Sei sulla pagina 1di 20

Applied Catalysis A: General 246 (2003) 4968

Oxidative dehydrogenation of isobutane over V2 O5-based catalysts prepared by grafting vanadyl alkoxides on TiO2 SiO2 supports
V. Iannazzo a , G. Neri a , S. Galvagno a , M. Di Serio b , R. Tesser b , E. Santacesaria b,
a b

Dipartimento di Chimica Industriale e Ingegneria dei Materiali dellUniversit di Messina, Salita Sperone 31, Messina 98166, Italy Dipartimento di Chimica dellUniversit di Napoli Federico II, Complesso di M. te S. Angelo, Via Cinthia, Napoli 80126, Italy Received 31 July 2002; received in revised form 9 December 2002; accepted 10 December 2002

Abstract Following four different procedures many vanadium-based catalysts have been prepared by using the grafting technique and have been tested on the oxidative dehydrogenation of isobutane. The best results of selectivity have been obtained with catalysts prepared by grafting bimetallic vanadiumtitanium alkoxides directly on silica. The alkoxide precursors have been obtained by partially hydrolysing titanium alkoxide, dissolved in isopropanol, with a stoichiometric amount of water and reacting then with vanadyl tri-isopropoxide, or alternatively by mixing the two mentioned alkoxides in isopropanol and submitting both to controlled partial hydrolysis. The bimetallic alkoxide grafted on silica show a prevalence of isolated VOTi bonds with respect to polyvanadylic VOV bonds that are prevalent, on the contrary, when vanadyl tri-isopropoxide dissolved in n-hexane is grafted on a TiO2 SiO2 support. Catalysts characterised by the prevalence of VOTi bonds are slightly less active but two times more selective than catalysts in which VOV bonds prevail. The preparation of vanadium-based catalysts with a favourable TiO2 environment has been largely simplied by avoiding the use of a TiO2 SiO2 support obtaining, in the meantime, a remarkable improvement in the selectivity. 2003 Elsevier Science B.V. All rights reserved.
Keywords: Isobutane; Oxidative dehydrogenation; Grafting alkoxides; V2 O5 ; Hydrolysis

1. Introduction Isobutene is an important feedstock for petrochemical, polymer and chemical industries [1,2]. Light olens can be produced by dehydrogenation, at high temperature, of the corresponding alkanes. However, the catalytic dehydrogenation still suffers from a number of limitations including high energy input and catalyst deactivation. The light alkane oxidative dehydrogenation (ODH) represents an alternative for the production of these chemicals, provided that highly selective catalysts are developed. A number of studies on the ODH of isobutane to isobutene are reported

Corresponding author.

in the literature [39]. It can be pointed out that supported V2 O5 , a well-established catalyst for the ODH of propane [1016], was less investigated in the ODH of isobutane. With conventional supported V2 O5 catalysts, low selectivity to the desired olenes were reported. Hoang et al. [17] obtained over V2 O5 /Al2 O3 catalysts a selectivity of less than 15%, at 7% of isobutane conversion. However, a very recent paper of Zhang et al. [18] reports that vanadium-containing MCM-41 catalysts prepared by a direct hydrothermal (DHT) method show selectivities to isobutene higher than 40%, at a conversion of about 10%. On the other hand, it has been clearly shown that activities and selectivities of vanadia-based catalysts in the ODH of alkane strongly depend on the VOx environment,

0926-860X/03/$ see front matter 2003 Elsevier Science B.V. All rights reserved. doi:10.1016/S0926-860X(02)00668-3

50

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

on the VOx dispersion, on the acidbase properties of the support, and therefore, on the method of catalyst preparation [19]. Activities and selectivities also depend on the size of the alkane molecule, being different for the same catalyst for respectively ethane, propane and butane [20]. Vanadia catalysts prepared by grafting vanadyl tri-isopropoxide on a titaniasilica support were investigated with the aim of developing catalysts for the selective ODH of isobutane to isobutene. It is well known that titania allows to obtain a very good dispersion of V2 O5 on its surface [21,22], but the resulting samples show low surface area and low resistance to sintering. On the contrary, SiO2 weakly interacting with V2 O5 favours the thermally induced agglomeration of VOx species on the surface [22]. Hence, the preparation of mixed TiO2 /SiO2 supports by co-precipitation, by impregnation of titanium salts on silica or by grafting titanium alkoxides on silica surface, is a common practice. In particular, the grafting technique is an interesting route to obtain TiO2 /SiO2 supports [2327] with: (i) a surface area higher than that usually obtained with titania (typically 50100 m2 /g); (ii) a high resistance to sintering and good mechanical properties; (iii) a higher dispersion of the surface active species. The grafting method, can also be used then to anchor vanadyl chloride or vanadyl tri-isopropoxide, in a water-free solvent, on the hydroxyl groups of a support. This technique has been used to obtain the VOx phase on respectively: SiO2 [28,29], Al2 O3 , [28,29], TiO2 [2832], TiO2 /SiO2 [3335]. In the present work, vanadium-based catalysts have been prepared by using four different grafting procedures: (a) grafting vanadyl tri-isopropoxide, dissolved in n-hexane, on TiO2 /SiO2 ; (b) grafting vanadyl tri-isopropoxide, dissolved in n-hexane and partially hydrolysed by a controlled procedure before grafting on TiO2 /SiO2 ; (c) grafting vanadyl tri-isopropoxide, dissolved in isopropanol and partially hydrolysed by a controlled procedure before grafting on TiO2 /SiO2 ; (d) grafting directly on SiO2 mixtures of titanium and vanadylic alkoxides, dissolved in isopropanol and

submitted to partial hydrolysis, performed in different ways, before grafting. Partial hydrolysis of vanadyl tri-isopropoxide has the objective of inducing a moderate molecular aggregation, before grafting, with the aim of modifying vanadium dispersion on the surface and verify, therefore, the effect of vanadia dispersion on the catalytic activity and selectivity in the ODH of isobutane. Partial hydrolysis of mixtures of titanium and vanadyl alkoxides brings, on the contrary, to bimetallic alkoxides in isopropanol solution that have been anchored directly on SiO2 support. In this way, the catalyst preparation of a vanadium-based catalyst having an intimate TiO2 environment is largely simplied without loosing, as it will be seen, activity but increasing selectivity. An increased selectivity has been observed also for catalysts in which vanadyl tri-isopropoxide is grafted on silica coated with a sub-monolayer of TiO2 . As it will be seen, the conclusion is that for obtaining good selectivities we need a good dispersion of vanadia and a prevalence of VOTi bonds with respect to VOV ones. A simplied kinetic approach has shown that a complex reaction scheme is operative in which CO and CO2 are produced by both isobutane and isobutene. The inuence of the reaction parameters such as, temperature, isobutane and oxygen partial pressure and contact time on the reaction products distribution, over all the mentioned catalysts, is reported. The structure of both the titaniasilica supports, prepared by a multi-step grafting procedure, and the vanadia catalysts were characterised by several techniques including scanning electron microscopy (SEM) with EDX, X-ray diffraction (XRD), BET surface area measurements, temperature programmed reduction and oxidation (TPR, TPO), and FTIR- and UV-diffuse reectance. In the present work, an effort has been made to correlate the observed properties with the obtained performances.

2. Experimental 2.1. Catalysts and supports preparation methods The TiO2 SiO2 support was prepared in three steps by grafting, rst of all, titanium isopropoxide

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

51

dissolved in toluene on commercial SiO2 (furnished by Grace, type S 432). The amount of titanium isopropoxide, dissolved in toluene and used for the rst step of grafting, roughly corresponds to a monolayer or to a moderate excess with respect to a monolayer, by assuming a conventional stoichiometry of one hydroxyl for one alkoxide molecule. Silica, after calcination at 500 C for 28 h, was contacted with the mentioned solution by reuxing, at the boiling point of the solvent for 6 h. The density of the OH groups on the silica surface was determined by means of thermogravimetric analysis (TGA), as previously reported [27]. The solid obtained after the rst grafting step was recovered by ltration, washed with toluene, dried at about 105 C, steamed at 190 C for 2 h, for eliminating residual alkoxide groups from the surface by hydrolysis and nally calcined at 500 C for 2 h. The same procedure was repeated other two times to obtain a support of silica coated with a multi-layer of TiO2 . The properties of the silica coated supports in correspondence of each of the three different grafting steps are reported in Table 1, together with some properties of the solids obtained after each step. The differences observed in the two series of TiO2 SiO2 supports prepared and reported in Table 1 correspond to the differences in the adopted calcination times (2 h at 500 C for the TS11 to TSM1 series and 8 h for TS1 to TSM series) and in the amounts of titanium alkoxide used for each step. A support Tsm S containing a sub-monolayer of TiO2 has also been prepared. The supports of silica coated with multi-layers of titania were then contacted with different solutions of vanadyl tri-isopropoxide in n-hexane or isopropanol at room temperature for 24 h under He atmosphere. After reaction the samples were ltered, washed in

n-hexane, dried at 105 C, steamed at 190 C for 2 h and calcined at 500 C for 2 h. In Table 2, a list of all the prepared catalysts is reported together with the conditions adopted for the preparation, the loading of TiO2 and V2 O5 and other properties. The acronyms of the catalysts are easily interpretable and summarise both the preparation methods and the compositions. We have named for example: S = SiO2 , T = TiO2 , TSM and TSM1silica coated with a multi-layer of TiO2 , Tsm Ssilica coated with a sub-monolayer of TiO2 , Hn-hexane solvent of the precursor, I isopropanol solvent. Sufx h corresponds to a hydrolysed alkoxide. Catalysts of VH/TSM and VH/TSM1 series, for example, were prepared by grafting vanadyl tri-isopropoxide dissolved in n-hexane on TiO2 SiO2 supports. Grafting was made under an inert and dry atmosphere, at room temperature, by contacting the support with the stirred solution. The grafting yields in this case is almost quantitative. Catalysts of the Vh H/TSM1 series were prepared by owing moistured nitrogen in a vanadyl tri-isopropoxide solution in n-hexane, for different times, inferior to the time necessary for obtaining a precipitate of vanadium oxide hydrate (about 1617 min). Partially hydrolysed vanadium alkoxides, so obtained, are then grafted on the TiO2 SiO2 support by contacting solution and solid at room temperature, for 24 h under a dry helium atmosphere. After grafting reaction, the catalyst was recovered by ltration, washed with the used solvent, dried at 105 C, steamed for 2 h at 150 C and nally calcined at 500 C for 2 h. Catalysts of the Vh I/Tsm S and Vh I/TSM type were prepared by grafting partially hydrolysed vanadyl tri-isopropoxide, dissolved in 2-propanol, on the

Table 1 Properties of the prepared supports TSM1, TSM and Tsm S and of the solids obtained during multi-step titanium alkoxide grafting on silica Sample SiO2 TiO2 SiO2 TiO2 SiO2 TiO2 SiO2 TiO2 SiO2 TiO2 SiO2 TiO2 SiO2 TiO2 SiO2 Acronyms S TS11 TS21 TSM1 TS1 TS2 TSM Tsm S Grafting step 0 1 2 3 1 2 3 1 TiO2 (wt.%) 7.0 13.7 17.8 5.9 9.7 11.3 2.3 Metal initial amount (mmol/g) 1.37 1.37 1.37 0.95 0.95 0.95 0.30 SSA (m2 /g) 282 Pore volume (cm3 /g) 1.02 0.23 0.26 0.27 0.81 OH density (mmol/g) 0.92

289 237 267 299 245

0.71 0.79 0.82

52

Table 2 List of the prepared catalysts, their properties and preparation modalities Catalyst V2 O5 /TiO2 SiO2 V2 O5 /TiO2 SiO2 V2 O5 /TiO2 SiO2 V2 O5 /TiO2 SiO2 V2 O5 /TiO2 SiO2 Acronyms VH/TSM VH/TSM1(1) VH/TSM1(2) VH/TSM1(4) Vh H/TSM1 (5)a TiO2 (mmol/g) 1.41 2.20 2.20 2.20 2.20 V2 O5 (wt.%) 0.80 0.65 2.00 3.60 1.03 Preparation modalities Grafting in n-hexane Grafting in n-hexane Grafting in n-hexane Grafting in n-hexane Grafting of partially hydrolysed vanadyl tri-isopropoxide in n-hexane Grafting of partially hydrolysed vanadyl tri-isopropoxide in n-hexane Grafting of partially hydrolysed vanadyl tri-isopropoxide in n-hexane Grafting of partially hydrolysed vanadyl tri-isopropoxide in isopropanol Grafting of partially hydrolysed vanadyl tri-isopropoxide in isopropanol Grafting of partially hydrolysed vanadyl tri-isopropoxide reacted with titanium tetra-isopropoxide in isopropanol Grafting of partially hydrolysed vanadyl tri-isopropoxide in mixture with titanium tetra-isopropoxide in isopropanol Grafting of partially hydrolysed titanium tetra-isopropoxide reacted with vanadyl tri-isopropoxide in isopropanol BET surface area (m2 /g) 249 243 250 258 280 Pore volume (cm3 /g) 0.29 0.30 0.38 0.28 (Oads /Vsupp ) 0.73 1.00 0.59 V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

V2 O5 /TiO2 SiO2

Vh H/TSM1 (10)a

2.20

1.05

269

V2 O5 /TiO2 SiO2 V2 O5 /TiO2 SiO2 (H2 O/Vh = 1)

Vh H/TSM1 (15)a

2.20

0.98

239

Vh I/TSM

1.41

0.90

314

0.26

0.62

V2 O5 /TiO2 SiO2 sub-monolayer of TiO2 (H2 O/Vh = 1) V2 O5 TiO2 /SiO2 (H2 O/Vh = 1)

Vh I/Tsm S

0.30

0.56

(Vh -T)I/S (Ti/V = 4)

0.22

0.80

335

0.45

V2 O5 TiO2 /SiO2 (H2 O/(V-T)h = 1)

(V-T)h I/S (Ti/V = 12)

0.91

0.80

303

0.70

1.00

V2 O5 TiO2 /SiO2 (H2 O/Th = 1)

(Th -V)I/S (Ti/V = 12)

0.35

0.90

348

0.58

1.00

Hydrolysis time in minutes.

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

53

corresponding supports, again by taking the solution and the solid in contact at room temperature for 24 h under a dry helium atmosphere. Hydrolysis was carried out by dissolving 1 ml of vanadyl tri-isopropoxide in 4 ml of 2-propanol and by adding a stoichiometric amount of water (1 mol of water/1 mol of vanadyl tri-isopropoxide) containing traces of HCl acting as catalyst. The operation has been made at room temperature for 5 h, under stirring, always by keeping the solution under a dry helium atmosphere. After the grafting reaction, the catalyst was recovered by ltration, washed with the used solvent, dried at 105 C, steamed for 2 h at 150 C and nally calcined at 500 C for 2 h. The grafting yields in these cases are not quantitative for the inuence of the solvent (the parent alcohol) on the following equilibrium: surface OH + Me(OR)n surface O Me(OR)n1 + ROH (1)

As a consequence, small amounts of vanadium oxide (<1 wt.%) are loaded on the support. Catalysts of (V-T)I/S series have been prepared, on the contrary, by grafting a bimetallic vanadium titanium alkoxide dissolved in isopropanol directly on SiO2 . The bimetallic alkoxide has been obtained in different ways by reacting, for example, partially hydrolysed vanadyl tri-isopropoxide, with titanium tetra-isopropilate (Vh -T)I/S or by hydrolysing rstly titanium tetra-isopropilate and reacting it with vanadyl tri-isopropoxide (Th -V)I/S, or at last by hydrolysing a mixture of the two alkoxides both dissolved together in isopropanol (V-T)h I/S. The grafting operations were performed in all cases at room temperature by stirring in the presence of a dry helium atmosphere for 24 h. The molar ratio vanadiumtitanium has been kept in the range 1:4 to 1:12, as it can be seen in Table 2. After grafting, all the catalysts were recovered by ltration, washed with 2-propanol, steamed at 150 C and calcined at 500 C for 2 h in the usual way. 2.2. Catalysts and supports characterisation methods The morphological characterisation and elemental composition of the catalyst was carried by SEM on a JEOL JSM 5600LV instrument coupled with an EDX analyser.

XRD patterns were recorded on an X3000 Seifert diffractometer equipped with a lithium uoride monochromator on the diffracted beam. The scans were collected within the range of 444 (2) using Mo K radiation. Diffuse reectance spectra were obtained on a UV-Vis scanning spectrophotometer Shimatzu AV2101, equipped with an integrating sphere, using BaSO4 as reference. UV-Vis spectra were recorded in the diffuse reectance mode (R) and transformed to a magnitude proportional to the extinction coefcient (K) through the KubelkaMunk function (F(R)). FTIR and DRIFT spectra have been made by using a NICOLET AVATAR 360 instrument. Samples were examined in the DRIFT cell, at room temperature, after the outgassing in vacuum at 400 C for 1 h. TPR proles were obtained under the following conditions: sample weight 100160 mg, heating rate (from 373 to 873 K) 7 K/min, ow rate 60 cm3 /min, H2 N2 6 vol.%. Prior to the determination of O2 uptake, the reduced samples were cooled down to 653 K under He ow (60 cm3 /min). O2 chemisorption data were achieved by dosing small amounts (0.15 cm3 ) of O2 over the pre-reduced catalysts. Pulses were repeated until peaks of constant area were obtained. Textural analysis was carried out on a Sorptomatic 1990 System (Fisons Instruments) by determining the nitrogen adsorption/desorption isotherms at 77 K. Before analysis, the samples were heated overnight under vacuum up to 473 K (heating rate = 1 K/min). The specic surface area data were assessed by the BET and BJH methods [36,37]. 2.3. Catalytic runs Experimental runs, lasting 8 h, were performed in a quartz-glass microreactor (1.1 cm of internal diameter) at P = 1 atm and T = 350500 C (catalyst weight in the range 0.110.40 g). Prior to the catalytic test, samples were activated in situ at 500 K for 4 h, under air ow (60 cm3 /min). The reaction mixtures normally contained 6.214% of C4 H10 and about 1317% of O2 diluted in He. Total ow rates have been kept in the range 200500 m3 /min. On-line GC analyses were carried out by using two different gas chromatographs equipped with two different gas-chromatographic columns one suitable for O2 , CO, CO2 analysis (a packed column 60/80 of

54

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

Carboxen 1000) using a TCD detector and another one suitable for the hydrocarbons analysis (a column of fused silica Plot with Al2 O3 /KCl of 50 m length, 0.53 mm of diameter, using a stationary phase Chrompack) using a FID detector. Blank runs, performed in the empty reactor at 500 C, allowed to rule out the occurrence of homogeneous reactions to a signicant extent. Conversions and selectivities were dened as it follows: Y= Si = molisobutane reacted molisobutane in the feed 100 100
Fig. 1. Evolution of the conversion with temperature for some different catalysts.

molisobutane converted to i molisobutane reacted

where i = i-C4 H8 , CO, CO2 and others (CH4 , C2 H6 , C3 H8 , C3 H6 , n-C4 H10 , etc.). Catalytic runs have been performed by changing for each catalyst the temperature and the residence time. In some runs the partial pressure of the reagents have been changed too.

3. Results 3.1. Catalytic activities and selectivities Results obtained in the ODH of isobutane over the investigated catalysts have shown that, under the experimental conditions adopted, isobutene, carbon monoxide and carbon dioxide were the main reaction products. Only small amounts (<2%) of methane, ethane, propane, propene, butadiene, n-butene isomers were observed. No oxygenated products (metacrolein and metacrylic acid) were instead detected. A blank test carried without catalyst in the reactor showed no conversion of isobutane up to 500 C. No deactivation of the catalysts was noted for time on stream up to 8 h. Table 3 reports the results of a preliminary screening of all the catalysts investigated in the ODH of isobutane carried out under the specied conditions PC4 H10 = 6, 7 kPa, PO2 = 13 kPa, Ftot = 300 cm3 /min, Wcat = 0.4 g, that is, only temperature has been changed in the different runs performed. Table 4 reports, on the contrary, kinetic runs performed at constant temperature of 450 C by changing the residence time of both the overall feed rate and of the isobutane reagent and in some cases by changing

the oxygen concentration entering in the reactor. From the runs reported in Table 3, it can easily be seen that the most active catalysts are: Vh I/TSM, VH/TSM, VH/TSM1(2) and VH/TSM1(4). This can be better appreciated by comparing the activities reported in Figs. 1 and 2, while the selectivities are much higher for the catalysts (Th -V)I/S, (V-T)h I/S and Vh I/Tsm S. As the activity of Vh I/Tsm S catalyst is quite low, we can conclude that the best catalysts resulted (V-T)h I/S and (Th -V)I/S having activities comparable with the other catalysts but higher selectivity. Selectivities to isobutene are poorly affected by the temperature for all the catalysts, while selectivities to CO increases at the expense of selectivities to CO2 . It is interesting to observe from Fig. 3 that we can individuate, for

Fig. 2. Evolution of the conversion with temperature for some different catalysts.

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968 Table 3 Catalytic runs performed at different temperatures Catalyst code V2 O5 (wt.%) T ( C) Isobutane conversion (%) Selectivities (%) i-C4 H8 VH/TSM 0.8 0.8 0.8 0.8 0.65 0.65 0.65 0.65 2 2 2 2 3.6 3.6 3.6 3.6 0.9 0.9 0.9 0.9 0.9 0.56 0.56 0.56 0.56 0.9 0.9 0.9 0.9 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 325 340 380 430 380 420 450 520 335 380 390 450 310 330 350 500 340 350 370 380 470 405 420 455 480 380 420 470 490 380 420 465 500 380 410 450 480 515 3.3 5 12.3 16.9 3.7 7.9 11.9 22.6 2.6 5.2 9.1 21 3.4 5.8 6.9 34.9 3.6 4.5 7.9 16.6 36.1 2.4 3.1 4.9 5.9 3.8 6.2 9.4 11.2 3.1 6.5 11.2 14.2 4.4 6.1 9.4 9.8 16.4 25.2 17.5 10 8.3 13.3 9 8.7 8.2 9.3 11.9 11.6 9 20.2 14.6 13.6 3.7 18.1 14.3 11.2 8.3 10.1 24 26.4 24.5 23.6 29 23.4 20.9 19.7 15.3 12.6 12.4 13.9 29.5 26.9 23 25.2 19.4 CO 30.5 36.8 48.2 53.2 43 46.7 52.5 60.6 36.1 36.3 42.4 57 37.6 41.9 42 64.4 35.6 39.7 44.6 54.6 61.1 32.5 37.6 40.4 41.3 41.6 43.6 47.4 48.3 39.6 47.8 53.4 59.5 35.5 40.3 45.3 46.6 50.1

55

CO2 41.2 44.8 40.3 37.4 41.2 41.6 36 27.4 52.1 46.6 41.6 31.1 39.4 42.3 42.7 31.3 44.3 44.1 42.1 35 26.5 35.1 28.2 27.9 28.8 27 30.6 30 30.2 37.3 35.6 30.3 22.7 30.9 29.4 30 27.3 28.8

VH/TSM1(1)

VH/TSM1(2)

VH/TSM1(4)

Vh I/TSM

Vh I/Tsm S

(Th -V)I/S

(Vh -T)I/S

(V-T)h I/S

Other reaction conditions have been kept constant: catalyst amount = 0.4 g; Ftot = 300 ml/min.

the examined catalysts, two different ranges of selectivity and this is probably related to the prevalence on the surface of two different catalytic sites. The most selective catalysts have in the temperature range 350500 C selectivities changing from 30 to 20% by

increasing the conversion, that is about the double of the less selective ones. From the runs of Table 4 other interesting information can be obtained. In Fig. 4, for example, it is possible to appreciate the evolution of the conversions and selectivities with the residence

56

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

Table 4 Catalytic runs made at 450 C by changing residence time and reagents partial pressures Catalyst code Catalyst weight (g) 0.3 0.3 0.3 0.1 0.1 0.1 0.3 0.3 0.1 0.1 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.4 0.3 0.4 0.3 0.4 0.3 0.3 0.3 0.3 0.25 0.25 0.31 0.31 0.3 0.3 0.4 0.4 0.4 V2 O5 (wt.%) 0.65 0.65 0.65 0.65 0.65 0.65 2 2 3.6 3.6 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.8 0.98 0.98 1.05 1.05 1.03 1.03 0.56 0.56 0.56 Tot (ml/min) Pi-C4 H10 (kPa) PO2 (kPa) Isobutane conversion (%) 9.5 7.6 9.4 6.1 4.9 4.3 28.8 14.6 24.3 18.2 10.3 6.4 4.1 4.6 5.1 7 7.5 7.4 8.5 9.4 7.4 5.2 9.1 7 5.1 4.9 5.4 4.6 23.9 5.1 8.1 13.8 7.1 14.6 10.7 14.9 4.9 8.2 6.7 3.9 Selectivities (%) i-C4 H8 8.7 10.6 10.2 10.2 19.1 20.5 7 12.3 8.4 10 23.5 29.5 43.9 42 40.3 29.4 26.6 26.1 29 21.5 31.9 41.6 29.2 29 38.3 35.9 36.4 31.8 11.5 25.6 22.6 12 14.6 10.1 15.5 12.2 19.1 22.2 23.5 33.1 CO 54.4 54.8 54.3 48.7 43.7 37.4 51.7 50.4 58.5 59.6 47.7 42.5 33.8 35.3 33.9 41.1 44.8 41.8 43.3 25.9 27.6 33.6 39.4 40.1 32.4 38.8 33.2 40.5 54.8 44.2 46.1 46.8 43.4 50.7 46.8 53 43.7 38.1 38.2 38.7 CO2 35.4 32.7 33 36.5 30.2 28.7 38.8 33.5 31.4 28.8 27.8 26.8 21.3 20.8 25.2 28.7 27.7 31 27.6 52.1 20.7 23.8 29.1 27.2 27.2 26.4 26.8 32.5 28.6 24.7 22.7 33.4 28.7 35.2 33.7 31.7 30.2 27.6 33.6 34.1

VH/TSM1(1)

200 300 500 200 300 490 200 500 200 300 200 300 380 540 300 300 300 300 300 300 300 300 200 200 265 300 480 490 200 300 500 200 490 200 300 200 300 200 300 500

7.2 7.69 7.09 7.2 7.77 6.2 7.6 7.09 7.6 7.62 9.63 6.01 7.95 8.26 13.3 13.3 13.3 13.3 5.9 7.6 10.2 14.5 7.2 7.2 7.26 7.77 6.33 6.2 7.2 7.77 7.09 7.2 6.2 7.2 7.77 7.2 7.77 7.2 7.77 7.09

13.17 13.04 13.17 13.17 13.17 12.41 13.17 13.17 13.17 12.91 13.07 13.07 13.07 13.07 3.1 7.6 9.9 12.9 7.85 7.85 7.85 7.85 13.17 13.17 13 17.17 17.72 12.41 13.17 13.17 13.17 13.17 12.41 13.17 13.17 13.17 13.17 13.17 13.17 13.17

VH/TSM1(2) VH/TSM1(4) (Th -V)I/S

(Th -V)I/S

(Th -V)I/S

(V-T)h I/S

(Vh -T)I/S

Vh H/TSM1 (15) Vh H/TSM1 (10) Vh H/TSM1 (5) Vh I/Tsm S

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

57

Fig. 3. Evolution of the selectivity with the temperature for different catalysts.

Fig. 5. Inuence of the vanadium load on the activity for VH/TSM1 type catalysts.

time for a given catalyst (VH/TSM1(1)), at a xed temperature of 450 C. Similar trends are observable also for other catalysts, that is, conversion obviously increases with the residence time, while selectivity to isobutene slightly decreases favouring the formation of CO and CO2 . In Fig. 5, it is possible to appreciate the effect of vanadium loading on the activity of the catalysts of the VH/TSM1 type. The activity seems to increase about linearly with the vanadium load in agreement with the very high dispersion of these catalysts. Fig. 6, reporting selectivity to isobutene as a function of the conversion for different catalysts, conrms the existence of two ranges of selectivity for the examined catalysts and the most selective catalysts resulted again (Th -V)I/S and (V-T)h I/S.

A kinetic approach applied to the overall isobutene conversion has then been made, derived from the Mars and Van Krevelen model [38], simplied by adopting a pseudo-rst-order kinetic law in agreement with the suggestion of Boisdron et al. [39]. We have interpreted with this model the kinetic runs of some of the catalysts reported in Tables 3 and 4. The obtained results are reported in Table 5. As it can be seen, the apparent activation energy for the overall reaction resulted about 914 Kcal/mol with signicant changes from one to another catalyst. It is also interesting to observe that the activation energy is much smaller than the one obtained for propane on similar

Fig. 4. An example of the evolution of conversion with residence time for the VH/TSM1(1) catalyst.

Fig. 6. Selectivity as a function of conversion for different catalysts.

58

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

Table 5 Kinetic parameters and tting errors on conversion Catalyst Activation energy (Kcal/mol) 14.3 11.8 8.7 8.8 13.8 Pre-exponential factor (mol/gcat. /h/atm) 5229 1881 53.6 44.7 1236 Mean absolute percent error on isobutane conversion 24.6 12.7 15.8 21.6 14.9

VH/TSM1(1) VH/TSM1(4) (Th -V)I/S (V-T)h I/S Vh I/Tsm S

catalysts [40] (about 20 Kcal/mol) in agreement with a more easy hydrogen abstraction from the hydrocarbon molecule. Moreover, the high average errors given by the model on the conversions calculations introduce some doubts on the correctness of the kinetic analysis. As a matter of fact, by extrapolating to zero contact time the selectivities of the main reaction products (see, for example, Fig. 4), it is easy to observe that high values are obtained not only for isobutene but also for CO and CO2 . This means that a complex kinetic scheme is operative in which CO and CO2 are obtained not only from isobutene but also from isobutane. The kinetic approach could be deepened, therefore, only by performing much more kinetic runs in order to determine the kinetic parameters of the several parallel-consecutive occurring reactions. 3.2. Supports and catalysts characterisation 3.2.1. Specic surface area and pores distributions The support TiO2 SiO2 has been obtained, as previously described, by repeating three times the grafting procedure. The amount of grafted TiO2 for each grafting step is reported in Table 1 together with the specic surface area, the pore volume and the surface hydroxyls density (obtained by TGA as described elsewhere [27]). It is interesting to observe, rst of all, that the specic surface area of TiO2 SiO2 remain very high after titania grafting, comparable with the specic surface area of the original silica support. Pore volume is, on the contrary, reduced for the appearance of micropores (2030% of the pore volume) that was not present in the original silica support. However, a TiO2 SiO2 support of high surface area and stable to thermal treatment is obtained. As mentioned before, the difference in TSM and TSM1 supports consists

only in the duration time of SiO2 calcination and in the different concentration of the titanium alkoxide solution put in contact with the solid during the successive grafting steps. OH density does not change too much in the rst two steps of grafting. A 2/2 stoichiometry can, therefore, be suggested as prevalent in the grafting reaction, i.e. two hydroxyls reacting with two alkoxide groups of a molecule of titanium alkoxide. Vanadyl alkoxide grafting in some cases reduces further the specic surface area, but this remains always high and comparable with that of the original support, in other cases an increase of the surface area is observed, as it can be seen in Table 2. In particular, catalysts prepared by grafting bimetallic alkoxide directly on silica do not show the strong decrease of pore volume observed in the case of TiO2 SiO2 supports. 3.2.2. XRD analyses XRD analyses have been performed on different supports and catalysts such as: TSM, TSM1, VH/ TSM, Vh I/TSM, (Vh -T)I/S, (V-T)h I/S and (Th -V)I/S. For the support TSM and the catalysts of the type VH/TSM and Vh I/TSM, XRD analyses show a signal corresponding to the presence of small crystallites of TiO2 in the form of anatase. The size of these crystallites is somewhat higher in the presence of vanadium oxide. However, the amorphous part of TiO2 is largely predominant. The catalysts of the type (Vh -T)I/S, (V-T)h I/S and (Th -V)I/S resulted completely amorphous. It is important then to point out that we have never observed crystalline V2 O5 in all the examined catalysts. The conclusion is that TiO2 coating is mainly composed of amorphous TiO2 with some small crystallites of anatase, while V2 O5 is always highly dispersed and completely amorphous. More details about the properties of the TiO2 SiO2 support are reported elsewhere [27]. 3.2.3. Morphological analyses at SEM and EDX SEM analyses performed on the TSM and TSM1 supports have shown the presence of particles of irregular shape having a wide size distribution (in the range 160 m). No large cluster of TiO2 were observed on the surface. This, coupled with EDX elemental mapping analysis, indicates an homogeneous distribution of grafted titanium species on the surface of silica. Catalysts prepared by grafting vanadium on supports TSM and TSM1 through procedures ac show

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

59

Fig. 7. SEM micrograph showing the morphology of the sample Vh I/Tsm S.

the same morphological features observed on the parent supports (see micrograph in Fig. 7 taken on the sample Vh I/Tsm S). Grafted vanadium species are then supposed to be highly dispersed on the carriers. On the contrary, on catalysts of (V-T)I/S series, prepared by procedure d, has been also noticed the presence of separate aggregates composed, according to EDX analysis mainly of titanium and vanadium (Fig. 8 referred to catalyst (Th -V)I/S). These aggregates show no crystalline shape, suggesting they are likely amorphous. 3.2.4. Spectroscopic analyses FTIR and DRIFT spectra have been collected for many supports and catalysts reported in Tables 1 and 2. It is known from the literature [41] that pure V2 O5 shows FTIR absorption bands at respectively 515, 603, 827, 950, 1019, 1600, 3666 cm1 . The absorption bands between 550 and 670 cm1 are attributed to the rocking modes of the VOV bonds, while the bands between 670 and 770 cm1 correspond to the stretching of the same bonds. The absorption band at 980 cm1 is normally attributed to the symmetric

stretching of V=O bond in amorphous V2 O5 , while the band at 1020 cm1 corresponds to the vibration of the same bond and is characteristic of the crystalline V2 O5 . At last, the band at 3670 cm1 is associated to the vibration of the bond VOH. For monolayers of V2 O5 , the band at 1020 cm1 disappears and that at 980 is enhanced [41]. So it is possible to recognise the presence of crystalline V2 O5 in a catalyst. For very low charge of V2 O5 , as in our case, crystalline V2 O5 is completely absent and the band at 1020 cm1 is never observed, while absorption at 960990 cm1 conrms the formation of polyvanadylic species of low nuclearity [42]. The FTIR spectra of the supports TSM and TSM1 show four absorption bands. The most intense, at 1104 cm1 , corresponds to the asymmetric stretching vibration of the bonds SiOSi for the tetrahedric SiO4 units [4345]. Other bands are observable at 1200 cm1 (asymmetric stretching vibration of the SiO bonds) [46], 800 cm1 (symmetric stretching vibration of the same bond) and at about 950 cm1 corresponding to the asymmetric stretching vibration of the SiOTi bond [45,47,48]. SiO2 , the original support shows an intense band of

60

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

Fig. 8. SEM micrographs and EDX elemental analysis of the sample (Th -V)I/S. Numbers refer to regions where corresponding EDX patterns have been collected.

absorption at 3747 cm1 corresponding to isolated silanol groups SiOH [49,50]. This peak disappears as a consequence of the titanium alkoxide grafting, as it can be appreciated in Fig. 9 where the spectra for SiO2 and TiO2 SiO2 (TSM) are compared. In the same gure are also reported, for comparison, the spectra of (V-T)h I/S and of VH/TSM. As it can be seen, the spectrum for VH/TSM is quite similar to the one for TSM, while the spectrum for (V-T)h I/S show many silanol groups in agreement with the fact that in this case a bimetallic alkoxide of vanadium and titanium is directly grafted on silica without forming a monolayer. The presence of intense bands characteristic of the support and the presence of small amounts of loaded V2 O5 in the prepared catalysts, give place to DRIFT spectra of low intensity in the wavenumber ranges that are peculiar of vanadium bonds. However, some useful information can equally be derived from the observation of these spectra made in the range 4002000 cm1 . In Fig. 10, the DRIFT spectra for the catalysts of the VH/TSM1 type, containing

different amounts of V2 O5 , from 0.65 to 3.6 wt.%, are reported. In this gure, it is possible to observe a slight increase of the VOV absorption bands in the range 550800 cm1 suggesting an increase of the aggregation degree and a consequent decrease of the V2 O5 dispersion. The effect of hydrolysis of vanadyl alkoxide in n-hexane solution performed for different times (catalysts Vh H/TSM1 (5,10,15)) has the effect of forming vanadium aggregates and we also observe in this case a slight increase of the VOV bands in the range 550800 cm1 . Hydrolysis in a polar solvent such as isopropanol, has a more relevant effect that can be appreciated in Fig. 11, where spectra of catalysts obtained from precursors either subjected to hydrolysis or not are compared. It is interesting to observe, for catalysts obtained from bimetallic precursors subjected to hydrolysis (catalysts (V-T)h /I/S, (Vh -T)/I/S), the disappearance of the polyvanadylic VOV absorption bands. This means that the adopted procedure of preparation prevents the V2 O5 aggregation and gives place to isolated

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

61

Fig. 9. Disappearance of silanol SiOH groups as a consequence of titanium alkoxide grafting as it can be seen by DRIFT analysis.

Fig. 10. DRIFT spectra for three VH/TSM1 catalysts characterised by a different load of vanadium.

62

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

Fig. 11. A comparison of DRIFT spectra obtained for different catalysts.

monovanadylic species or conversely to a prevalence of VOTi bonds with respect to VOV ones. Some supports and catalysts have also been submitted to DR-UV analysis. In Fig. 12 examples of the spectra obtained are reported. The effect of V2 O5 in altering the band of reectance shown by the support TSM1 is clear. The band between 250 and 350 nm corresponds to titanium ion in octahedral coordination [51,52]. The broadening of the band is due to V2 O5 absorbing at wavelengths higher than TiO2 and broader. Crystalline V2 O5 absorbs at 500550 nm [41], and as this band is never present in the examined catalysts we can exclude the presence crystals of V2 O5 in agreement with the XRD analyses. V5+ in octahedral coordination exhibits a charge transfer transition at 400480 nm, while tetrahedral coordination shows absorption bands at 300350 nm, absorption bands at 270300 nm are characteristic of isolate V5+ in tetrahedral form [5357]. It is clear from the observation of Fig. 10 that in our dispersed catalysts vanadium is mainly present in a tetrahedral form. This is more evident in Fig. 13 where it is possible to appreciate the effect of vanadium loading on the reectance. In this spectrum the reectance of the support has been subtracted. As it can be seen, the vanadium load

changes both the intensity of the band and the position of the maximum falling in the range typical of tetrahedral coordination. In Fig. 14, a comparison between VH/TSM1(1) and Vh H/TSM1 (for three different times of hydrolysis) is reported. It is interesting to observe the strong difference in the maximum position between VH/TSM1(1) and Vh H/TSM1 catalysts. 3.2.5. TPR and oxygen uptake by pulse technique Different catalysts of the ones reported in Table 2 have been submitted to thermal programmed reduction (TPR) with hydrogen, showing in some cases very different behaviours. Samples reduced at about 550 C with hydrogen and then frozen in Helium at 370 C have been re-oxidised with oxygen pulses at 370 C so determining the oxygen uptake for each vanadium atom. By assuming a stoichiometry O/V = 1, chemisorbed oxygen becomes a roughly evaluation of V2 O5 dispersion [58], even if some authors do not consider this measure reliable because it depends on many parameters [59]. Examples of TPR plots obtained are reported in Figs. 15 and 16. In Fig. 15, TPR obtained

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

63

Fig. 12. A comparison of DR-UV spectra obtained for respectively the TSM1 support and different vanadia catalysts.

Fig. 13. DR-UV spectra obtained for VH/TSM1 catalysts containing different amount of vanadia. Spectrum of the support has been subtracted.

64

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

Fig. 14. A comparison of DR-UV spectra respectively obtained for a VH/TSM type catalyst and three Vh H/TSM catalysts.

for respectively VH/TSM, Vh I/TSM, (Vh -T)I/S, (Th -V)I/S and (V-T)h I/S catalysts are compared, while in Fig. 16 a comparison between the TPR of two VH/TSM1 catalysts with different vanadium load is re-

ported. Reduction initiates at about 250280 C in all cases. In Fig. 15, it is possible to observe that the obtained curves correspond to the addition of two or three different peaks, with a maximum falling at different

Fig. 15. A comparison of TPR plots obtained for different catalysts.

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

65

Fig. 16. Effect of the vanadium load on the TPR plots.

temperatures, that are reasonably related to more or less reducible catalytic sites. Curve related to (V-T)h I/S, for example, can be dissociated in two peaks, one greater having a maximum at about 445 C and another one smaller at 565 C. Catalyst (Th -V)I/S shows three different peaks at respectively 320, 410 and 465 C, that is, the catalyst surface is more heterogeneous and the different vanadium species anchored on the surface are more reducible. Two peaks are also present at respectively 400 and 540 C also for VH/TSM catalyst but with a large predominance of the less reducible sites. A dispersion index for the different catalysts expressed as oxygen uptake for vanadium atom is reported in Table 2 for the catalysts submitted to TPR analysis. As it can be seen, all the catalysts are largely dispersed, in particular, (V-T)h I/S and (Th -V)I/S showing a complete accessibility of the vanadium atoms. The large dispersion of catalyst VH/TSM1(1) is probably due to the very low vanadium charge. In Fig. 16 the effect of vanadium loading on TPR behaviour can be appreciated. As it can be seen, by increasing the amount of vanadium on the surface the catalyst becomes more reducible.

4. Discussion and conclusions Many vanadium-based catalysts have been prepared following different preparation procedures. All the prepared catalysts have been tested in the ODH of isobutane giving place to great differences of both activities and selectivities. A rst group of catalysts has been prepared, for example, by grafting different amounts of vanadyl tri-isopropoxide, dissolved in n-hexane, on a support of silica coated with a multi-layer of TiO2 . The grafting reaction, in this case, is almost quantitative and allows to obtain catalysts containing an increasing amount of vanadium on the surface (see catalysts of VH/TSM1 and VH/TSM type reported in Table 2). The activities of these catalysts resulted roughly proportional to the vanadium content, as it can be seen in Fig. 5. This behaviour suggests a good dispersion for all the considered catalysts. This is not contradictory with the DRIFT analyses for these catalysts showing the presence of polyvanadylic groups because the tendency of V2 O5 to give monolayered structures on TiO2 surfaces [22] until it reaches a complete coverage is well known. The apolar solvent n-hexane, favours the

66

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

molecular aggregation giving place in solution at least to dimeric structures of the type:

as it has been shown for vanadyl tri-ethoxide [60]. Selectivities to isobutene of these catalysts remain relatively low (from 10 to 30%). Moreover, by extrapolating selectivity data to zero contact time (see Fig. 4) it is possible to observe high values for all the main reaction products, that is, isobutene, CO and CO2 This means that CO and CO2 are produced not only by isobutene but also by isobutane, even if in a less extent. A similar behaviour has been observed also for the other catalysts. It is possible to suggest, therefore, a reaction scheme of the type:

This complex scheme of reaction well explains the poor adherence of the pseudo-rst-order kinetic approach to the experimental behaviour. Another group of catalysts has been prepared starting from a dened amount of vanadyl tri-isopropoxide, always dissolved in n-hexane and submitted to an appropriate ow of moistured nitrogen for different times. The times of exposure to moisture has been chosen 5, 10 and 15 min, considering that 1617 min is the time occurring for observing the formation of a precipitate. In this way, hydrolysis and condensation of vanadyl tri-isopropoxide is induced and catalysts obtained have the same vanadium load (about 1 wt.%, see catalysts of type Vh H/TSM1 in Table 2) but would have different dispersions. Activities and selectivities are not particularly affected by the described treatment. Probably, condensation occurs but giving place to bidimensional structures that leave all the vanadium

atoms accessible without changing, therefore, neither the activity nor the selectivity. Other two catalysts have been prepared by grafting partially hydrolysed vanadyl tri-isopropoxide, dissolved in isopropanol and treated with a stoichiometric amount of water (1:1), on respectively silica coated with a multi-layer of TiO2 and silica coated with a sub-monolayer of TiO2 . It is interesting to observe that in the case of using the support of silica coated with a sub-monolayer of TiO2 , partially hydrolysed vanadium alkoxide is grafted exclusively on the islands of TiO2 and not on silica as shown by SEM with EDX analysis. Catalyst Vh I/TSM is one of the most active catalysts but it is not much selective. On the contrary, Vh I/Tsm S is less active but much more selective. The last group of catalysts have been prepared by grafting directly on silica, a vanadiumtitanium bimetallic alkoxide, dissolved in isopropanol, prepared by following three different alternative routes. In the rst case, vanadium alkoxide has been partially hydrolysed and then reacted with titanium alkoxide (Vh -T)I/S. In the second case, titanium alkoxide has been partially hydrolysed and reacted with vanadium alkoxide (Th -V)I/S. In the last case, both the alkoxides of vanadium and titanium in mixture have been treated with a stoichiometric amount of water before grafting on silica (V-T)h I/S. With the exception of (Vh -T)I/S, the catalysts of this group resulted as the most selective ones and activities are inferior but comparable with those of the most active catalysts. It is interesting to observe that the best catalysts (Th -V)I/S and (V-T)h I/S do not show at the DRIFT analysis the presence of polyvanadylic groups in the wavenumber range 550800 cm1 , while these groups are present for both VH/TSM and Vh I/TSM catalysts. This can be interpreted with the formation of isolated vanadium oxide groups directly bounded to titanium oxide grafted on the silica support. Therefore, the prevalence of VOTi bonds in these catalysts with respect to VOV bonds in the others seems to be the reason of the observed high selectivity. Similar results are obtained for both (Th -V)I/S and (V-T)h I/S catalysts probably because the same active sites are formed. This occurs because in the polar solvent, isopropanol hydrolysis is strongly contrasted by the solvent, acting negatively on the equilibrium (1). Titanium alkoxide reacts more fastly than vanadium alkoxide giving place to mononuclear species of

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968

67

the type: Ti(OR)4 + H2 O Ti(OR)3 OH + ROH

Then, Ti(OR)3 OH can react with vanadyl tri-isopropoxide to give: 1, 2 or 3 Ti(OR)3 OH + OV(OR)3 OV[OTi(OR)3 ]1,2 or 3 Being all the catalysts examined well dispersed, the differences in both activity and selectivity are too many, probably due to differences in the active sites. This is conrmed by both the TPR plots reported in Fig. 15, where the most selective catalysts (Th -V)I/S and (V-T)h I/S show a reduction peak at lower temperature and also by the apparent activation energies reported in Table 5 that are lower for the same catalysts. It is worth pointing out in conclusion that the preparation method of vanadium-based catalysts with a TiO2 favourable environment has been realised in a simpler and cheaper way by using a silica support and by obtaining in the meantime a remarkable improvement of the selectivity in the ODH of isobutane.

Acknowledgements Thanks are due to MIUR for the nancial support. References
[1] [2] [3] [4] [5] [6] [7] [8] D. Seddon, Catal. Today 15 (1992) 1. L. Dunn, Hydrocarbon Process. June (1994) 27. S.S. Bharadwaj, L.D. Schimdt, J. Catal. 155 (1995) 43413. M. Huff, L.D. Schmidt, J. Catal. 155 (1995) 8294. F. Cavani, M. Koutyrev, F. Trir, J. Catal. 158 (1996) 23 250. A. Kaddouri, C. Mazzocchia, E. Tempesti, Appl. Catal. A: Gen. 169 (1998) L3L7. P. Moriceau, B. Grzybowska, Y. Babaux, G. Gwrobel, G. Hecquet, Appl. Catal. A: Gen. 168 (1998) 26277. B. Grzybowsa-Swierkosz, J. Sloczynski, R. Grabowski, K. Wcislo, A. Kozlowska, J. Stooch, J. Zielinski, J. Catal. 178 (1998) 8797. S.M. Al Zahrami, N.O. Elbashir, A.E. Abasaeed, M. Abdulwahed, Ind. Eng. Chem. Res. 40 (2001) 781784. A. Corma, J.M. Lpez Nieto, N. Paredes, Appl. Catal. A 97 (1993) 159. R. Burch, E.M. Crabb, Appl. Catal. A 100 (1993) 111. P.M. Michalakos, M.C. Kung, I. Jahan, H.H. Kung, J. Catal. 140 (1993) 226.

[9] [10] [11] [12]

[13] J.G. Eon, R. Olier, J.C. Volta, J. Catal. 145 (1994) 318. [14] E.A. Mamedov, V. Corts Corbern, Appl. Catal. A 127 (1995) 1. [15] N. Boissdron, A. Monnier, L. Jalowiecki-Duhamel, J. Barbaux, J. Chem. Soc., Faraday Trans. 91 (17) (1995) 2899. [16] R. Grabowski, B. Grzybowska, K. Samson, J. Sloczynski, J. Stoch, K. Wcislo, Appl. Catal. A 125 (1995) 129. [17] M. Hoang, J.F. Mathews, K.C. Pratt, React. Kinet. Catal. Lett. 661 (1997) 2126. [18] Q. Zhang, Y. Wang, Y. Ohishi, T. Shishido, K. Takehira, J. Catal. 202 (2001) 308. [19] B. Grzybowsa-Swierkosz, Appl. Catal. A: Gen. 157 (1997) 409420. [20] T. Blasco, J.M. Lopez Nieto, Appl. Catal. A: Gen. 157 (1997) 409420. [21] N.E. Quaranta, J. Soria, V. Corts Corbran, J.L.G. Fierro, J. Catal. 171 (1997) 1. [22] G.C. Bond, Appl. Catal. A: Gen. 157 (1997) 91103. [23] M.G. Reichmann, A.T. Bell, Appl. Catal. 32 (1987) 315. [24] A. Fernandez, J. Leyrer, A.R. Gonzalez-Elipe, G. Munera, H. Knozinger, J. Catal. 112 (1988) 489. [25] A. Munoz, G. Munera, in: G. Poncelet, P.A. Jacobs, P. Grange, B. Del Mon (Eds.), Preparation of Catalysts V, Stud. Surf. Sci. Catal. 63 (1991) 627. [26] R. Castillo, B. Koch, P. Ruiz, B. Delmon, J. Catal. 161 (1996) 524. [27] P. Iengo, G. Aprile, M. Di Serio, D. Gazzoli, E. Santacesaria, Appl. Catal. A: Gen. 178 (1999) 97109. [28] J. Haber, A. Kozlowska, R. Kozlowski, J. Catal. 102 (1986) 52. [29] J. Kijenski, A. Baiker, M. Glinski, P. Dollenmeier, A. Wokaun, J. Catal. 101 (1986) 1. [30] G. Bond, J. Perez Zurita, S. Flamerz, P.J. Gellings, H. Bosch, J.G. Van Hommen, B.J. Kip, Appl. Catal. 22 (1986) 361. [31] M. Schraml-Marth, A. Wokaun, A. Baiker, J. Catal. 124 (1990) 86. [32] G.C. Bond, S.F. Tahir, Appl. Catal. 71 (1991) 1. [33] A. Sorrentino, S. Rega, D. Sannino, A. Magliano, P. Ciambelli, E. Santacesaria, Appl. Catal. A: Gen. 209 (2001) 4557. [34] R. Monaci, E. Rombi, V. Solinas, A. Sorrentino, E. Santacesaria, G. Colon, Appl. Catal. A: Gen. 214 (2001) 203212. [35] M.A. Reiche, E. Ortelli, A. Baicher, Appl. Catal. B: Environ. 23 (1999) 187203. [36] S. Brunauer, P.H. Emmet, J. Am. Chem. Soc. 60 (1938) 309. [37] E.P. Barret, L.G. Joiyner, P.P. Halenda, J. Am. Chem. Soc. 73 (1953) 373. [38] P. Mars, D.W. Van Krevelen, Chem. Eng. Sci. 3 (1954) 41. [39] N. Boisdron, A. Monnier, L. Jalowlecki-Duhamel, Y. Barbaux, J. Chem. Soc., Faraday Trans. 91 (17) (1995) 28992905. [40] A. Comite, A. Sorrentino, G. Capannelli, M. Di Serio, R. Tesser, E. Santacesaria, J. Mol. Catal. A: Chem., in press. [41] Y. Nakagawa, J. Ono, H. Miyata, Y. Kubokawa, J. Chem. Soc., Faraday Trans. I 79 (1983) 2929. [42] G.T. Went, L.-J. Leu, A.T. Bell, J. Catal. 134 (1992) 479491.

68

V. Iannazzo et al. / Applied Catalysis A: General 246 (2003) 4968 [51] S.A. Holmes, F. Quignard, A. Choplin, R. Teissier, J. Kervennal, J. Catal. 176 (1998) 173. [52] R. Hutter, T. Mallat, A. Baiker, J. Catal. 153 (1995) 177. [53] J.G. Eon, R. Olier, J.C. Volta, J. Catal. 145 (1994) 318. [54] G. Lischke, W. Hanke, H.G. Jerschkewitz, G. Ohlaman, J. Catal. 91 (1985) 54. [55] T. Lindblad, B. Rebenstorf, Z.G. Yan, S.L.T. Andersson, Appl. Catal. A: Gen. 112 (1994) 187. [56] P. Concepcion, J.M. Lopez Nieto, J. Perez-Pariente, J. Mol. Catal. A: Chem. 99 (1995) 173. [57] K. Wada, H. Yamada, Y. Watabe, T. Mitsudo, J. Chem. Soc., Faraday Trans. 94 (1998) 5852. [58] G.T. Went, L. Leu, T. Bell, J. Catal. 134 (1992) 479. [59] B. Dyakova, B. Mehandzhiev, B. Grzybowska, I. Gasior, J. Haber, Appl. Catal. 3 (1982) 255. [60] D.C. Bradley, R.C. Mehrotra, D.P. Gaur, Metal Alkoxides, Academic Press, New York, 1978.

[43] M.R. Boccun, K.M. Rao, A. Zecchina, G. Leofanti, G. Petrini, in: Proceedings of the European Conference on Structure and Reactivity of Surfaces, Trieste, 1988, Stud. Surf. Sci. Catal. 48 (1989) 133. [44] A. Duran, C. Serna, V. Fornes, J.M. Fernandez-Navarro, J. Non Cryst. Solids 82 (1986) 69. [45] A. Duran, J.M. Fernandez-Navarro, P. Casariego, A. Joglar, J. Non Cryst. Solids 82 (1986) 75. [46] R.B. Laughlin, J.D. Joannopoulos, Phys. Rev. B16 (1977) 2942. [47] F.L. Galeneer, A.J. Leadbetter, M.W. Stringfellow, Phys. Rev. B27 (1983) 1052. [48] A. Bertoluzza, G. Fagnano, M.A. Morelli, V. Gottardi, M. Guglielmi, J. Non-Cryst. Solids 48 (1982) 117. [49] E. Astorino, J.B. Peri, R.J. Willey, G. Busca, J. Catal. 157 (1995) 482. [50] C.U.I. Odenbrand, S.L.T. Andersson, L.A.H. Andersson, J.G.M. Brandin, G. Busca, J. Catal. 125 (1990) 541.

Potrebbero piacerti anche