You are on page 1of 14

Applied Catalysis A: General 205 (2001) 201214

Structure and activity of Sn-Mo-O catalysts: partial oxidation of methanol


N. Graciela Valente, Luis A. Arra, Luis E. Cads
INTEQUI, Instituto de Investigaciones en Tecnologa Qumica (UNSLCONICET), Chacabuco y Pedernera, 5700 San Luis, Argentina Received 22 October 1999; received in revised form 24 March 2000; accepted 29 March 2000

Abstract This work deals with the catalytic behavior of a series of Sn-Mo-O catalysts in the partial oxidation of methanol. The catalysts of different Sn:Mo ratios, were prepared by co-precipitation and they were investigated by means of dynamic experiments, test reactions (methanolformaldehyde partial oxidation, isopropyl alcohol decomposition) and physico-chemical characterization (XRD, BET, TPD of NH3 , TPR, XPS and EPR). It was observed that interdispersion between MoO3 and SnO2 favors a supercial architecture in which the Sn-Mo interaction plays a major role modifying the reactivity of the lattice oxygens and the reducibility of Mo ions and, therefore, the catalytic behavior. The partial oxidation of methanol induces a reordering of the catalyst structural organization leading to a Mo surface enrichment. The absence of chemical shifts for Sn (XPS) suggests that the OMo bond is mainly responsible for the methanol reaction. 2001 Elsevier Science B.V. All rights reserved.
Keywords: Methanol partial oxidation; Mo-Sn catalysts; Methyl formate

1. Introduction The rst part of the present work on the Sn-Mo-O catalytic system focused on the catalytic behavior of mechanical mixtures of pure Sn and Mo oxides [1]. Various systems based on multicomponent oxides have been the object of study in the oxydehydrogenation (ODH) methanol reaction. In particular, Ai has worked on MoO3 -Bi2 O3 -P2 O5 [2], SnO2 -V2 O5 [3], SnO2 -P2 O5 and SnO2 -MoO3 [4] catalysts. It is well-known that in several reactions catalysts based on multicomponent oxides exhibit a better performance than component oxides separately. This has been accounted for in the literature by means of mechanisms, such as Remote Control by spill over species
Corresponding author. Fax: +54-2652-426711. E-mail address: lcadus@unsl.edu.ar (L.E. Cad s). u

[5] as well as by means of physico-chemical characterizations. However, the role of component oxides in these catalysts is not yet completely understood. The state of the catalyst surface (composition, chemical state of the metals, surface structure, etc.) cannot usually be studied during the catalytic reaction. Besides, the supercial composition in interdispersed systems might be different than that of the bulk and may change in the course of the reaction. An important factor to be considered in a study based on a reaction of oxidative dehydrogenation is the characteristics of the absorbed oxygen species as well as their capacity to migrate in the lattice of the oxide used as catalyst or in its boundary. A further factor to be taken into account is the lability of the metaloxygen bonds. Some authors have proposed that both the acid base properties of the metal oxide and the strength of the metaloxygen bond are

0926-860X/01/$ see front matter 2001 Elsevier Science B.V. All rights reserved. PII: S 0 9 2 6 - 8 6 0 X ( 0 0 ) 0 0 5 6 5 - 2

202

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

directly responsible for catalytic activity [6]. Therefore, changes in the structure cause modications in the catalytic behavior. SnO2 is not a good catalyst for ODH of methanol due to its low activity. However, combination with other metallic oxides, such as MoO3 improves its catalytic behavior. Ai [7] has attributed this activity to the presence of acid and base sites distributed on the catalyst in some particular way. However, this author does not offer a description of the surface that could allow one to identify phases and account for active sites. Weng and Wolf [8], in studies performed by XPS, have suggested the possibility that SnMoO2 coexists with MoO3 in samples treated at temperatures of 873 K, which is signicantly hotter than typical conditions employed for alcohol dehydrogenation. Niwa et al. [9] have noted that, in MoO3 catalysts supported on SnO2 , the increase of activity in the ODH of MeOH can be attributed to changes in the reducibility of the supported Mo. Previously, Niwa et al. [10] had prepared and described a catalyst which consisted of a monolayer of MoO3 supported on SnO2 . By observing the turnover frequency of methanol oxidation versus the Mo surface concentration, these authors concluded that the Mo sites are the active sites for the reaction. At Mo loadings higher than those of the monolayer, they observed formation of large isolated crystals with the consequent increase of acidity while the turnover frequency remains constant. These observations are of great value for catalysts based on mixed oxides, for example, obtained by co-precipitation. However, Valente et al. [1] obtained excellent results with mechanical mixtures prepared under conditions such that the interactions between molybdenum and tin oxides were minimal. The formation of mixed oxides was not observable by the characterization techniques used. Therefore, the observation by Niwa et al. [10] as regards a close relation between structure and activity should be supplemented with explanations that take into account the observations on mechanical mixtures. Thus, additional factors as the extent of Mo reduction and the inuence of tin oxide on such reduction should be included together with geometrical considerations. The description of this system might correspond to complex catalyst whose structure is difcult to determine. The possibilities of the presence of an amorphous phase or contamination of one phase by another cannot be left out. In order to understand

the behavior of this catalyst, it becomes necessary to investigate the physico-chemical properties of the involved phases by mean of catalyst characterization after different pre-treatments. In this way, it would be possible to know the probable surface reordering. In fact, in the ODH of methanol, the proposed mechanisms [11] indicate that the proton and electron transfer between the alcohol and the catalyst is important. The purpose of the present work is to determine the possible physico-chemical transformations of the catalyst. These transformations might be related with the reaction of the adsorbed phases through the oxidationreduction cycles of the catalytic process (dynamic feature). We will try to explain the solid reactivity and the catalyst activity and selectivity according to the possible mechanisms of methanol reaction. Taking into account the complexity of the system, a series of measurements were performed, by means of dynamic experiments, test reactions and physico-chemical characterization. The preparation of catalysts with different Sn:Mo atomic ratios by co-precipitation of their respective precursors had the objective of achieving close contact between the phases of such elements. In this way, it is possible to reach a stable architecture towards which a system of mechanical mixtures might tend.

2. Experimental 2.1. Catalyst preparation 2.1.1. Preparation of pure oxides The reagents used were (NH4 )6 Mo7 O24 4H2 O (Merck AR grade with Fe<0.0005%, Pb<0.001% and Cu<0.001%), and SnCl2 2H2 O (May and Baker AR grade with As<0.0001%). MoO3 was prepared by dissolving ammonia heptamolybdate in distilled water. After evaporation of the solvent at 323 K and reduced pressure, the solid obtained was dried at 373 K overnight in a vacuum oven and then calcined at 823 K for 8 h. SnO2 was prepared from SnCl2 and urea. SnCl2 was dissolved in NH4 OH solution at pH 8.5 with magnetic stirring. Then, urea was added and stirred for 15 min. The obtained suspension was then ltered and washed with a very dilute NH3 solution in order to eliminate Cl (complete elimination was checked with AgNO3

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

203

solution). After evaporation of the solvent in a rotating vessel at about 323 K and reduced pressure, the solid obtained was dried at 373 K and then calcined at 823 K for 8 h. 2.1.2. Preparation of Sn-Mo catalysts The reagents used were the same as those for pure oxides preparation. The catalysts were prepared by precipitating tin hydroxide in excess of urea in aqueous ammonia solution at pH 8.5. After being ltered and washed, the precipitate was mixed with a saturated solution of (NH4 )6 Mo7 O24 4H2 O in adequate amounts so as to obtain different Sn:Mo molar ratios. The solvent was evaporated by using a rotary evaporator at 323 K and at reduced pressure. Then, the solid was dried at 373 K overnight and calcined in air at 823 K for 8 h. The catalysts were referred to by their Sn:Mo atomic ratios. Thus, the following catalysts were prepared: 0:1; 3:7; 1:1; 7:3; 1:0. 2.2. Catalytic test 2.2.1. Selective oxidation of methanol Selective oxidation of methanol was measured under atmospheric pressure in a conventional xed bed reactor equipped with a jacket furnace. The reactor temperature was measured and controlled by a PID controller with a coaxial thermocouple. The reaction temperatures were 433, 448 and 473 K. The feed was a mixture of MeOH, O2 and He at 2.5:1:10.5 molar ratio. The total gas feed rate was 28 sccm. The catalyst (W=1.0 g) was held in place by a glasswool sandwich. The ows of O2 and He were controlled by Matheson mass ow controllers. MeOH was fed from a saturator that was immersed in a water bath. Its temperature was controlled by means of a Lauda RCSR20C cryostat, ensuring temperature variations lower than 0.1 K. Analyses of products were made by on-line gas chromatography (KONIC-3000) with a thermal conductivity detector by using Porapak T and 5 molecular sieve columns. In all cases, formaldehyde (FA), methyl formate (MF), dimethyl ether (DME), water and CO2 were the only measurable reaction products. Other by-products, such as dimethoxymethane (DMM) and CO were not detected. The catalytic measurements were obtained after running the reaction for at least 2 h, when the steady state was essentially reached.

The experiments at different contact times (W/FV ) were carried out with 0.7 g of 7:3 catalysts. The independent variables investigated were molar fractions of methanol, YMe (0.14 and 0.18) and oxygen YO2 (0.07) of feed, weight of catalyst to total gas ow rate ratio, W/FV (0.9, 1.8 and 2.7 g cat s cm3 ) and reaction temperatures, T (453 and 463 K). 2.2.2. Catalytic evaluation feeding FA Measurements were performed with 7:3 catalyst at 453 K using the different feed conditions listed as follows: FA; FA+O2 ; FA+O2 +MeOH; FA+MeOH; MeOH+O2 ; MeOH. 2.2.3. Catalytic activity on catalysts subjected to pre-treatments The 7:3 catalyst was subjected to different reducing conditions using MeOH and H2 and also to oxidizing conditions. The total gas ow rate was 30 sccm and pressure was 93.3256 kPa. The working temperature was 453 K. The solid was subjected to the following conditions: (1) oxidizing conditions: MeOH+O2 and (2) reducing conditions: (i) A 5 vol.% H2 in He was fed for 15, 30, 45, 60 and 90 min, respectively; (ii) A 18 vol.% MeOH in He was fed for 60 min. In both cases after the pre-treatment, He was owed for 5 min and the reaction was studied under the usual conditions 2.3. Catalyst characterization 2.3.1. Surface area Specic surface areas (SBET , m2 g1 ) of all samples were determined from nitrogen adsorption isotherms at 77 K by the BET method. A Micromeritics Accusorb 2100 E was used. 2.3.2. X-ray diffraction (XRD) XRD patterns were obtained by using a Rigaku diffractometer operated at 30 kV and 20 mA by employing Ni-ltered Cu K radiation (=0.51418 nm). 2.3.3. Decomposition of isopropyl alcohol (IPA) Decomposition of IPA was used for determining the acidbase properties of the samples. The reaction was carried out at 363 and 373 K in a xed-bed continuous ow reactor under atmospheric pressure. The feed consisted of 4.5% IPA and the balance helium.

204

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

The ow rate was 40 sccm and the conversion of IPA was <15% in all experiments. Product analysis was performed by gas chromatography using a Carbowax 20 M on Chromosorb W column and a thermal conductivity detector. 2.3.4. Temperature programmed reduction (TPR) TPR studies were performed in a conventional unit. The apparatus consisted of a gas handling system with mass ow controllers (Mathesson), a tubular reactor, a linear temperature programmer (Omega, model CN 2010), a PC for data retrieval, a furnace and various cold traps. Samples of ca. 80 mg were rst oxidized in a 30 sccm ow of 20 vol.% O2 in He at 600 C for 30 min and then, cooled at room temperature. After that, helium was admitted at room temperature to remove oxygen. The samples were subsequently contacted with a 30 sccm ow of 4.5 vol.% H2 in N2 and heated, at a rate of 10 K min1 , to a nal temperature of 1000 K, while hydrogen consumption was monitored by a thermal conductivity detector after removing the water formed. 2.3.5. Electron paramagnetic resonance (EPR) The EPR measurements were obtained with a Bruker spectrometer at room temperature and a Klystron frequency of 9.7 GHz and 100 kHz magnetic eld modulation. All the catalysts were evaluated prior and after being subjected to reaction conditions. The 7:3 catalyst was also evaluated by EPR after being pre-treated at 453 K with: (i) 5% H2 /N2 mixture, 40 sccm (15 min); (ii) 18% MeOH/He mixture, 40 sccm (15 min); (iii) 0.18:0.07:0.75 MeOH/O2 /He mixture, 40 sccm (30 min); (iv) 0.18:0.07:0.75 MeOH/O2 /He mixture, 40 sccm (30 min) and then treated with 5% H2 /He. 2.3.6. X-ray photoelectron spectroscopy (XPS) XPS spectra were obtained with a Shimadzu spectrometer employing a Mg K X-ray excitation source (h=1253.6 eV). The sample, as ne powder, was pressed. The reported binding energies were referenced to C1s at 284.6 eV. Referencing was performed both before and after a complete set of spectra had been obtained. Ratios of atomic concentration in the outer layers of the samples were expressed as the corresponding XPS area ratios by using the effective ionization cross section of ejected electrons tabu-

lated by Scoeld and the formulas given by Seah and Dench [12]. 2.3.7. Temperature programmed desorption (TPD) TPD was performed with a dynamic apparatus and detection by thermal conductivity connected to a data acquisition system, using NH3 as probe molecule.

3. Results 3.1. Surface area (SBET ) The specic surface areas (m2 g1 ) obtained for the different catalysts are shown in Table 1. The SBET exhibited by co-precipitated catalysts was greater than that of pure MoO3 , independent of the tin load. The surface area increased as the Sn:Mo ratio increased from 3:7 to 7:3, becoming closer to that of SnO2 . 3.2. X-ray diffraction X-ray diffractograms are shown in Fig. 1. For all catalysts, only two phases were observed, corresponding to orthorhombic MoO3 and SnO2 (cassiterite). When the amount of MoO3 was increased from 7:3 to 3:7, the intensity ratio of the MoO3 -SnO2 peaks was higher and crystallinity of SnO2 decreased. 3.3. Temperature programmed desorption The results for the different catalysts are shown in Fig. 2. The spectrum exhibited two well-dened peaks. Total acidity could be calculated by integrating the area under the peaks. In this case, the following order was found 1:1>3:7>7:3. However, the area corresponding to the rst peak would correspond
Table 1 Specic surface area (SBET ) of fresh 0:1, 3;7, 1:1, 7:3 and 1:0 catalysts Sn:Mo 0:1 3:7 1:1 7:3 1:0 SBET (m2 g1 ) 1.74 14.36 21.53 30.75 27.39

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

205

Fig. 1. X-ray diagram of fresh 3:7, 1:1 and 7:3 catalysts. To compare: ( ) SnO2 (JCPDS 21-1250) and () MoO3 (JCPDS 5-508).

Fig. 2. TPD of NH3 of fresh catalysts: () 3:7; ( ) 1:1 and () 7:3.

206

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

Fig. 3. TPR of fresh catalysts: () 3:7; ( ) 1:1; () 7:3; ( ) SnO2 and () MoO3 .

to physisorption (up to 400 K). Another aspect that has to be considered in the interpretation of ammonia desorption experiments, is the possible ammonia decomposition on the solid surface, which can mask the ammonia TPD data. Niwa et al. [9], observed a N2 signal by mass spectroscopy at ca.700800 K during TPD of NH3 from pure SnO2 and MoO3 supported on SnO2 catalysts with less than 4 wt.% of MoO3 . During TPD of catalysts with 4 wt.% of MoO3 or more, the N2 signal was not found. Studying a mixed (massive) MoO3 -SnO2 catalyst, the same authors did not observe the N2 signal, but a new desorption peak between 600700 K was observed. This peak was assigned to acidic sites of MoO3 , which are present only when MoO3 is in close contact with SnO2 , because pure MoO3 does not show this desorption peak. As a consequence, in this study, the acidity was evaluated by integration of the ammonia desorption peaks between 400 and 700 K, showing the order 1:1 =3:7>7:3. 3.4. Temperature programmed reduction The TPR results (Fig. 3) showed that the binary catalysts (SnO2 -MoO3 ) have enhanced reducibili-

ties compared with the pure oxides, which showed very low reducibilities. The reduction temperature of the main peak was different for each catalysts, the higher being the one corresponding to the 3:7 catalyst (850 K). Catalyst 1:1 exhibited a value of 843 K while that of Sn:Mo=7:3 was 808 K. It can be observed that there appears to be a peak with lower reduction temperature (around 700 K) common to all catalysts. Also, catalyst 7:3 exhibited a shoulder at 643 K. 3.5. Electronic paramagnetic resonance The results obtained by EPR for the different catalysts prior (fresh) and after being subjected to reaction conditions (used) are shown in Fig. 4. The spectra of the used catalysts showed a resonance signal with an average g value of 1.925, corresponding to paramagnetic species attributed to Mo5+ ions. However, it can be observed that fresh catalysts exhibited a different g value (1.96), which could indicate an axial symmetrical environment. Used catalysts showed a sharper signal, which would indicate lower anisotropy in the system caused by a reordering of the ions Mo5+ during the reaction. In the used 3:7 catalyst, a signal with a g

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

207

could be compared from their respective intensities. Thus, according to pre-treatment, the intensities followed the order (MeOH+O2 )>MeOH 2 . These =H results indicate that the reacting mixture that included oxygen generated a more generalized reduction from Mo6+ to Mo5+ . It cannot be discarded that when strong reducing agents such as H2 were used, Mo6+ was reduced to Mo4+ (MoO2 ), giving a lower Mo5+ signal. 3.6. X-ray photoelectron spectroscopy The results are presented in Table 2. The binding energies corresponding to Mo 3d5/2 were lower in all mixed catalysts compared to pure molybdenum oxide. The difference of 1 eV, might indicate the presence of Mo5+ in addition to Mo6+ . In fact, the line widths of the peak corresponding to Mo 3d5/2 exhibited a small shoulder that might indicate the existence of more than one species. The binding energy for Mo 3d5/2 in the mixed catalysts was independent of the catalyst composition and did not change signicantly after catalytic test. The binding energy values of Sn 3d5/2 were also independent of the catalyst composition and did not change after catalytic test, being similar to those obtained for pure tin oxide. The surface atomic ratios did not match the expected value from the bulk composition. The surfaces of the 1:1 and 3:7 catalysts showed enrichment in Sn when compared to the starting materials. However, overall, the surfaces are still enriched in Mo, compared to the bulk.
Table 2 Binding energies and Sn:Mo atomic surface ratios of fresh and used 3:7, 1:1, and 7:3 catalysts, and binding energies of fresh MoO3 and SnO2 , from XPS measurements Catalyst Mo:(Mo +Sn) bulk Binding energies (eV) Mo 3d5/2 Sn 3d5/2 7:3 Fresh 7:3 Used 1:1 Fresh 1:1 Used 3:7 Fresh 3:7 Used MoO3 SnO2 0.30 0.30 0.50 0.50 0.70 0.70 232.0 232.1 232.2 232.0 232.1 232.0 233.1 486.8 486.8 486.8 486.9 486.9 486.8 486.9 1.24 1.20 0.18 0.48 0.18 0.34 2.33 2.33 1.00 1.00 0.42 0.42 Sn:Mo surface Sn:Mo bulk

Fig. 4. EPR spectra of fresh and used catalysts: () 3:7; ( ) 1:1 and () 7:3.

average value of 2.0017 assigned to O species was observed [13]. In a previous study on mechanical mixtures, an increase in the concentration of Mo5+ species was observed after the catalyst was subjected to the reacting atmosphere. It was attributed to the MeOH reaction or to some of the intermediates on the catalyst surface [1]. Catalyst 7:3 was subjected to a series of oxidizing and reducing pre-treatments and subsequently characterized by EPR. The results obtained are shown in Fig. 5. In all the spectra, a resonance signal with a g value of 1.925 was observed, which is attributed to an unpaired Mo5+ electron. The spectra were normalized on the base of sample weight at equal measurement conditions. Taking into account that the bandwidths are similar, the concentrations

208

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

Fig. 5. EPR spectra of 7:3 catalyst after different treatments: () H2 ; ( ) MeOH and () MeOH+O2 .

3.7. IPA decomposition test This test gives information about the surface acidbase properties of the catalysts. The IPA decomposition proceeds by two parallel routes: dehydration to propene on acidic sites and dehydrogenation to acetone on redox (basic sites). This reaction cannot distinguish between the Brnsted and Lewis sites. Its

advantage lies in the fact that it can be applied to low specic surface area catalysts for which the spectroscopic techniques cannot provide reliable data [14]. The propene and acetone production rates together with their respective ratios are shown in Table 3. Assuming that propene formation rate is a measure of the surface acidity, as suggested by Ai [4], then the acidity order of the catalysts is as follows: 1:1 =3:7>7:3.

Table 3 Isopropyl alcohol decomposition on catalysts 0:1, 3:7, 1:1, 7:3 and 1:0 Treaction (K) Sn:Mo 7:3 363 373 363 373 363 373 Rates of propene production 1.02105 1.28105 (mol m2 1:1 1.97105 3.07105 min1 ) 3:7 1.84105 2.70105 3.44106 5.97106 1.12104 1.18104

Rates of acetone production (mol m2 min1 ) 2.46106 4.23106 4.65106 5.02106

Rate of acetone production to rate of propene production ratio 0.24 0.21 0.19 0.36 0.16 0.22

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

209

If the activation energy of the propene formation is considered as a measure of surface acidity [15], the same trend is found. The ratio of dehydrogenation rate (to acetone) to the dehydration rate (to propene) has been regarded [4] as a measure of basicity. Assuming this correlation, the basicity of the catalysts shows a reverse order to that found for acidity. 3.8. Catalytic activity The results of methanol conversion (X) and selectivities to MF, FA, DME and CO2 as a function of reaction temperature for the different catalysts are shown in Fig. 6. As expected, all the catalysts showed strong increase of conversion with temperature. Irrespective of the catalysts, the MF and FA selectivities showed a slight dependence with the temperature, while CO2 selectivities clearly increase as the temperature increases, accompanying the decrease of DME selectivities. It is worth to mention that the 7:3 catalyst showed the highest conversion (64% at 498 K) with MF and FA selectivities in the order of 48 and 21%, respec-

tively. This catalyst also showed the lowest selectivities to DME and CO2 (16 and 12%, respectively). The results of catalytic evaluation of Sn:Mo=7:3 at different contact times (W/FV ) are shown in Table 4. For a MeOH molar fraction (YMeOH ) of 0.14, it was observed that at low W/FV values, selectivity to FA was the highest, while at elevated W/FV values, the main product was MF at the expense of FA. When working with YMeOH of 0.18, a decrease of conversion was observed. However, the moles of methanol converted were slightly smaller than in the case of YMeOH 0.14. The reason could be attributed to a lower YO2 :YMeOH ratio present when YMeOH of 0.18 was used. Under all conditions (except for W/Fv 0.9 g s cm3 and YMeOH 0.14), the main product was MF. Table 5 shows the results obtained by feeding reagents and products in different orders. By feeding only FA, a small amount of MF was observed, which might come from a dimerization reaction of FA by a Tischenko type mechanism. This effect was immediately suppressed by incorporation of O2 into the feed. The subsequent incorporation of MeOH caused high

Fig. 6. Catalytic activity of 3:7, 1:1 and 7:3 catalysts: ( ) MeOH conversion; ( ) FA selectivity; ( ) MF selectivity; ( ) DME selectivity and ( ) CO2 selectivity.

210

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

Table 4 Catalytic activity of catalyst 7:3 as a function of W/FV (g cat s cm3 ) at 453 and 463 K W/FV 2.7 1.7 0.9 2.7 1.7 0.9 2.7 1.8 0.9 2.7 1.9 0.9
a

T (K) 453 453 453 453 453 453 463 463 463 463 463 463

YMeOH :YO2 0.18:0.07 0.18:0.07 0.18:0.07 0.14:0.07 0.14:0.07 0.14:0.07 0.18:0.07 0.18:0.07 0.18:0.07 0.14:0.07 0.14:0.07 0.14:0.07

Xa 38.135 27.98 15.61 52.56 39.32 22.12 60.026 47.354 30.764 81.04 69.38 45.48

SFA 27.24 27.8 32.04 29.38 35.69 42.98 28.80 32.15 36.376 33.35 33.68 42.823

SDME 24.24 24.07 23.33 22.33 22.35 26.22 20.09 19.49 20.042 22.81 19.98 19.14

SFM 47.00 47.34 44.58 47.15 41.96 30.79 49.55 46.15 43.51 41.045 44.95 37.87

SCO2 1.5 0 0 1.06 0 0 1.57 1.21 0 2.79 1.40 0

X correspond to conversion of methanol.

activity and more MF production, which was proportional to the MeOH concentration in the feed. This might indicate that MeOHFA constitutes a probable route of MF formation. When only MeOH was fed, no activity was observed and consequently there was no DME formation. DME might form via dehydration on acid sites. This might indicate that, after treatment with methanol, the acid sites on the surface of these catalysts were poisoned or destroyed, or became too weak to produce MeOH dehydration. Oxygen in the feed is necessary for DME production. By feeding MeOH and FA without the presence of O2 , a decrease in conversion with time on stream was observed with an accompanying decay in MF yield. The catalytic activity studies of the catalyst 7:3 subjected to different pre-treatments under reducing conditions with MeOH and H2 and under oxidizing conditions can be summarized as follows. When the

catalyst was treated with H2 prior to MeOH ODH, the initial activity was high, decreasing with operation time until reaching constant values. The reducing atmosphere (H2 ) might have favored the formation of catalytically active sites, which is shown by the elevated initial activity of the catalyst. The methanol pre-treatment did not show any observable effect on the methanol ODH. 4. Discussion The X-ray diffractograms only show the crystalline phases corresponding to SnO2 and MoO3 . No new XRD peak was observed, indicating the absence of denite Sn-Mo mixed oxide compounds. However, by comparing the XRD intensities of the mixed catalysts with pure oxides MoO3 and SnO2 , a clear decrease of SnO2 crystallinity was observed as molybdenum loading was increased. For catalyst 7:3, the MoO3 :SnO2 intensity ratio fell below that expected by a linear calculation based on its relative atomic ratio. Baares et al. [16] note that after catalytic evaluation, higher MoO3 crystallization might occur. The changes exhibited by binary oxides with respect to pure oxides indicate that this catalyst is more complex than a simple mixture of oxides. In fact, the changes in the surface area of catalysts SnO2 -MoO3 with composition do not follow a linear trend. The TPR data showed that the reducibilities of the pure oxides were very low, which differ from the data

Table 5 Catalytic activity of catalyst 7:3 at 453 K using different feed conditions Feed FA FA+O2 FA+O2 +MeOH FA+MeOH MeOH+O2 MeOH Results MF is produced (Tischenko reaction) MF production is not observed High MeOH conversion and high MF yield The MeOH conversion and the MF yield decrease proportionally The observed products are MF (the highest yield), FA and DME DME production is not observed

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

211

reported by Niwa et al. [9]. These authors observed reduction of pure SnO2 above 663 K. The TPR of the co-precipitated catalysts exhibited large peaks of hydrogen consumption, indicating that their reducibilities are strongly enhanced compared with the pure oxides. The maximum reduction temperature observed is different for the three catalysts and follows the order: T7:3 <T1:1 <T3:7 . For catalyst 7:3, a marked decrease of reduction temperature was observed. Three temperature maxima can be distinguished, at 643, 700 and 808 K, all of which are lower than the single maximum exhibited by catalyst 3:7 at 850 K, which is close to that obtained for catalyst 1:1 (843 K), but with a lower amount of consumed hydrogen. According to these observations, the strength of the metaloxygen bonds in these three catalysts should be signicantly different. The TPR results indicate that the Mo load determines the system reducibility, and the combination of the tin and molybdenum oxides confers to the system different characteristics from those of pure phases separately. From these TPR data, it is not possible to determine when the MoO3 reduction nished and the SnO2 reduction began. Moreover, it is probable that in these massive catalysts both MoO3 and SnO2 undergo reduction simultaneously over a wide temperature range. In addition, the reduction of these massive catalysts, which are characterized by their low surface areas, is far to be quantitative. Taking into account that these catalysts are evaluated in the selective oxidation of methanol at temperatures at 500 K or below, the more valuable information is obtained from the reduction trends observed below 700750 K. From these data it is clear that the reducibilities of the massive (co-precipitated) catalysts are strongly enhanced respect to the pure oxides, having the order 7:3>1:1>3:7. This is consistent with the XRD results that indicate a strong Sn-Mo interaction. Niwa et al. [9] noted that when MoO3 is impregnated on alumina, observation by XAFS reveals that MoO3 keeps its original structure and the lengths of the MoO bonds of the MoO4 unit are 0.167, 0.173, and 0.195 nm. When the support is SnO2 , MoO3 loses the 0.195 nm bond. Since this is a very active catalyst, it could be thought that such activity is due to its structure. In a previous work [1], we stated that SnO2 shows high deep oxidation activities, while MoO3 is selective to products of partial oxidation although its

activity is relatively low. Combination of these component oxides leads to higher activities and selectivities for partial oxidations as observed for other mixed oxide catalysts. One rst conclusion to be drawn from the SBET , XRD and TPR results, is that interdispersion between MoO3 and SnO2 favors a supercial architecture in which the Sn-Mo interaction plays a major role. There exists a probable preferential ordering by close contact between phases. Further characterizations are required in order to more accurately describe these interactions. The EPR results indicate the presence of Mo5+ in the fresh catalysts, probably produced by the high calcination temperature [17]. This observation would agree with reports by Niwa et al. [9] on the formation of Mo5+ when molybdenum is impregnated on Sn. The spectra obtained by EPR both for fresh and used samples indicate that the intensities of the signals attributed to paramagnetic species are similar, exhibiting an average g of 1.925 assigned to Mo5+ . However, there is a clear difference between fresh and used catalysts. In the former, Mo5+ appears to be in an axial symmetrical environment, since two g values can be identied. In used samples, on the other hand, the spectra show a single g value, indicating lower anisotropy related to a possible reordering of the coordination sphere of surface Mo5+ . Although EPR is a bulk technique, any observation after the test reaction indicates changes in the surface, in which the reaction occurs. The reordering of surface atoms might inuence product distribution in a structure-sensitive reaction. Mo5+ might be part of the Mo6+ arrays in its crystalline structure. The existence of the coupled pair Mo5+ Mo6+ , presumed necessary to maintain activity, reveals a favorable balance to the reduced state. It is surprising to observe from the EPR results that, after H2 treatment, catalyst 7:3 presents a lower concentration of Mo5+ species than the one subjected to reaction (MeOH+O2 ). Pure MoO3 does not suffer an appreciable reduction with H2 neither does it exhibit signal by EPR. Thus, it can be inferred that SnO2 facilitates that reduction. Additionally, the EPR spectra of catalyst pre-treated with MeOH and O2 and then treated with H2 , showed that the Mo5+ signal remains almost constant. This might indicate that, in spite of the treatment with hydrogen, Mo5+ , which is produced under reaction conditions, is very stable to subsequent reductions even under 5% H2 /N2 mixture.

212

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

The lower intensity of the EPR signal for 3:7 catalyst might indicate its greater stability to reduction, or its reduction to Mo4+ , which does not give an EPR signal. The signal corresponding to Sn3+ was not observed in any of the catalysts. According to Niwa et al. [10], the formation of MF is signicant for Mo loadings on SnO2 under 3%, while loadings of 4% or above produce almost exclusively FA. It could be supposed that the catalysts studied in this paper which show signicant MF yields, could have similar supercial architecture to that of until 3% Mo loadings. It must be made clear that this might not be so in terms of strictly dened compositions and that it is not easy to compare data obtained from methanol ODH, since this reaction is very sensitive to operation conditions. The partial conclusions derived from the EPR results might be summarized as follows: In Sn-Mo catalysts, at low Mo loadings, the Mo5+ produced under reaction conditions is very stable to subsequent reduction even under severe conditions. Sn facilitates reduction. At high Mo loadings, reduction is more difcult, and when reduction occurs, it might be to Mo4+ . The rst conclusion stated above would indicate the existence of a particular architecture at atomic level. With the help of XPS, we attempted to account for this surface change which does not appear to involve a solid state reaction but substantially modies the physicochemical characteristics of the surface. The binding energies are not modied in any of the analyzed catalysts after being tested in the methanol ODH reaction. The binding energy values for Sn obtained by XPS were independent of the catalyst composition and matched the values for pure oxide SnO2 within a 0.2 eV range. For Mo 3d5/2 , however, even though the binding energy values are also independent of composition, they are 1 eV below those of pure molybdenum oxide. All the catalysts showed a Sn:Mo surface atomic ratio lower than those of the bulk, indicating a surface enrichment of Mo. After catalytic test, the only surface atomic ratio, which remains practically unaltered, is that of 7:3. Catalysts 1:1 and 3:7, on the other hand, suffer major changes increasing the Sn:Mo surface atomic ratio. The Sn:Mo surface atomic ratio of used 3:7 catalyst is relatively close to that of bulk Sn:Mo atomic ratio. Okamoto et al. [18] observed a similar behavior and attributed it to an exclusion effect due to MoO3

crystallization. Our results for Sn:Mo surface atomic ratio obtained by XPS for used catalysts are close to those obtained by Okamoto et al. [18] for fresh catalysts. Line widths of level Mo3d show a small shoulder, which would indicate a slight reduction of Mo6+ during the reaction. This phenomenon was not observed on the tin spectrum. In brief, the catalyst surface is strongly altered by the test reaction (methanol ODH), which induces a supercial enrichment of Sn without reaching the bulk values. The shift in the maximum of BE Mo 3d5/2 for the binary catalysts with respect to pure oxides (from 233 to 232 eV) would indicate the formation of an intermediary between Mo6+ and Mo4+ , which could be Mo5+ . This one can be actually observed by EPR. We found that at Mo:(Mo+Sn)=0.5 atomic ratios, the characteristics are similar to those reported by Okamoto et al. [18] for Mo:(Mo+Sn)=0.7 catalysts. Probably, differences in the preparation method contribute to the formation of a more crystalline MoO3 in close contact with SnO2 . However, it can be postulated that the reaction of methanol ODH plays a major role in the redenition of the catalytic surface nal architecture, as indicated by the changes in the surface atomic ratio measured by XPS, after catalytic test. It has been observed [18] that Sn suffers a chemical shift of at least 0.4 eV when it passes from Sn4+ to Sn2+ . A further partial conclusion to be proposed is that since no chemical shifts are detected for Sn (XPS), the bond of oxygen to molybdenum may be the main specie responsible for alcohol dehydrogenation. According to Niwa et al. [9], the MoO bond in MoO3 supported on SnO2 and on Fe2 O3 , is weaker than in TiO2 , ZrO2 and Al2 O3 supported MoO3 . The strength of the MoO bond can be inuenced by the metaloxygen bond of the support. This suggests that there may be oxygen exchange between the supported oxide and the support, or else, a fast replacement in the reduced site from the support [19]. It can be postulated that oxygen diffuses not in the SnO2 lattice but in the boundary between SnO2 and MoO3 . Mechanical mixtures of Sn and Mo oxides showed higher activities and selectivities than the pure oxides [1]. In these systems, the possibilities of Mo dissolution in the Sn lattice are low. It is evident that the results of this work cannot be interpreted independently of those obtained for mechanical mixtures. According to the above comments, the strength of the

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214

213

metaloxygen bonds in the three catalysts here studied are different, which results in different catalytic behaviors. If impregnated catalysts were used, the reducibility observed in the MoO3 on monolayer could be correlated with its structure, since it can be considered as a distorted form of MoO3 species, which can be identied as reducible due to their low co-ordination number. However, in these co-precipitated catalysts, molybdenum is probably included in the tin oxide lattice, favoring the subsequent stability of the formed Mo5+ ions. For Mo loadings above Sn:Mo=7:3, Mo would segregate forming crystalline MoO3 and giving a more acidic catalyst capable of activating methanol by mechanisms other than ODH. In fact, it is clear the increase of the DME yield in 1:1 and 3:7 catalysts. The DME is produced by dehydration in acid sites. These results are coherent with the higher acidity observed by the IPA test reaction and ammonia TPD in these catalysts. Considering that MeOH is a molecule with a highly electronegative oxygen atom and, therefore, exhibits well differentiated electron donor and acceptor characteristics, C+ O H+ , it is susceptible to be activated by acid and basic sites of metallic oxides. Activity and selectivity could be described in terms of the acidbase properties of the catalyst, considering that the combination of the metal oxides contributes to the modication of the acid properties. The decomposition of IPA indicates that the catalyst possesses associated acid and basic sites and isolated acid sites From these results any clear correlation with methanol conversion or with the product distribution in the methanol ODH was found, except for the above mentioned of DME yield. Finally, in an attempt to elucidate the behavior of the catalytic system under study, a series of tests were performed for catalyst Sn:Mo=7:3 with different feed sequences of reagents and products. By feeding FA, it was observed that dimerization practically does not contribute to MF formation and that MeOH is necessary for its formation. However, the presence of O2 notably increases the yield to MF. The formation of MF would take place by reaction in series, from FA with methoxy species contributed by MeOH, without discarding a small contribution by FA dimerization (Tischenko). This is corroborated by the activity measurements for different W/FV , since at higher W/FV values MF selectivity is increased at the expense of FA. MeOH or some of its derivatives

(FA) would be responsible for producing catalytically active sites (Mo5+ ) and impeding their reoxidation by O2 . According to the strength of the sites, FA would be retained to form an intermediary that would nally be released as MF. It can be concluded that in catalysts with low Mo content, oxygen is necessary for the MF formation mechanism.

5. Conclusions The ODH methanol reaction induces a reordering of the catalyst structure (EPR). In fact, the reaction plays a major role in redening the nal architecture of the catalytic surface, as indicated by the changes in the supercial atomic ratio measured by XPS, after catalytic test. Both the results from sequential feed of reagents and/or products and from EPR after MeOH and O2 treatment would indicate that the reaction is involved in regeneration of Mo5+ centers. The close combination of SnO2 and MoO3 modies the reactivity of the lattice oxygens and the reducibility of Mo ions, and therefore, the catalytic properties. The synthesis method here used results in high dispersion of Sn and Mo oxides. The interaction between both elements and their relative proportions determine the crystalline ordering. Molybdenum is probably included in the tin oxide lattice favoring the subsequent stability of the Mo5+ ions formed. When high quantities of molybdenum are used, the segregation of MoO3 both during the catalyst preparation process and after the reaction modies the product distribution due to changes in the reaction mechanisms. The surface atomic ratios of fresh catalysts did not match the expected value from the bulk composition, showing in all cases Mo enrichment. The surfaces of the 1:1 and 3:7 catalysts, after methanol ODH, showed enrichment in Sn when compared to the starting materials. However, overall, the surfaces are still enriched in Mo, compared to the bulk. The absence of chemical shifts for Sn (XPS) suggests that the OMo bond is the main responsible for alcohol dehydrogenation. The possible surface transformation of the catalyst by the reaction constitutes the main limitation of

214

N.G. Valente et al. / Applied Catalysis A: General 205 (2001) 201214 [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] M. Ai, T. Ikawa, J. Catal. 40 (1975) 203. M. Ai, J. Catal. 40 (1975) 318. M. Ai, J. Catal. 40 (1975) 327. L.-T. Weng, B. Delmon, Appl. Catal. A 81 (1992) 141. J.M. Tatibout, Appl. Catal. A 148 (1997) 213. M. Ai, J. Catal. 77 (1982) 279. E. Weng, E. Wolf, Appl. Catal. A 96 (1993) 383. M. Niwa, M. Sano, H. Yamada, Y. Murakami, J. Catal. 151 (1995) 285. M. Niwa, H. Yamada, Y. Murakami, J. Catal. 134 (1992) 331. Z. Sojka, M. Che, J. Phys. Chem. 99 (1995) 5418. M.P. Seah, W.A. Dench, Surf. Int. Anal. 1 (1979) 2. M.B. Ward, M.J. Lin, J.H. Lunsford, J. Catal. 50 (1977) 306. B. Grzybowska-Swiekosz, Appl. Catal. A: Gen. 157 (1997) 263. A. Gervasini, A. Auroux, J. Catal. 131 (1991) 21. M. Baares, H. Hu, I.E. Wachs, J. Catal. 150 (1994) 407. W. Oganowski, Y. Hanuza, B. Jezowska-Trzebiatowskaa, Y. Wrzyszcz, J. Catal. 39 (1975) 161. Y. Okamoto, K. Oh-Hiraki, T. Imanaka, S. Teranishi, J. Catal. 71 (1981) 99. Y. Moro-oka, Catal. Today 45 (1998) 3.

the use of methanol oxidation as a catalytic surface probe. Acknowledgements The authors are grateful to Japan International Cooperation Agency for the grant of the X-ray Photoelectron Spectrometer to CENACA. Financial support is also acknowledged to the Consejo Nacional de Investigaciones Cientcas y Tcnicas of Argentina (CON ICET), Agencia Nacional de Promocin Cientca y Tecnolgica (ANPCYT) and Universidad Nacional de San Luis. References
[1] N.G. Valente, L.E. Cads, O.F. Gorriz, L.A. Arra, J.B. Rivarola, Appl. Catal. A 153 (1997) 119.

[19]

You might also like