catalysts were reduced (500 K). The XANES results indicated that the reduction of CuO occurred in two steps (Cu 2 + Cu 1 + Cu 0). The equilibrium Cu 1 + Cu 0 in small metallic Cu crystallites therefore depends on the standard free energy of formation, and hence on the morphology of

Fig. 1. X-ray diffraction patterns for calcined Cu/SiO2, ZrO2 support, Cu/ZrO2, and Cu2O/SiO2.

Fig. 2. TPR-H2 proles for calcined Cu/SiO2, Cu/ZrO2, and Cu2O/SiO2 samples.

124

A.G. Sato et al. / Catalysis Communications 26 (2012) 122126

Fig. 3. Time-resolved XANES spectra collected during TPR-H2 for Cu/SiO2 and Cu/ZrO2 (AB), and semi-quantitative evolution of copper species with increased temperature (CD).

the crystallite [15]. The oxidation of the metal is favored by a decrease in both metal particle size and temperature [16,17]. Adsorption of CO on Cu surfaces in different oxidation states has been shown to give rise to various species that exhibit IR absorption bands in the 20002200 cm 1 spectral region [18]. Fig. 4 shows FT-IR spectra of CO adsorbed at 323 K on (A) Cu/SiO2 and (B) Cu/ZrO2 catalysts, where the IR band of adsorbed CO can be seen at 2125 and 2102 cm 1, respectively. The observed shift in the IR band suggests the existence of signicant structural and electronic differences between the Cu/SiO2 and Cu/ZrO2 samples. The adsorption of CO on Cu/SiO2 produces the Cu0\CO band at 2122 cm 1 in samples reduced at high temperatures [19]. On the other hand, Cu 0\CO species usually absorb at a lower frequency than that observed for Cu1 +\CO species. However, when copper is highly dispersed, Cu0 forms carbonyls that can absorb at the same frequency as Cu 1 +\CO carbonyls. In this case, the criterion of stability of the carbonyls has to be applied in order to distinguish Cu0\CO from the Cu+ 1\CO species [20]. Higher stability characterizes a stronger -component of the Cu1 +\CO bond, and low stability characterizes a Cu0\CO -bond [19]. The spectra of adsorbed CO and residual CO after desorption at different temperatures are shown for the Cu/SiO2 and Cu/ZrO2 catalysts by the spectra below the thicker lines in Fig. 4A and B, respectively. For the Cu/ SiO2 catalyst, a low thermal stability CO fraction was desorbed at 373 K, and another CO fraction was desorbed at a higher temperature (423 K). These results suggest that for Cu/SiO2, the CO band at 2125 cm1 corresponds to the overlap of Cu0\CO and Cu1 +\CO species, which have low and high thermal stabilities, respectively. For the Cu/ZrO2 catalyst, most of the adsorbed CO was of low thermal stability, suggesting that CO was mainly adsorbed on Cu 0 sites. The Cu 0\CO band for the Cu/SiO2 and Cu/ZrO2 catalysts was detected

at 2125 and 2102 cm 1, respectively. The shift of around 23 cm 1 in the IR band position, to a lower frequency, suggests stronger back-donation from Cu 0 to CO, characterizing a higher electron density of Cu sites for the Cu/ZrO2 catalyst than for the Cu/SiO2 catalyst. Carbon monoxide can interact with either the Lewis acid Zr + cations or the Brnsted acid hydroxyl groups present at the surface of ZrO2 [21]. The spectra of adsorbed CO in the ZrO2 spectral region (17501150 cm 1) are shown in Fig. 5 for the Cu/ZrO2 catalyst. Analogously to the interaction of CO2 with ZrO2, the formation of carbonate and bicarbonate species could be an indication of surface basicity associated with the presence of oxygen atoms [22] and OH groups. The different surface species formed upon CO adsorption can yield information on the existence of basic surface sites (basic hydroxyl groups and coordinatively unsaturated sites (cus) O 2 centers) or of acidbase pair sites (cus Zr4 +O2 centers) [23]. The CO adsorbed onto the Cu/ZrO2 sample showed well-resolved bands at 1635, 1546, 1493, 1394, 1340, and 1218 cm 1, related to the presence of different bonded CO species caused by the presence of basic centers and acidbase pair sites. The bands at 1635 and 1218 cm 1 were assigned to bidentate bicarbonates (b-HCO3 ), where1 1 as the bands at ca. 1546 and 1340 cm , and at ca. 1394 cm , were re2 2 lated to bidentate (b-CO3 ) and monodentate (m-CO3 ) carbonate species, respectively [23,24]. The band at 1493 cm 1 could be associated with the formation of ionic bicarbonate species (i-HCO3 ). The bicarbonate species are usually formed by adsorption of CO onto basic hydroxyl groups. The formation of OH groups on the monoclinic ZrO2 surface is assumed to be predominantly terminal [21,23]. Carbonate structures are formed via interaction of the oxygen in CO with Zr 4 + cations in the lattice, as well as with a surface oxygen atom [21,23]. The bidentate carbonate

A.G. Sato et al. / Catalysis Communications 26 (2012) 122126

125

Fig. 5. FT-IR spectra of CO adsorbed at 323 K on Cu/ZrO2. Spectral region of zirconia.

Fig. 4. FT-IR spectra of adsorbed CO, and TPD, for (A) Cu/SiO2 and (B) Cu/ZrO2. Spectral region of copper. Thicker line: adsorption at 323 K.

surface complexes involve acidbase pair sites (cus Zr 4 +O 2 centers), while the monodentate carbonates involve strongly basic cus O 2 centers [2527]. The changes in catalytic activity and selectivity for ZrO2, and for Cu supported on SiO2 or ZrO2, are shown in Table 1. The products formed during ethanol conversion on Cu/SiO2 and Cu/ZrO2 were ethyl acetate (AcOEt), acetaldehyde (AcH), methyl ethyl ketone (MEK), crotonaldehyde (CROT), propanone (PrO), CO, CO2, ethylene (ETE), diethyl ether (DEE), and butanol (BOL). The action of Cu/SiO2 was mainly in the dehydrogenation of ethanol, as shown by a selectivity for acetaldehyde (SAcH = 87%). As it has been reported previously [28], an increased loading of Cu on SiO2 (not shown) resulted in increased selectivity for ethyl acetate, with the catalyst containing 35% of Cu loading (35Cu/SiO2) showing a selectivity for ethyl acetate of SAcOEt = 45%. The Cu2O/SiO2 catalyst was highly selective for dehydrogenation of ethanol to acetaldehyde (Table 1, run 7). Dehydrogenation activity on Cu2O can be attributed to adsorption of ethanol to form CH3CH2O\Cu 1 + species, and the ability of Cu 1 +\O sites to undergo CH scission to form acetaldehyde. Therefore, the high activity and selectivity for dehydrogenation of ethanol on Cu/SiO2 was related to a predominance of Cu 1 + sites. Unlike Cu/SiO2, Cu/ZrO2 showed high activity and selectivity for AcOEt, even at a low Cu loading. The Cu/SiO2 and Cu/ZrO2 catalysts showed activities and selectivities for ethyl acetate of rAcOEt = 0.34 mmol/gcat h, SAcOEt = 5.3%, and rAcOEt = 5.12 mmol/gcat h,

SAcOEt = 73.0%, respectively. Nevertheless, Cu/SiO2 and Cu/ZrO2 showed low selectivities for products of AcH condensation (such as MEK, BOL, and CROT). The products of ethanol dehydration were DEE and ETE. In an attempt to understand the nature of the active sites involved in formation of AcOEt, the ethanol was placed in contact with the ZrO2 support. The results in Table 1 show that for ZrO2 there was mainly dehydration of ethanol to DEE, while the activity for ethanol dehydrogenation was much lower for ZrO2 than for Cu/ZrO2. This nding is in agreement with DFT calculations [29], and indicates the low effectiveness of ZrO2 for dehydrogenation of ethanol, due to a high barrier to CH scission. As a result, adsorbed CH3CH2OZr(a) ethoxy species stable are formed on ZrO2. Since ZrO2 was unable to dehydrogenate ethanol, and was consequently not active for AcOEt involving AcH as a reactant, the ethanol was passed through a double bed reactor (two catalytic beds in series, separated with quartz wool), the rst bed containing Cu/SiO2, and the second containing ZrO2. Although the formation rate of AcOEt increased from rAcOEt = 0.34 mmol/gcat h, for Cu/SiO2, to rAcOEt = 0.70 mmol/gcat h, for Cu/SiO2//ZrO2 (run 4), the formation rate was still lower than the rAcOEt = 5.12 mmol/gcat h obtained for the Cu/ ZrO2 catalyst. This result suggests that although AcOEt can be formed from ethanol and AcH on ZrO2, the sites highly active for AcOEt are formed by the presence of Cu species on ZrO2. Interestingly, comparison of runs 4 and 5 revealed that there was a higher selectivity for products of AcH condensation (CROT, MEK, and BOL) when the Cu/SiO2 catalyst was physically mixed with ZrO2. This suggests that intermediates of AcH formed due to the presence of Cu 1 + species at the Cu 0 surface can be transferred onto the ZrO2 surface by spillover, promoting formation of the condensation products of acetaldehyde [13]. The results of runs 2, 3, and 4 showed that there was suppression of AcH condensation products (CROT, MEK, and BOL) when Cu was impregnated onto ZrO2, suggesting that the Cu particles could have interacted with acid sites of the ZrO2 surface, preventing the condensation by decreasing the density of acid sites. Reaction of ethanol on the ZrO2 support formed products of the dehydration of ethanol (DEE and ETE). When ethanol and acetaldehyde in a ratio of 1.65 (EtOH:AcH) were passed over ZrO2, the AcH condensation products (CROT, MEK, and BOL) were formed. This behavior suggests that strongly acidic sites were responsible for ethanol dehydration [30], while dehydrogenation required moderately acid sites and strongly basic sites. On the other hand, acid sites can promote the -aldolization reaction of acetaldehyde [31]. The DRIFTS analyses showed the presence of Cu 1 + species with lower electron density over Cu/SiO2, which should favor adsorption of ethoxy species, while for Cu/ZrO2 the presence of a higher electron

126

A.G. Sato et al. / Catalysis Communications 26 (2012) 122126

Table 1 Product distribution in the ethanol reaction over Cu/ZrO2, Cu/SiO2 and Cu2O/SiO2. Run T (K) 498 473 498 498 498 473 523 Reactant feed EtOH EtOH EtOH EtOH EtOH EtOH + AcHc EtOH Catalyst Ethanol conversion (%) 41 45 0.7 41 47 43 15 Selectivity (%) AcH 87 23 30 66 51 32 100 AcOEt 5.3 73 11 11 54 MEK/BOL 6.3 2.0 7.8 18 5.3 CROT 0.7 1.3 14 19 7.4 DEE 0.4 0.4 46 0.5 0.6 0.2 ETE 0.3 17 0.2 0.1 0.1 Othersa 0.3 7.0 0.5 0.3 1.0 AcOEt formation rate (mmol/gcat h) 0.34 5.12 0.70 0.81 3.62

1 2 3 4 5 6 7

Cu/SiO2 Cu/ZrO2 ZrO2 (Cu/SiO2//ZrO2)b Cu/SiO2 + ZrO2 mixed Cu/ZrO2 Cu2O/SiO2

Selectivity (%) to: Acetaldehyde (AcH); ethyl acetate (AcOEt); methyl-ethyl-ketone and/or butanol (MEK/BOL); crotonaldehyde (CROT); diethyl ether (DEE); and ethylene (ETE). a Others (minor accounts): propanone, CO and CO2. b Double bed reactor: Cu/SiO2 (rst layer = 100 mg) and ZrO2 (second layer = 90 mg). c Ratio of EtOH:AcH = 1.65.

density Cu 0 species might favor the adsorption of acyl species. The results suggested that ethyl acetate formation occurred due to the combination of the Cu 1 +/Cu 0 pair and Cu 0\ZrO2 interface sites, which promoted the reaction between ethoxy species, mainly adsorbed over Cu 1 + and Zr + cations, and acyl species, mainly adsorbed over Cu 0. A detailed analysis of this effect will be the subject of a future paper. 4. Conclusions In the Cu/SiO2 catalyst, the Cu 1 + species was more prevalent than Cu at the surface and in the bulk material, in contrast to Cu/ZrO2, where the main species was Cu 0. Accordingly, the Cu 1 + provided by the Cu2O/SiO2 catalyst was highly selective for the dehydrogenation of ethanol, and the Cu 1 +/Cu 0 pair provided by Cu/SiO2 generated sites of very low selectivity for ethyl acetate and high selectivity for acetaldehyde. The efcient direct transformation of ethanol into ethyl acetate requires a combination of the Cu 1 +/Cu 0 pair and high electron density Cu 0 chemically interfaced to ZrO2.
0

Acknowledgments The authors thank CNPq and CAPES for nancial support. The XPD facilities were provided by the National Synchrotron Light Laboratory (LNLS), Brazil (beam line D06A-DXAS). References
[1] A.M. Ruppert, K. Weinberg, R. Palkovits, Angewandte Chemie. International Edition 51 (2012) 25642601. [2] J. Sun, K. Zhu, F. Gao, C. Wang, J. Liu, C.H.F. Peden, Y. Wang, Journal of the American Chemical Society 133 (2011) 1109611099. [3] T. Tsuchida, J. Kubo, T. Yoshioka, S. Sakuma, T. Takeguchi, W. Ueda, Journal of Catalysis 259 (2008) 183189. [4] D.P. Volanti, A.G. Sato, M.O. Orlandi, J.M.C. Bueno, E. Longo, J. Andrs, ChemCatChem 3 (2011) 839843. [5] F.-W. Chang, W.-Y. Kuo, K.-C. Lee, Applied Catalysis A: General 246 (2003) 253264.

[6] N. Iwasa, N. Takezawa, Bulletin of the Chemical Society of Japan 64 (1991) 26192623. [7] K. Inui, T. Kurabayashi, S. Sato, Journal of Catalysis 212 (2002) 207215. [8] K. Inui, T. Kurabayashi, S. Sato, Applied Catalysis A: General 237 (2002) 5361. [9] K. Inui, T. Kurabayashi, S. Sato, N. Ichikawa, Journal of Molecular Catalysis A: Chemical 216 (2004) 147156. [10] L. Wang, W. Zhu, D. Zheng, X. Yu, J. Cui, M. Jia, W. Zhang, Z. Wang, Reaction Kinetics, Mechanisms, and Catalysis 101 (2010) 365375. [11] A.B. Gaspar, A.M.L. Esteves, F.M.T. Mendes, F.G. Barbosa, L.G. Appel, Applied Catalysis A: General 363 (2009) 109114. [12] A.B. Gaspar, F.G. Barbosa, S. Letichevsky, L.G. Appel, Applied Catalysis A: General 380 (2010) 113117. [13] P.C. Zonetti, J. Celnik, S. Letichevsky, A.B. Gaspar, L.G. Appel, Journal of Molecular Catalysis A: Chemical 334 (2010) 2934. [14] J.Y. Kim, J.A. Rodriguez, J.C. Hanson, A.I. Frenkel, P.L. Lee, Journal of the American Chemical Society 125 (2003) 1068410692. [15] E. van Steen, M. Claeys, M.E. Dry, J. van de Loosdrecht, E.L. Viljoen, J.L. Visagie, The Journal of Physical Chemistry. B 109 (2005) 35753577. [16] A.M. Saib, A. Borgna, J. van de Loosdrecht, P.J. van Berge, J.W. Geus, J.W. Niemantsverdriet, Journal of Catalysis 239 (2006) 326339. [17] R. Knapp, S.A. Wyrzgol, A. Jentys, J.A. Lercher, Journal of Catalysis 276 (2010) 280291. [18] A. Dandekar, M.A. Vannice, Journal of Catalysis 178 (1998) 621639. [19] K. Hadjiivanov, H. Knozinger, Physical Chemistry Chemical Physics 3 (2001) 11321137. [20] A. Milushev, K. Hadjiivanov, Physical Chemistry Chemical Physics 3 (2001) 53375341. [21] K.T. Jung, A.T. Bell, Journal of Molecular Catalysis A: Chemical 163 (2000) 2742. [22] W. Hertl, Langmuir 5 (1989) 96100. [23] D. Bianchi, T. Chak, M. Khalfallah, S.J. Teichner, Applied Catalysis A: General 112 (1994) 5773. [24] K. Pokrovski, K.T. Jung, A.T. Bell, Langmuir 17 (2001) 42974303. [25] G. Cerrato, S. Bordiga, S. Barbera, C. Morterra, Surface Science 377379 (1997) 5055. [26] B. Bachiller-Baeza, I. Rodriguez-Ramos, A. Guerrero-Ruiz, Langmuir 14 (1998) 35563564. [27] J.I.D. Cosimo, V.K. Dez, M. Xu, E. Iglesia, C.R. Apestegua, Journal of Catalysis 178 (1998) 499510. [28] R.M. Duran, Thesis: Desidroacoplamento do etanol para acetato de etila sobre catalisadores de cobre suportados. Departamento de Engenharia Qumica, Universidade Federal de So CarlosUFSCar, So Carlos, 2000, pp. 135. [29] Y.-W. Chen, J.-J. Ho, Journal of Physical Chemistry C 113 (2009) 61326139. [30] M.E. Manrquez, T. Lpez, R. Gmez, J. Navarrete, Journal of Molecular Catalysis A: Chemical 220 (2004) 229237. [31] A. Yee, S.J. Morrison, H. Idriss, Catalysis Today 63 (2000) 327335.

Vous aimerez peut-être aussi