Sie sind auf Seite 1von 7

Applied Catalysis A: General 268 (2004) 2531

In situ preparation of zirconium sulfate pillared clay: study of acidic properties


S. Ben Chaabene a , L. Bergaoui a, , A. Ghorbel a , J.F. Lambert b , P. Grange c
a

Laboratoire de chimie des matriaux et catalyse, Facult des sciences de Tunis, Universit Tunis-El Manar, Campus universitaire, 1060 Le Belvdre, Tunis, Tunisia b Laboratoire de ractivit de surface, UMR 7609 CNRS, Universit Pierre et Marie Curie, 4 place Jussieu, 75252 Paris Cedex 05, France c Unit de catalyse et chimie des matriaux diviss, Universit Catholique de Louvain, Croix du Sud2/17, B-1348 Louvain-la-Neuve, Belgium Received 14 March 2003; received in revised form 14 March 2003; accepted 10 March 2004

Abstract Sulfated zirconium clays were prepared by adding ammonium sulfate to the intercalation solution. The main controlled parameter in this study is the SO4 :Zr ratio in solution. The characterization of these catalysts, their acidic properties and their catalytic activities was examined by N2 -BET adsorption, chemical analysis, the adsorptiondesorption of basic molecules, the conversion of n-hexane and the transformation of isopropanol. Two different types of SO4 Zr polycation binding were advanced and correlated with the acidic properties of these solids. To enhance the acidity of the sulfated zirconium clay, the SO4 :Zr molar ratio must be up than 0.125. The higher this ratio, the higher is the activity of these solids. It appears that highly polymerized entities are more active than individual sulfated polycations. The isopropanol dehydration to propene as well as the isomerization of n-hexane seems to be related to the number and the strength of Brnsted acidity. 2004 Published by Elsevier B.V.
Keywords: Sulfated Zr-pillared clay; Acidity (TPD and IR); n-Hexane conversion; Isopropanol dehydration

1. Introduction In order to protect the environment, several sets of regulations have been established. Owing to this legislation, great interest has been devoted to the substitution of unfriendly and corrosive liquids (used in chemical and petrochemical industries for important reactions) by solid catalysts. On this basis, clays may constitute very promising substitutes. When inorganic species are introduced into the interlayers of the clay, the resulting nanocomposite can be used as a catalyst for specic reactions. For instance, pillared interlayered clays (PILCs) including metal clusters such as (Zr, Al and Ti) generate micropores [15] larger than those of zeolites [6,7]. Besides, clay constitutes a good and cheap
Corresponding author. Present address: INSAT, Cimie et Biologie Appliques, Centre Urbain Nord, B.P. 676, Tunis 1080, Tunisia. Tel.: +216-71-703-829; fax: +216-71-704-329. E-mail address: latifa.bergaoui@insat.rnu.tn (L. Bergaoui).

support to obtain a good dispersion of metal species. The dispersion of sulfated metal oxide such as zirconium oxide, which is known to have a very strong acidity [810], and the elucidation of their acidic properties is the aim of this work. Synthesis of this type of materials has been the subject of few studies [1116]. Sulfated Zr-PILCs have been prepared and impregnated with sulfates [1114]. Nevertheless, this method of sulfation leads to solids having low surface areas and low thermal stability of sulfates and developing moderate acidity. Farfan-Torres and Grange modied the acidity of the ZrOCl2 -montmorillonite by adding (NH4 )2 SO4 during the intercalation reaction [11,13]. The incorporation of zirconium sulfate hydroxyl complex in Na-montmorillonite using zirconium acetate as a precursor was also studied [15]. In this work, a new preparation method is investigated, including an initial sulfation by creation of sulfated species in the ZrOCl2 intercalation solution.

0926-860X/$ see front matter 2004 Published by Elsevier B.V. doi:10.1016/j.apcata.2004.03.015

26

S. Ben Chaabene et al. / Applied Catalysis A: General 268 (2004) 2531

2. Experimental 2.1. Preparation method The preparation of sulfated zirconium pillared clay is an attempt to intercalate zirconium sulfate species between the clay layers. A sample of montmorillonite (KC2) was obtained from the CECA (France) and its <2 m fraction was separated by gravity sedimentation. It was then exchanged three times with a 1 mol/l NaCl solution and washed thoroughly. Chemical analysis yielded an exchange capacity of 80 meq/100 g. Ammonium sulfate was added to freshly prepared 0.1 mol/l ZrOCl2 solutions, with SO4 :Zr molar ratios varying from 0 to 0.2. The solution was then reuxed during 4 h. Intercalated clays were then prepared by adding the intercalation solution dropwise to 10 g/l clay suspensions. The slurry was stirred for 4 h at boiling temperature, washed by successive dialyses, and dried during one night at 120 C. The samples were then calcined in owing air. The temperature was raised at 60 C/h to 500 C, and the nal temperature was maintained for 5 h. The term of pillared clays will here be reserved to samples stabilized by calcination. The pillared montmorillonite will be called ZrP-r (where r is the SO4 :Zr molar ratio in the intercalation solution). 2.2. Textural properties Surface area measurements were performed by nitrogen physisorption at 77 K using a static volumetric apparatus (Micromeritic ASAP 2000 adsorption analyzer); the BET equation was applied to the adsorption isotherm. All samples were evacuated in vacuum at 110 C prior to nitrogen physisorption. 2.3. Chemical analysis The total zirconium was analyzed by atomic adsorption after dissolution of the solids. Sulfur was analyzed by using a Coulomat 702 apparatus. Sulfur was thermally decomposed at high temperature, carried away by an oxygen ow and dissolved in solution. The amount of dissolved sulfur was measured by a specic electrode. The analysis of the surface for zirconium, silicon and sulfur was performed in an SSX 100/206 photoelectron spectrometer. The data treatment was performed with the CasaXPS program. 2.4. Temperature-programmed desorption of ammonia The total acidity of the sulfated zirconium pillared clays was determined by temperature-programmed desorption of ammonia (NH3 -TPD). Before adsorption of ammonia, the samples were treated under oxygen at 500 C. The samples were then cooled down to 80 C in He ow, then treated with NH3 pulses until sat-

uration. Weakly adsorbed NH3 was eliminated by treatment under He at the same temperature for 1 h. The temperature-programmed desorption of ammonia was performed between 80 and 500 C at 10 C/min and followed by an on-line catharometer. 2.5. Pyridine thermodesorption Surface acidity was determined using probe molecules adsorption (pyridine). The calcined sample was evacuated at 500 C for 2 h, then cooled and contacted for 15 min with pyridine at room temperature. After this step, the sample was outgassed for 1 h, and then heated to the desired temperature using a linear temperature programmation. IR spectra were recorded at each stage on a Bruker IFS 66V spectrometer. 2.6. n-Hexane conversion test The n-hexane isomerization was performed in a microow reactor under atmospheric pressure, using a mechanical mixture (1 w/w) of catalyst and a standard Pt/Al2 O3 reforming catalyst (0.35% Pt in weight). The latter was previously reduced at 500 C for 4 h. The catalytic test was performed as follows: the mixture of Pt/Al2 O3 + catalyst was pretreated rst under O2 ow at 500 C for 1.5 h and then under H2 ow for 30 min at 220 C. H2 saturated with n-hexane was passed over the catalyst. The analysis of the products was carried out on-line using FID gas chromatography. At the reaction temperature (280 C) Pt/Al2 O3 alone is not active. 2.7. Isopropanol conversion test The relative acidity of the catalysts was also evaluated using the 2-propanol transformation reaction. In a xed bed reactor, 100 mg of catalyst were treated under O2 ow during 3 h at 400 C and with a temperature ramp of 10 C/min. The sample was then cooled under He down to 100 C. Isopropanol-saturated helium was passed over the catalyst and the reaction was followed at 100 C using on-line chromatography.

3. Results and discussion 3.1. Composition and textural properties of the catalysts The amount of zirconium xed by ZrP0 is 18.5 wt.% or 25 wt.% of ZrO2 (Fig. 1). Former studies of Zr-pillared clays have reported very different amounts of intercalated Zr, depending on the host clay and the pillaring method used. In the pioneer work of Yamanaka and Brindley, where the intercalation reaction was carried out at room temperature, the amount of zirconium xed was about 13.9 wt.% of ZrO2 [17]. The higher amount of xed zirconium in our work is probably do to the higher temperature used for the intercalation reaction. An amount of xed zirconium closer to our

S. Ben Chaabene et al. / Applied Catalysis A: General 268 (2004) 2531


Total zirconium content (wt %) 45 40 35 30 25 20 15 0 0.05 0.1 0.15 SO4:Zrsolution 0.2

27

Total sulfur content (wt %)

1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0.05 0.1 0.15 SO4:Zr solution 0.2

Fig. 1. Fixed zirconium (wt.%) vs. SO4 :Zr ratio in the intercalation solution.

Fig. 2. Fixed sulfur (wt.%) vs. SO4 :Zr ratio in the intercalation solution.

value has been reported where the suspension was stirred at 40 C during the intercalation [18,19]. In fact, it has been reported by Ohtsuka et al. that at room temperature the intercalated species in a synthetic clay is the zirconium tetramer [Zr4 (OH)8 (H2 O)16 ]8+ , giving a 7 interlayer spacing. More highly polymerized zirconium species are generated after an elevated temperature aging step [20]. Singhal et al. have suggested the existence in a ZrOCl2 solution of an equilibrium between tetrameric and octameric species [21]. In our case, we postulate that the intercalated species are octamers (Zr8 ) because an interlayer spacing higher than 7 is obtained. The speciation in solution will be discussed elsewhere in more detail. The textural and chemical properties of the catalysts are reported in Table 1. All the solids have a BET surface area between 200 and 273 m2 /g and a microporous volume between 0.03 and 0.05 cm3 /g. For r values higher than 0.125 a decrease in the surface is noticed when r increases. The zirconium and the sulfur contents (Figs. 1 and 2, respectively) remain constant when the sulfated zirconia pillared clays are prepared with SO4 :Zr ratios ranging from 0.05 to 0.125. For SO4 :Zr ratios lower then 0.125, less then one sulfate ion per Zr8 unit are present. This could be interpreted by equilibrium, in the intercalation solution, between sulfate-free Zr8 polycations and polycations with at least one sulfate attached. Since the sulfate-free polycations are more charged than the sulfated polycations and cation-exchange is expected to favor high-charge species, the amount of reTable 1 Surface area and chemical analysis for different ZrPr solids SO4 :Zr molar ratio in solution 0 0.025 0.05 0.075 0.1 0.125 0.15 0.175 0.2 Surface area (m2 /g) 228 222 273 245 246 267 242 223 199 Si atomic % at the surface 19.8 18.8 16.7 15.8 17.3 15.8 15.6 16.7 10.9 S/Zr atomic ratio of the bulk 0 0.04 0.04 0.03 0.04 0.05 0.06 0.08 0.08 S/Zr atomic ratio at the surface 0 0.05 0.09 0.12 0.15 0.13 0.57 0.65 1.03

tained zirconium and sulfur can remain almost constant even if the concentration of sulfated species increases. For r > 0.125, the sulfur and zirconium amounts increase rapidly which suggest the presence of a more polymerized species. Table 1 also shows that the surface S/Zr ratios are higher than the bulk S/Zr ratios. We have to precise here that the surface region probed by XPS corresponds mainly to the surface of tacto ds consisting in the stacking of several clay layers. We have previously demonstrated for aluminum pillared clay, that when the intercalation solution is added to a clay suspension, the solid occulates quickly and the intercalation process takes place during the dialysis step [22]. The same phenomenon may be operating for sulfated zirconium-intercalated montmorillonite. Since the unsulfated polycations are more charged and less voluminous than other, they have probably higher mobility during the dialysis process, and can migrate easier between clay layers. Therefore sulfated polycations should be xed preferentially at the surface of tacto ds, while unsulfated polycations are intercalated between clay layers. For r = 0.2, the low Si content detected at the surface, the high total amounts of zirconium and sulfur and the low surface area suggest that a bulk Zr-containing polymeric phase has been precipitated. This phase is highly sulfated since the S/Zr atomic ratio at the surface is equal to 1.03. In summary, we can say that for r lower than 0.125 the clay contains both unsulfated and sulfated polycationic species. This sulfated polycations will be called type I species (Fig. 3). When r is higher than 0.125, the amount of SO4 2 in solution is sufcient to form bridges between the zirconium polycations and develops a polymeric phase (type II in Fig. 3). For the highest amounts of SO4 2 , a highly polymeric phase is observed. The presence of two modes of interaction between sulfate ions and zirconium polycations was conrmed elsewhere [16]. 3.2. TPD of NH3 The data displayed in Table 2 show that the bulk concentration of acidic sites is more important than for sulfated zirconia [23]. Since the zirconium in pillared clay is more

28

S. Ben Chaabene et al. / Applied Catalysis A: General 268 (2004) 2531


0.05<r<0.125 0.125<r<0.2 Sulfate ion

Unsulfated polycation Type I sulfated polycation Type II sulfated polycations

Fig. 3. Presentation of different zirconium species in interaction with the clay layers.

Transmittance

Table 2 Concentration of ammonia desorbing between 80 and 500 C for different ZrPr samples SO4 :Zr ratio in solution 0 0.075 0.125 0.15 0.175 0.2 [NH3 ] (mmol/g) 0.713 0.803 0.981 0.866 0.848 1.057

f e d c b B a L+B L

1600

1550

dispersed, this result is not surprising. The amounts of adsorbed NH3 obtained for our solids are very close to those reported in a former study of zirconium-pillared clay sulfated by impregnation [13]. The presence of sulfate in the intercalating solution obviously enhances the concentration of acid sites. It is not easy to correlate this concentration with any single parameter because it depends on the amount of both zirconium and sulfur between the layers and at the surface of tacto ds, and also on the surface area of the solids. If we plot the surface concentration of acid sites (in mol/m2 of solid) and the temperature of maximum desorption versus the SO4 :Zr ratio (Fig. 4), we can notice that, as long as there is no formation of a species type II (i.e., for 0.05 < r < 0.125), the amount of adsorbed ammonia is relatively constant and the energy of interaction with the surface does not vary either. We can hypothesize that the rst type of bridging sulfate has not any evident effect on the number and the strength of the acidic sites. How6 mol NH 3 /m 5 4 3 2 1 0 0.05 0.1 0.15 0.2 120 100 80 SO42-:Zr4+ Solution
2

1500 1450 1400 Weavenumber (cm-1)

Fig. 5. IR spectra of ZrP0 sample after pyridine adsorption followed by desorption at room temperature (a), 150 C (b), 225 C (c), 300 C (d), 375 C (e) and 450 C (f). B: Brnsted acidity; L: Lewis acidity.

ever, when the polymeric phase is present (r > 0.125), the amount and the strength of the acid sites abruptly increase. 3.3. Adsorptiondesorption of pyridine Pyridine has been used as a probe molecule for the determination of the nature of acid sites present on the surface of catalysts. Figs. 57 present the evolution of the transmittance IR spectrum in the 16001400 cm1 range upon pyridine adsorption followed by desorption at increasing temperature, respectively for the ZrP0, ZrP0.125 and ZrP0.2 samples.
f e d c b a B L+B

160 140

Temperature of maximum desorption (C)

180

Transmittance

L 1600 1550 1500 1450 1400 Wavenumber (cm-1 )

Fig. 4. The amount of retained ammonia and the temperature of maximum desorption for different ZrPr.

Fig. 6. IR spectra of ZrP0.125 sample after pyridine adsorption followed by desorption at room temperature (a), 150 C (b), 225 C (c), 300 C (d), 375 C (e) and 450 C (f). B: Brnsted acidity; L: Lewis acidity.

S. Ben Chaabene et al. / Applied Catalysis A: General 268 (2004) 2531


f e d c b
Zr Zr O O Zr Zr Zr S O Zr O Zr Zr H O O Zr Zr O H O S O O O H H Zr O H O Zr H O O Zr Zr O H O S O Zr O H H O

29

Transmittance

Zr O H

(a)
B L+B L

(b)

(c)

Fig. 8. Schematic view of the rst type of SO4 Zr polycations binding (a) and the second type of SO4 Zr binding (b, c).

1600

1550

1500 1450 1400 Wavenumber (cm-1)

Fig. 7. IR spectra of ZrP0.2 sample after pyridine adsorption followed by desorption at room temperature (a), 150 C (b), 225 C (c), 300 C (d), 375 C (e) and 450 C (f). B: Brnsted acidity; L: Lewis acidity.

The spectrum recorded after pyridine desorption at 150 C shows the existence of both Lewis and Brnsted acidity. The decrease of the intensity of the 1448 cm1 band after evacuation at 300 C for ZrP0 and ZrP0.125 indicates that the Lewis acid sites are not very strong. For ZrP0.2 the 1448 cm1 band is still observed after desorption at 375 C which indicates a higher Lewis acidity for this solid. The Brnsted acidity of ZrP0.125 is lower than that of ZrP0. This fact suggests that SO4 2 substitutes groups which are responsible for the Brnsted acidity in the sulfate-free polycations. In samples where we have postulated a polymeric phase (ZrP0.2), a new type of Brnsted acidity appears, probably as a result of a different type of sulfate/polycation binding. It is believed that dissolution of ZrOCl2 8H2 O in water give rise to polycations having a structure built up from [Zr4 (OH)8 (H2 O)16 ]8+ tetramers [24]. Indeed, Mieh-Brendl and al. have shown by EXAFS that the nearly square frame of Zr4 zirconyl units is preserved after intercalation between clay layers and calcination at moderate temperatures [25]. Using NMR to study the tetranuclear hydroxo zirconium complex in aqueous solution, Aberg and Glaser have proposed that there are two inert and two labile water molecules per Zr. In addition two exchanging protons per Zr were observed and assigned to a terminal water molecule in the tetramer [26]. They also suggest that octamer can form by stacking two tetramers on top of each other and substituting some labile water molecules to give hydroxy bridges. But some labile waters are still present in the octamer [26]. The release of the protons of labile water molecules makes the tetramer a strong acid. When the polycation is intercalated, the corresponding ZrOHterminal groups are probably responsible for the Brnsted acidity observed for ZrP0 solid. Now, in our experimental conditions, when a low amount of SO4 2 is added to the intercalation solution, these labile groups are probably substituted by sulfate groups. These sulfate groups cannot give rise to a Brnsted acidity because all sulfate oxygens are probably coordinated to Zr atoms (Fig. 8a). When the amount of sulfate becomes higher, a new

source of Brnsted acidity appears. The type II of SO4 Zr polycation binding provides most likely this new Brnsted acidity. In this second type of binding, two oxygen atoms may be linked to zirconium, while the other two are free. The SO4 groups may be bridging between zirconium atoms of two different polycations which explain the formation of a high polymeric phase. The latter type of linkage, is probably locally similar to surface groups of bulk SO4 ZrO2 catalysts. An important number of studies have attempted to determine the nature of acid sites in these catalysts [8]. The Brnsted acidity can come from an OH groups linked to zirconium (Fig. 8b). The donation ability of the hydroxyl group is strengthened by the electron-inductive effect of the S=O double bonds in the sulfate group [27,28]. Alternatively, Brnsted acidity has been proposed to arise directly from the SOH group (Fig. 8c) in Clearelds model [29]. 3.4. n-Hexane isomerization The catalytic behavior of the sulfated zirconium pillared clays in the n-hexane isomerization reaction has been studied in order to evaluate the acidity of the catalysts. The evolution of n-hexane conversion as a function of sulfur amount is plotted in Fig. 9. A good correlation is observed between the amount of xed sulfur and the catalytic activity of the resulting solids. When SO4 2 groups are the rst type (0 < r < 0.125), the conversion is very low. On the other hand, the higher the amount of sulfate type II, the higher is the conversion. The isomerization products selectivity is not correlated with the total amount of sulfate retained by the clay, nor with the S/Zr total ratio or the S/Zr surface ratio. In contrast,

20 n-hexane conversion (%) 15 10 5 0

r = 0.2 r = 0.175 r = 0.15 r = 0.125 0 < r < 0.125 0.6 0.8 1 1.2 1.4 1.6 1.8 Total sulfur content (% wt)

Fig. 9. n-Hexane conversion as function of the total amount of xed sulfur.

30
Isomerization Selectivity (%) 70 60 50 40 30 20 10

S. Ben Chaabene et al. / Applied Catalysis A: General 268 (2004) 2531


8 Propene activity (10-8 .mol.g-1 .s -1 ) 7 6 5 4 3 2 1 0 1.6 S/Zr atom. ratiosurface 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 0.05 0.1 0.15 SO4:Zr Solution 0.2

0 1 2 3 4 5 Sulfur contentsurface (atom. %)

Fig. 10. Isomerization selectivity in n-hexane transformation reactions on sulfated zirconium clays as function of the amount of xed sulfur at the surface.

Fig. 11. Evolution of the propene activity and the total S/Zr ratio at the surface for different SO4 :Zr ratios in solution.

it does appear to be positively correlated with the amount of sulfur at the surface, as illustrated in Fig. 10. We can conclude that the isomerization reactions need the Brnsted acidity generated by the second type of sulfur. Only the ZrP0.2 solids can be considered to exhibit a strong acid behavior because it has an appreciable isomerization activity at low temperature. In other words, it appears that only the polymeric phase (possessing the second type of SO4 2 Zr polycation bonding) is able to catalyze the isomerization of n-hexane efciently. The activity of these pillared clay catalysts remains substantially lower than that of bulk SO4 ZrO2 solids [30,31]. 3.5. Isopropanol conversion test The transformation of isopropanol is a widely used reaction to test the acidbase and redox properties of catalysts. The reaction can give diisopropyl ether, propene and acetone as product. Acetone is formed in the presence of basic or redox sites via oxidative dehydrogenation. Ether formation must involve an inner-molecular coupling reaction. Many contradictory interpretations concerning the mechanism of isopropanol dehydration have been presented and isopropanol transformation cannot be a simple test of acidity [32]. In our experimental conditions only propene and ether are produced at a temperature reaction of 100 C. As expected, our solids do not have any basic or redox properties. In Fig. 11 the propene activity is plotted in regard with S/Zr surface ratio. A parallel evolution between these two parameters is observed. The propene activity seems to depend on the strength of acid sites at the surface. To conrm this observation, the activity of propene is also plotted for some catalysts versus the isomerization products activity in Fig. 12. The good correlation between the propene and the isomerization products activities shows that the two reactions need the same active sites. Since it is assumed that the isomerization reactions need Brnsted sites, the same sites are involved in propene formation.

6 Propene activity (10-8 .mol.g -1.s -1 ) 5 4 3 2 1 0 0 5 10 15 20 Isomerization products activity (10-5 .mol.g -1 .s -1)

Fig. 12. Evolution of the propene activity as a function of the isomerization products activity for different ZrPr catalysts.

4. Conclusion The total number of acid sites as measured using a small sized probe molecule (NH3 ) is very important in sulfated Zr-clay catalysts. Those sites are present in both sulfated and unsulfated Zr oxyhydroxide species. Unsulfated Zr species are mostly located between the layers while sulfated species are situated on the surface of tacto ds. Two types of SO4 Zr polycations binding have been evidenced. The rst one (type I) does not generate any new Brnsted acidity because all oxygen atoms of SO4 2 are linked to zirconium ions. This type of binding is predominant for r < 0.125 where less then one sulfate ion per Zr8 unit are present. The second type of SO4 Zr polycation binding (type II) generates both Brnsted and Lewis acidity; it probably corresponds to a sulfate group bridging two zirconium atoms and still possessing free SO groups. Type I species are not very active for n-hexane and isopropanol transformation. The n-hexane isomerization occurs principally at the surface of tacto ds. The dehydration of isopropanol seems to depend on the number and strength of Brnsted acidic sites.

References
[1] D.E.W. Vaughan, Catal. Today 2 (1988) 187. [2] F. Figueras, Catal. Rev.-Sci. Eng. 30 (1988) 457.

S. Ben Chaabene et al. / Applied Catalysis A: General 268 (2004) 2531 [3] I.V. Michell, Pillared Layered Structure: Current Trends and Applications, Elsevier, London, 1990. [4] E.M. Farfan-Torres, P. Grange, J. Chim. Phys. 87 (1990) 1547. [5] J.-F. Lambert, G. Poncelet, Top. Catal. 4 (1997) 43. [6] S. Cheng, Catal. Today 49 (1999) 303. [7] H.J. Chae, I.S. Nam, S.W. Ham, S.B. Hong, Catal. Today 68 (2001) 31. [8] X. Song, A. Sayari, Catal. Rev.-Sci. Eng. 38 (1996) 329. [9] A. Corma, H. Garcia, Catal. Today 38 (1997) 257. [10] G.D. Yadave, J.J. Nair, Micropor. Mesopor. Mater. 33 (1999) 1. [11] E.M. Farfan-Torres, P. Grange, C.R. Acad. Sci. Belg. 45 (1990) 113. [12] E.M. Farfan-Torres, P. Grange, Catal. Sci. Technol. 1 (1991) 103. [13] E.M. Farfan-Torres, E. Sham, P. Grange, Catal. Today 15 (1992) 515. [14] M. Katoh, H. Fujisawa, T. Yamaguchi, Stud. Surf. Sci. Catal. 90 (1994) 263. [15] S. Ben Chaabene, L. Bergaoui, A. Ghorbel, J.-F. Lambert, Stud. Surf. Sci. Catal. 143 (2002) 1053. [16] L. Bergaoui, A. Ghorbel, J.-F. Lambert, Stud. Surf. Sci. Catal. 142 (2002) 903. [17] S. Yamanaka, G.W. Brindley, Clays Clay Miner. 27 (1979) 119. [18] E.M. Farfan-Torres, P. Grange, in: Proceedings of the International Symposium on Chemistry of Microporous Crystals, Tokyo, 1991, p. 97.

31

[19] E.M. Farfan-Torres, O. Dedeycker, P. Grange, in: G. Poncelet, P.A. Jacobs, P. Grange, B. Delmon (Eds.), Preparation of Catalysis V, Elsevier Science, Amsterdam, 1991, p. 337. [20] K. Ohtsuka, Y. Hayashi, M. Suda, Chem. Mater. 5 (1993) 1823. [21] A. Singhal, L.M. Toth, J.S. Lin, K. Affholter, J. Am. Chem. Soc. 118 (1996) 11529. [22] L. Bergaoui, J.-F. Lambert, R. Franck, H. Suquet, J. Chem. Soc., Faraday Trans. 91 (1995) 2229. [23] A. Corma, V. Forns, M.I. Juan-Rajedell, J.M. Lopez Nietro, Appl. Catal. A: General 116 (1994) 151. [24] A. Cleareld, P. Vaughan, Acta Crystallogr. 9 (1956) 555. [25] J. Mieh-Brendl, L. Khouchaf, J. Baron, R. Le Dred, M.-H. Tuilier, Micropor. Mater. 11 (1997) 171. [26] M. Aberg, J. Glaser, J. Inorg. Chem. Acta 206 (1993) 53. [27] K. Arata, Adv. Catal. 37 (1990) 165. [28] D.A. Ward, E.I. Ko, J. Catal. 150 (1994) 18. [29] A. Cleareld, G.P.D. Serrette, A.H. Khazi-Syed, Catal. Today 295 (1994) 295. [30] L. Ben Hammouda, A. Ghorbel, F. Figueras, Stud. Surf. Sci. Catal. 130 (2000) 971. [31] R. Akkari, A. Ghorbel, Stud. Surf. Sci. Catal. 143 (2002) 1045. [32] D. Haffad, A. Chambellan, J.C. Lavelley, J. Mol. Catal. A: Chemical 168 (2001) 153.

Das könnte Ihnen auch gefallen