Vous êtes sur la page 1sur 16

Applied Catalysis B: Environmental 22 (1999) 91106

Selective catalytic reduction of NO with C3 H6 over Rh/alumina in the presence and absence of SO2 in the feed
Evangelos A. Efthimiadis , Sophia C. Christoforou, Apostolos A. Nikolopoulos, Iacovos A. Vasalos
Chemical Process Engineering, Research Institute and Department of Chemical Engineering, Aristotelian University of Thessaloniki, P.O. Box 1517, 54006 University City, Thessaloniki, Greece Received 25 December 1998; received in revised form 5 March 1999; accepted 4 April 1999

Abstract The selective catalytic reduction of NO in excess O2 with C3 H6 as reductant was studied over a Rh/alumina catalyst. The effect of SO2 on the extent of NO reduction and product formation was extensively examined. TPD and FT-IR studies using different feeds aimed at the identication of species adsorbed on alumina and Rh/alumina samples. Experiments were also carried out in a differential xed-bed reactor to estimate apparent orders of reaction and activation energies. The ratecontrolling step of the proposed reaction mechanism for the NO reduction is the reaction of RhNCO and RhNO+ for an SO2 -free feed. In the presence of SO2 in the feed the NO and C3 H6 oxidation over the catalyst were inhibited and a less complicated reaction mechanism is proposed, based on the reaction of RhNO+ and partially oxidized C3 H6 . 1999 Elsevier Science B.V. All rights reserved.
Keywords: Nitric Oxide; Propene; Rhodium; Alumina; SCR of NO; Reduction; SO2 effect; Mechanism

1. Introduction The selective catalytic reduction (SCR) of nitrogen oxides (NOx ) in the presence of excess O2 can be applied in the removal of gaseous pollutants from diesel and lean-burn engines. Depending on the nature and the preparation process of the fuel, sulfur may exist in exhaust streams in the form of SO2 . As a result, the tolerance of a deNOx catalyst to SO2 is important for its use in commercial applications.

Corresponding author. Tel.: +30-31-996-175; fax: +30-31-996184; e-mail: efthimia@alexandros.cperi.forth.gr

Catalysts used in the SCR of NOx can be roughly classied based on their support material to zeoliteand metal oxide-based. ZSM-5 is among the most extensively studied zeolitic supports. The combination of SO2 and H2 O in the feed was found to cause a signicant loss of deNOx activity over CoZSM5, that was attributed to the blockage of the Co2+ sites by SO2 [1]. This inhibition was not noticed over an overexchanged FeZSM5 catalyst at 500 C [2]. However, an elaborate preparation procedure had to be applied for the synthesis of the latter catalyst that aimed at the minimization of the Fe3+ sites and the inhibition of the formation of Brnsted sites. Tabata et al. [3] compared the durability of activity over Co-Beta, Cu-Beta

0926-3373/99/$ see front matter 1999 Elsevier Science B.V. All rights reserved. PII: S 0 9 2 6 - 3 3 7 3 ( 9 9 ) 0 0 0 4 0 - 5

92

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

and Co-ZSM5 catalysts using a model feed containing 0.3 ppmv SO2 . The rst catalyst was the only one that exhibited stable activity for about 4000 h under realistic reaction conditions. The promising performance of Co-Beta was attributed to the pore structure of Beta zeolite and the stable Co dispersion. The tolerance of metal oxide-based catalysts to the SO2 presence in the feed depends on the type and oxidation state of the deposited metal, and the nature of the support. In our earlier work [4] we classied the metal/-alumina catalysts in terms of their NOx reduction activity to those that are deactivated (e.g. transition metals on -alumina), not affected (Pt/alumina) and enhanced (Rh/-alumina) by the presence of SO2 in the feed. The poisoning effect of SO2 over Cu/Al2 O3 was noticed by Okazaki et al. [5], while the presulfation of Pt/Al2 O3 had minimal effect on the deNOx activity of the catalyst [6]. Zhang et al. [7] measured an enhancement in the deNOx activity over Pt/BSA (BSA: B2 O3 SiO2 Al2 O3 ) after the addition of 100 ppmv SO2 to the feed. Presulfation of -alumina caused a small decrease in the nitrogen yield when the reductant was CH3 OH [8]. On the contrary, almost no reduction to N2 was measured over the same catalyst when the reductant was C3 H8 [6]. The diversity in the deNOx activity over the presulfated samples was attributed to the difference in the adsorption capacity of the two reductants. Recently, Sumiya et al. [9] examined the effect of H2 O (10%) and SO2 (30 ppmv) on the activity of Ag/alumina mounted on a cordierite. They noticed a suppression of the NOx reduction activity over a temperature range of 250470 C that was attributed to the formation of sulfates on the catalyst. The reversible improvement in the deNOx activity of Rh/alumina when we added SO2 in the feed (NO/C3 H6 /O2 /He) [4] was our motive to examine any possible modications in the reaction mechanism of the SCR of NO in the presence and absence of SO2 in the feed. We applied surface characterization techniques to identify species adsorbed on the catalytic surface under different reaction conditions and we proposed different reaction schemes for SO2 -free and SO2 -contaning feeds. Our experimental results showed that the oxidation sites of Rh/alumina are signicantly affected by the presence of SO2 in the feed, thus, we also studied the effect of SO2 on the oxidation of NO over the same catalyst.

2. Experimental 2.1. Materials We used -alumina extrudates supplied by Engelhard (sample code: Al-3992 E 1/8) as the support of our catalysts. The extrudates were crushed and sieved to separate the particles of 180355 m. An aqueous solution of rhodium chloride was prepared from the dilution of pure RhCl3 3H2 O (38% Rh, Merck) for the catalyst formation. This solution was impregnated on the -alumina particles by applying the dry impregnation technique. The catalyst was dried at 120 C for 2 h and then calcined at 600 C for 9 h under air ow. Rh/alumina in the particle form was used in the TPD, FT-IR and activity studies. The following gas mixtures were mixed to prepare the feed in the TPD, FT-IR and activity studies: 2% NO/He, 2% C3 H6 /He, 20% O2 /He, 1% SO2 /He, He. During the catalyst preparation procedure we used synthetic air (20% O2 /N2 ). 2.2. Catalyst characterization We used the inductively coupled plasma and atomic emission spectroscopy (ICP/AES), the X-ray Photoelectron Spectroscopy (XPS), the H2 chemisorption, the N2 adsorption and the mercury porosimetry techniques to characterize our Rh/alumina catalyst. The ICP/AES analysis was carried out in a Plasma 400 (Perkin Elmer) spectrometer, equipped with Cetac 6000AT+ ultrasonic nebulizer. The XPS spectrum of fresh and post reaction samples was obtained by a SPECS, system LHS10, spectrometer. Mg K radiation (1253.6 eV) was the X-ray source. The Rh dispersion was measured using an Altamira AMI-1 dynamic apparatus, assuming the ratio of surface Rh to the chemisorbed hydrogen atom of 1 to 1. The N2 adsorption data measured in a Autosorb-1 (Quantachrome) apparatus were used in the calculation of the BET surface area. The pore structure of the sample was determined in an Autopore II (Micromeritics) porosimeter. 2.3. TPD procedure Temperature programmed desorption (TPD) studies were carried out in a quartz reactor (17 mm i.d.) loaded

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

93

with 0.5 g (90 mol of Rh) of fresh catalyst. The catalyst was held between two zones of quartz beads, each 1 cm in height. We used mass ow controllers to adjust the ow of the NO/He, C3 H6 /He, O2 /He, SO2 /He and He. The total ow rate was 1000 cm3 /min. The catalyst was initially heated at 320 C under pure helium ow. At this temperature the feed was sent to the reactor for 1 h and it was then cooled down to 100 C. The feed was switched to pure He and the catalyst was exposed to He ow at 100 C for 1 h to desorb the weakly (physically) sorbed species. The temperature was then increased to 600 C using a heating rate of 10 C/min. The exit of the reactor was connected to the gas analysis system described in Section 2.6. The variation of the NOx and SO2 concentrations with the temperature was recorded continuously. The N2 and N2 O concentrations were measured unsteadily and thus, these measurements could not be used in the TPD studies. Preliminary experiments showed no appreciable adsorption of NOx and SO2 on the empty reactor and the pipelines. 2.4. FT-IR procedure Fourier Transform Infrared (FT-IR) studies were performed on a NICOLET 740 spectrometer, equipped with an MCT detector, KBr beam splitter and a high temperature/high pressure (HTHP) environmental chamber (Spectra, 0030-102), designed to treat samples in situ. The cell was equipped with a furnace, a cooling water jacket, a ZnSe window and gas inlet and outlet streams. About 50 mg of the catalyst in powder form were placed in the chamber in each experiment. The sample was pretreated at 500 C for 30 min under 20% O2 in He ow and then cooled to the desired temperature (200 C or 250 C or 300 C). For sulfated samples the pretreatment temperature was lower (420 C) to avoid the decomposition of the sulfates. Absorbance spectra from 64 scans were averaged every 1 or 5 min, at a 4 cm1 resolution. Background spectra were recorded prior to each experiment under pure He ow. The reported spectra were calculated from the difference between the actual and the corresponding background spectra. 2.5. Activity measurements We used the same reactor as in the TPD experiments for the activity studies. The catalyst was ini-

tially heated at 500 C for 1 h under He. The typical composition of the reactive gas was 1000 ppmv NO, 1000 ppmv C3 H6 , 0 or 200 ppmv SO2 , 5% O2 in He. The reactor loading was 0.2 or 2 g and the catalyst weight over ow rate ratio (W/F) was 0.012 or 0.12 g s cm3 . All experiments were carried out under atmospheric pressure. 2.6. Gas product analysis The exit of the reactor was connected to analyzers and a gas chromatograph. The concentration of the NOx (NO and NO2 ) was measured with a chemiluminescence analyzer (Thermo Environmental, 42H). The concentration of CO and CO2 was measured with a dual channel, non-dispersive infrared spectroscopic (NDIR) gas analyzer (Rosemount, NGA 2000). A pulsed uorescent analyzer (Thermo Electron, 40) was used for the SO2 concentration measurements. The GC system (Varian, 3600 CX) was equipped with Thermal Conductivity Detector (TCD) and Flame Ionization Detector (FID). A molecular sieve 13X column was used for the separation of the inorganic species and a HayeSep N column for the separation of the organic species. The analyzers and the GC system were calibrated with gases of standard composition. In the presentation of the experimental results, NOx conversion is dened as the percentage of the inlet NO that is converted to N2 and N2 O, while C3 H6 and SO2 conversions are dened as the percentage of initial C3 H6 and SO2 concentrations that has reacted.

3. Experimental results The compositional analysis of the catalyst using ICP/AES technique showed that Rh/alumina contained 1.86 0.05 wt. % Rh. The dispersion of Rh atoms was about 60% (H2 chemisorption studies). Characterization of the catalyst using the X-ray Photoelectron Spectroscopy technique showed that the fresh catalyst consists of Rh2 O3 . The BET surface area of the fresh catalyst was 182 m2 /g. Mercury porosimetry measurements showed that the overall pore volume of the catalyst was 0.5 cm3 /g and the pore diameters varied in the range of 0.0050.02 m, mainly.

94

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

Fig. 1. Temperature programmed desorption experiment over the support (-alumina). Feed: 1000 ppmv NO in He.

3.1. TPD studies Initially we examined the NO sorption capacity of the support solid (-alumina) in the experiment shown in Fig. 1. The feed was 2000 ppmv NO in He during the sorption stage and pure He during the desorption stage of the experiment. Most of the NO was desorbed at 220 C, while a second smaller peak was detected at 480 C. The overall amount of the desorbed NOx was estimated from the integration of the corresponding curve in Fig. 1 and it was about 35 mol. About 40% of the desorbed NOx was nitrogen dioxide (NO2 ), implying some oxidation of the sorbed NO on the sample. The high percentage of the desorbed NO2 does not necessarily imply that an equally high percentage of the inlet NO is oxidized. It is possible that NO2 is adsorbed on the catalyst more strongly than NO. The NOx released at 480 C was attributed to the decomposition of aluminum nitrate formed during the sorption stage of the experiment. The NOx desorption curves over Rh/alumina are shown in Fig. 2, along with the corresponding ones over alumina. The same reaction conditions were applied in both experiments. The NOx curve over Rh/alumina exhibited a peak at 220 C that was attributed to the support solid (alumina), and a second peak at 380 C that was attributed to the presence of

Rh. The NO2 /NOx mole ratio in both peaks was about 0.35, that is, a value close to the one measured over alumina. Integration of the NOx curve gave the total amount of the desorbed NOx from the Rh/alumina sample (29 mol). This value was equal to the one calculated by the integration of the rst peak over the alumina sample. Therefore, the impregnated Rh did not change the overall amount of the sorbed NOx , but caused a shift of 1/3 of desorbed NOx to substantially higher temperatures. Our results are in agreement with the NO desorption data of Burch et al. [10] over oxidized Pt/Al2 O3 , where three NO peaks were detected at ca. 120 C, 230 C and 330 C. The peaks at the two lowest temperatures were attributed to the NO desoprtion from the sample, while the one at the highest temperature to the aluminum nitrate decomposition. We added 2% O2 to the NO/He sorption mixture and we repeated the TPD experiment (Fig. 2). Over the alumina sample we observed, qualitatively, the same peaks as in the experiment of Fig. 1 (peaks at 220 C and 480 C), though more NOx (50 mol) was desorbed from the porous solid than in the absence of O2 from the feed (35 mol). The presence of oxygen in the feed caused a large increase in the second NOx desorption peak (480 C) that was attributed to the formation of nitrates. The NO2 /NOx ratio (0.45) was slightly higher than that

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

95

Fig. 2. Temperature programmed desorption experiment over the support (-alumina) and Rh/alumina for NO/He and NO/O2 /He feeds.

in the absence of O2 (0.40). The same experiment was carried out over Rh/alumina (Fig. 2). The overall amount of the desorbed NOx was 54 mol (29 mol in the absence of O2 ), the NO2 /NOx ratio in the rst peak (220 C) was 0.47, while the corresponding ratio in the second peak (380 C) was 0.7. The height of the NOx desorption peak at 220 C was about the same in the presence and absence of O2 in the feed. On the other hand, the peak at 380 C was signicantly larger in the presence of O2 . The high percentage of NO2 was attributed to the impregnation of Rh on alumina. In the TPD experiments of Xue et al. [11] multiple NO and NO2 desorption peaks were measured over -alumina and Pt/-Al2 O3 in the temperature range of 250400 C. The supported samples exhibited NOx desorption curves of lower, in general, intensity of the unsupported ones and the authors postulated that the support governs the desorption characteristics of Pt/-Al2 O3 . In Fig. 3 we present the desorption of SO2 from the alumina and Rh/alumina samples. The feed gas was 2000 ppmv SO2 in He during the sorption stage and pure He during the desorption stage of the experiment. When the samples were exposed to the SO2 environment at 320 C we detected a rst peak at 220 C and a second one at 370 C. Larger amounts of SO2 were desorbed from alumina than from Rh/alumina. A possible explanation for these results is that SO2 is ad-

sorbed on the Rh sites and is subsequently oxidized to SO3 in a similar way as for the NO oxidation. SO3 can either desorb dissociatively (formation of SO2 and O2 ) or remain in the gas phase as SO3 , however, it cannot be detected by our analyzer. When the SO2 sorption took place at a higher temperature (550 C) we noticed that lesser amounts of SO2 were desorbed in the temperature range of 200400 C than when adsorption took place at 320 C, but in the former case we noticed a continuous SO2 emission at higher temperatures. We attributed this behavior to the decomposition of aluminum sulfates at elevated temperatures [12]. Oxidation of SO2 over Ag and migration of SO3 on Al2 O3 was also reported by Abe et al. [13]. In the same work they postulate that Ag promoted the decomposition of the sulfates during the desorption stage of the TPD experiment. We studied the simultaneous adsorption/desorption of NO and SO2 in the presence of O2 (Figs. 4 and 5). The feed in these experiments was 2000 ppmv NO, 500 ppmv SO2 , 2% O2 in He and the adsorption temperature was 320 C. We followed the same experimental procedure for alumina and Rh/alumina. We compared the NOx desorption data for SO2 -free and SO2 -containing feeds over alumina and we noticed that the location of the rst NOx desorption peak (220 C) and the NO2 /NOx mole ratio (45% and 50% in the absence and presence of SO2 , respectively),

96

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

Fig. 3. Temperature programmed desorption experiment over the support (-alumina) and Rh/alumina. Feed: 1000 ppmv SO2 in He.

Fig. 4. NOx desorption curves from TPD experiments over the support (-alumina) and Rh/alumina for an NO/SO2 /O2 /He feed.

were similar (Fig. 4). The main difference was that the overall amount of the desorbed NOx was substantially higher in the absence of SO2 . The NOx desorption data from the Rh/alumina sample (Fig. 4) were, qualitatively, similar to those over alumina. In the presence of Rh the concentration of NO2 was slightly higher than that in the absence of Rh (55% and 50% of the emitted NOx was NO2 over Rh/alumina and alumina, respectively). It is important that, contrary to the experimental data of Fig. 2, no NOx desorption peak was noticed at 380 C in Fig. 4 over Rh/alumina.

Moreover, we did not detect the peak that was attributed to the dissociation of nitrates (480 C) when the feed contained SO2 . The evolution of SO2 from the alumina and Rh/alumina samples is shown in Fig. 5 for the NO/SO2 /O2 /He feed. Comparison of these data with those in Fig. 3 clearly shows that there was no weakly-sorbed SO2 when the sample was exposed to the NO/SO2 /O2 /He mixture. The SO2 emission at high temperatures was attributed to the aluminum sulfate decomposition. Xue et al. [14] measured a similar SO2 desorption curve over Pt-based samples using a

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

97

Fig. 5. SO2 desorption curves from TPD experiments over the support (-alumina) and Rh/alumina for an NO/SO2 /O2 /He feed.

feed of composition similar to this study. They noticed that the presence of NO in the feed enhanced the SO2 oxidation, while the NO oxidation was inhibited by SO2 . Given that in none of the two samples of Fig. 5 we noticed any SO2 released at temperatures lower than 450 C we postulate that SO2 is oxidized in the presence of O2 and NO even in the absence of Rh. Comparison between the NO desorption data over alumina and Rh/alumina showed that the RhNO bond is stronger than the NOalumina one. Moreover, higher amounts of NO2 were desorbed from Rh/alumina than from alumina, implying that Rh and/or the Rh/alumina interface act as oxidation sites. Over the same sites SO2 is oxidized to SO3 that either remains on the sample surface or is transferred to the alumina via a spill-over mechanism, apparently leading to the formation of Al2 (SO4 )3 . The NO adsorption was inhibited by the presence of SO2 in the feed, thus, we postulate that SO2 is preferentially adsorbed on the Rh sites as compared to NO.

3.2. FT-IR studies In this series of experiments we studied the adsorbed species that are formed on Rhalumina fresh or sulfated. The latter sample was exposed to an SO2 containing feed (1000 ppmv NO, 1000 ppmv C3 H6 ,

1000 ppmv SO2 , 5% O2 , He balance) at 330 C for 33 h. The feeds in the FT-IR experiments were: 1000 ppmv NO in He; 1000 ppmv NO, 5% O2 in He; 1000 ppmv C3 H6 in He; 1000 ppmv NO, 1000 ppmv C3 H6 , 5% O2 in He. The sorption temperature was 250, 300 and 350 C. A spectrum was recorded initially every minute and then every 5 minutes. We considered that steady state was reached when the spectrum remained unchanged for 10 minutes. The steady state spectrum of the fresh Rh/alumina sample exposed to 1000 ppmv NO in He at different temperatures is shown in Fig. 6. A band at 1906 cm1 was initially formed followed by a band at 1592 cm1 . The former band was attributed to the adsorption of NO on Rh as RhNO+ [1518], while the latter one to the adsorption of NO2 on the support [15,19,20]. The transient behavior of the sample (not shown in Fig. 6) implies that NO adsorbed on Rh spills over the support where it is oxidized. The same bands were also observed at higher temperatures (300 C and 350 C). At 300 C the NO2 band was signicantly enhanced with respect to the one at 250 C, in contrast to that of NO, while the height of both bands decreased at 350 C. As a result, the spill over of NO from the metal to the support is favored at a moderate temperature (300 C), while further increase in temperature (350 C) enhances the desorption of both NO and NO2 from the sample.

98

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

Fig. 6. FT-IR spectra of species formed during the NO adsorption on fresh Rh/alumina at 250 (curve a), 300 (curve b) and 350 C (curve c). Feed: 1000 ppmv NO in He.

The band at 1722 cm1 was attributed to the linearly-sorbed NO on Rh, as RhNO [21]. This adsorbed species was not noticed at higher temperatures, in agreement with the work of Hecker and Bell [22]. The bands in the 15471300 cm1 region correspond to nitrate species on the support [21,23]. In accordance to the NO2 band, the intensity of the nitrates band increased at 300 C and almost zeroed at 350 C, probably due to decomposition of these species. We studied the NO oxidation over the fresh Rh/alumina catalyst at 250 C, 300 C and 350 C (Fig. 7). The NO2 band (1592 cm1 ) was signicantly enhanced by the presence of O2 in the feed. The intensity of the RhNO+ (1906 cm1 ) and RhNO (1722 cm1 ) bands was substantially lower in Fig. 7 than those in Fig. 6. The shoulder at 1624 cm1 could be attributed to the following nitrate: Al-O-NOO-Al [21,23]. The bands at 1564, 1387, 1294 and 1265 cm1 also corresponded to nitrates bonded with alumina [23]. At higher temperatures the intensity of all bands decreased. The IR spectrum for C3 H6 (feed 1000 ppmv C3 H6 in He) adsorbed on fresh Rh/alumina at 250 C is shown in Fig. 8. At the initial stage of the experiment we observed a band at 2092 cm1 and a second less intense one at 2020 cm1 . At steady state the latter band was higher than the former one (Fig. 8). Worley et al. [24] reported that the adsorption of CO on preoxidized Rh/alumina at 298 K gave bonds of the form Rh + CO in the region 20302135 cm1 .

They postulated that gem-dicarbonyl species on Rh+1 correspond to the region 20302100 cm1 , while CO bonded to Rh+2 and Rh+3 to bands at 2120 and 2135 cm1 , respectively. They repeated the experiment on prereduced Rh/alumina and they measured three bands in the region 20312100 cm1 (2031 and 2101 cm1 : gem-carbonyl CO; 2065 cm1 : linear CO) and a band at 1850 cm1 that was assigned to the CO bridged with Rh0 . Solymosi and Sarkany [16] measured the symmetric and the asymmetric stretching vibrations of gem-dicarbonyl species at 2040 and 2108 cm1 , respectively. Zhang et al. [25] attributed bands shown at 2075 and 1992 cm1 to CO adsorbed linearly and bridge-bonded on oxidized Rh+ sites. The bands at 2092 and 2020 cm1 (Fig. 8) were, therefore, attributed to asymmetric and symmetric stretching vibrations of the gem-carbonyl CO on Rh, respectively. CO can be formed from the partial oxidation of C3 H6 with oxygen atoms that are weakly bonded with Rh. The reduction of Rh sites by C3 H6 promotes the adsorption of CO on the metal sites and especially those that are assigned to the band at 2020 cm1 . Other products of the partial oxidation of C3 H6 on the support can be allylic ions, acrylates, carbonates, ethanol and acetic acid. These species correspond to the following bands: 1691, 1591, 1581 and 1461 cm1 [23,26,27]. At 300 C the height of the RhCO bands increased with respect to that at 250 C, while those bands attributed to products of the partial oxidation of C3 H6 adsorbed on alumina decreased. A further tem-

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

99

Fig. 7. FT-IR spectra of species formed during the NO adsorption on fresh Rh/alumina at 250 (curve a), 300 (curve b) and 350 C (curve c). Feed: 1000 ppmv NO, 5% O2 in He.

Fig. 8. FT-IR spectra of species formed during the C3 H6 adsorption on fresh (curve a) and sulfated (curve b) Rh/alumina at 250 C. Feed: 1000 ppmv C3 H6 in He.

perature rise (350 C) lowered the intensity of all the bands. The time-on-stream variation of the spectrum obtained during the CSR of NO is shown in Fig. 9 for a feed of 1000 ppmv NO, 1000 ppmv C3 H6 , 5% O2 in He. Initially a band was noticed at 1892 cm1 that was assigned to RhNO+ and a smaller one at 2132 cm1 that was assigned to RhCN [18]. The height of these bands increased until steady state was reached. In the vicinity of the RhCN band a plateau was developed in the region 22502132 cm1 that was indicative of formation of NCO. Eventually two bands could be identied at 2179 and 2222 cm1 .

We attribute the former band to the RhNCO species and the latter one to the AlNCO. The spill-over of NCO from the noble metal to the support could lead to AlNCO [18], though Solymosi and Rasko [28] postulate that this mechanism is inhibited by the presence of O2 in the feed. Ukisu et al. [29] distinguish the NCO formed by a hydrocarbon/NO/O2 feed from that formed by a CO/NO feed. Moreover, they report that NCO on Rh is more active than that on Al2 O3 for the NO reduction to N2 . Sumiya et al. [9] noticed two intense bands at 2260 and 2230 cm1 that were attributed to AgNCO and AlNCO, respectively.

100

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

Fig. 9. Time evolution of FT-IR spectra of adsorbed species formed during the NO reduction with C3 H6 over fresh Rh/alumina at 250 C. Feed: 1000 ppmv NO, 1000 ppmv C3 H6 , 5% O2 in He.

The intense band at 1590 cm1 was attributed to NO2 or carbonates adsorbed on the support, in accordance to our previous remarks. The bands at 1465, 1394 and 1307 cm1 correspond to carbonylic species on the support as a result of the partial oxidation of C3 H6 [30]. The asymmetric vibrations of the CH bond (CH2 and CH3 species) on the support cause the bands at 3000 and 2909 cm1 . At higher temperatures a shoulder was noticed at 1724 cm1 and a band at 1643 cm1 . The former can be assigned to the RhNO species and the latter one to RhONO [31]. The height of the NCO and CHx bands decreased at higher temperatures. We examined the effect of SO2 on the SCR of NO with C3 H6 in excess O2 as follows: A Rh/alumina sample was sulfated in a xed-bed reactor arrangement for 33 h using a feed of 1000 ppmv NO, 1000 ppmv C3 H6 , 1000 ppmv SO2 , 5% O2 in He. The sulfated sample was loaded in the FT-IR spectrometer, it was calcined at 420 C for 30 min under air ow and it was then exposed to different feeds. In curve (a) of Fig. 10 we present the spectra of the sulfated catalyst exposed to an NO/He feed. The sample exhibited the RhNO+ band at 1931 cm1 . It is important that neither NO2 nor nitrates exist on the sulfated sample. Chang [32] studied the adsorption of SO2 on -alumina in the presence of O2 and noticed a plateau in the region of 1370937 cm1 that was attributed to the presence of sulfates on alumina. A similar plateau was noticed in the background of the sulfated sample (not shown

in Fig. 10). A temperature rise from 250 C to 350 C caused a decrease in the height of the RhNO+ band of the sulfated Rh/alumina. The same band was also noticed when the feed was 1000 ppmv NO, 5% O2 in He (curve (b) of Fig. 10). In general, the sulfated samples did not exhibit any products of the NO oxidation even in the presence of excess O2 . We attribute this behavior to the coverage of the catalytic oxidation sites by sulfates that decompose only at high temperatures. On the other hand, NO can be adsorbed on Rh as RhNO+ . In curve (b) of Fig. 8 the feed was 1000 ppmv C3 H6 in He and the reaction temperature 250 C. The bands (2114 and 2047 cm1 ) were shifted 20 1 cm with respect to those over the fresh sample, in accordance to the remarks of Apesteguia et al. [33]. This shift was attributed by the above authors to a decrease in the electronic density of the metal in the presence of sulfur, an electron acceptor. The shift toward higher wavelengths is linked, according to Worley et al. [24], with the adsorption of gas species on a metal of higher oxidation state. Yao et al. [34] postulate that Pt catalyzes the oxidation of sultes to sulfates and the formation of aluminum sulfates limits the associative chemisorption of C3 H6 . This is in accordance to the lower bands of the sulfated samples with respect to those of the fresh samples in Fig. 8. A negative band was noticed in those samples that were exposed to a propene-containing feed at 1394 cm1 . We attribute this band to the reduction

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

101

Fig. 10. FT-IR spectra of species formed during the NO adsorption on sulfated Rh/alumina at 350 C for different feeds. Feeds: 1000 ppmv NO in He (curve a); 1000 ppmv NO, 5% O2 in He (curve b); 1000 ppmv NO, 1000 ppmv C3 H6 , 5% O2 in He (curve c).

of the sulfates under the reducing conditions of the experiments. Our explanation agrees with the band at 1390 cm1 measured by Chang [32] that was attributed to aluminum sulfates. It is possible that the reduction of the sulfates by the reductant produces those sites on the support where propene is adsorbed and then partially oxidized to Cx Hy Oz (bands <1700 cm1 ). Higher reaction temperatures decrease the height of the negative peak at 1394 cm1 and those of Cx Hy Oz . In curve (c) of Fig. 10 we present the spectrum of the sulfated Rh/alumina (at 350 C) exposed to a feed of 1000 ppmv NO, 1000 ppmv C3 H6 , 5% O2 in He. The biggest difference between this spectra and the corresponding in Fig. 6 is the disappearance of the bands: region 22502132 cm1 (NCO), region 30002900 cm1 (stretching vibrations of the CH bond), and band at 1590 cm1 (NO2 or carbonates). The formation of NCO species over Ag/alumina was strongly inhibited by the presence of SO2 in the feed according to Sumiya et al. [9], as well. The negative band at 1396 cm1 corresponds to sulfated species reduced by C3 H6 , in accordance to the results obtained during C3 H6 adsorption over the same sample (curve (b) of Fig. 8). The bands in the 17501450 cm1 region were attributed to carbonyl species on the support. It is interesting that only RhNO+ species were observed over the sulfated Rh/alumina (band at 1931 cm1 for NO/He and NO/O2 /He feeds and at 1919 cm1 for NO/C3 H6 /O2 /He feed).

3.3. NO oxidation We studied the NO oxidation over the Rh/alumina catalyst (Fig. 11) using a feed of 1000 ppmv NO, 5% O2 in He ( markers). The dashed line in this gure corresponds to the equilibrium NO conversion to NO2 at various temperatures for the above feed. The maximum conversion of NO to NO2 (60%) was measured at 337 C. At higher temperatures the experimental data coincided with the equilibrium values, while at lower temperatures we measured lower NO2 conversions. The deviation between the measured and the equilibrium conversions at low temperatures was attributed to kinetic limitations. In Fig. 11 we also present the NO conversion to N2 ( markers) and the NO oxidation to NO2 ( markers) for a feed of 1000 ppmv NO, 1000 ppmv C3 H6 , 5% O2 in He over the Rh/alumina catalyst. Addition of the above two curves results to a curve close to that of the NO oxidation ( markers). This remark may imply that under oxidation conditions all the Rh sites are used for the NO2 formation, while under reduction conditions (presence of 1000 ppmv C3 H6 ) part of the Rh sites change from oxidation to reduction sites. We examined the effect of the addition of 200 ppmv SO2 in the feed on the NO oxidation. The experimental data in Fig. 11 (+ markers) clearly show that the presence of SO2 in the feed severely inhibited the NO oxidation over Rh/alumina. About 8% of the inlet NO was oxidized to NO2 in the temperature range

102

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

Fig. 11. NO oxidation over Rh/alumina. () Feed: 1000 ppmv NO, 5% O2 in He; (+) Feed: 1000 ppmv NO, 200 ppmv SO2 , 5% O2 in He; (, ) Feed: 1000 ppmv NO, 1000 ppmv C3 H6 , 5% O2 in He.

of 200400 C, contrary to the experimental results in the absence of SO2 . When we removed the SO2 from the feed the initial activity of the catalyst was not recovered, implying that the inhibition effect was irreversible.

3.4. NO reduction In our previous work [4] we noticed that the addition of SO2 to the feed caused a reversible enhancement on the NO reduction over Rh/alumina at 320 C. Another interesting behavior associated with the presence of SO2 in the feed was the inhibition of NO oxidation. Typical NOx and C3 H6 conversion versus temperature curves are presented in Fig. 12(a) for a feed of 1000 ppmv NO, 1000 ppmv C3 H6 , 0 or 250 ppmv SO2 , 5% O2 in He. The reactor loading was 0.2 g and the reaction temperature changed from high to low values. We estimated the apparent activation energy for the NO reduction and the C3 H6 consumption from the experimental data of Fig. 12 for an SO2 -free feed: for NO reduction 31 3 kcal/mol; for C3 H6 consumption 34 8 kcal/mol. The values of these activation energies are similar, implying the possible existence of a common reaction intermediate for NOx reduction and C3 H6 oxidation. When the feed included 250 ppmv SO2 we noticed a sharp decrease of both the NOx and

the C3 H6 conversion with decreasing temperature at ca. 300 C. The apparent activation energy for NO reduction was almost double than the one of the SO2 free experiment. No precise value of the apparent activation energy is given because only a small number of experimental points met the differential conditions. The difference in the apparent activation energies in the presence and the absence of SO2 in the feed is evidence of a different reaction mechanism between the two reaction systems. In Fig. 12(b) we present the COx formation as a result of the C3 H6 oxidation. The presence of SO2 in the feed enhanced the formation of CO, though most of the consumed C3 H6 was converted to CO2 . We measured the activity of the sulfated Rh/alumina for the oxidation of NO and C3 H6 as follows: 0.5 g of Rh/alumina was initially exposed to a stream of 2000 ppmv NO, 500 ppmv SO2 , 2% O2 in He at 320 C for 1 h and it was then purged with He. This sample was then exposed to a 2000 ppm NO, 2% O2 in He mixture at 320 C and we measured no appreciable NO2 formation, implying the poisoning of the NO oxidation sites by preadsorbed SO2 . On the contrary, when the same sample was exposed to a 1000 ppmv C3 H6 , 2% O2 in He mixture, we measured almost complete oxidation of the hydrocarbon. Therefore, NO and C3 H6 are oxidized over different sites. The former sites are irreversibly deactivated by SO2 , while

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

103

Fig. 12. (a) Effect of SO2 on the NOx and C3 H6 conversion vs. temperature curves. (b) Effect of SO2 on the COx formation. () Feed: 1000 ppmv NO, 1000 ppmv C3 H6 , 5% O2 in He; (- - -) 1000 ppmv NO, 1000 ppmv C3 H6 , 250 ppmv SO2 , 5% O2 in He.

the latter ones are either not affected by SO2 or the interaction of the sample with the SO2 (e.g. aluminum sulfate) forms new sites for C3 H6 oxidation. Kinetic experiments were carried out at 306 C varying the NO concentration in the range of 2502000 ppmv using a reactor loading of 0.2 g. In Fig. 13 we present the variation of the NO reduction and C3 H6 oxidation rates with the mean NO concentration, dened as the average between the inlet and the exit concentration. Conversions lower than 10% were measured during the kinetic experiments for the NO and C3 H6 , assuring the differential operation of the reactor. The apparent order of reaction with respect to NO was 0.5, while the rate of the overall C3 H6 consumption was independent of the inlet NO concentration.

4. Discussion Our experimental results showed that the SCR of NO with C3 H6 over Rh/alumina is signicantly affected by the presence of SO2 in the feed, though this gas species is not directly involved in the above reaction. In the previous section we noticed that aluminum sulfate is formed on the catalyst when SO2 and O2 coexist in the feed. The ratio of the Al2 (SO4 )3 and Al2 O3 molar volumes is about 5, that is, pore plugging could take place. However, if diffusional limitations through the product layer are signicant, the extent

of this reaction would be limited to the outer surface of the catalyst only. We compared the pore structure (surface area, pore volume and pore size distribution) of Rh/alumina samples reacted in an SO2 -free and an SO2 -containing feed under identical reaction conditions and we did not observe any signicant difference between the pore structure of the two samples. Therefore, the formation of aluminum sulfate has not changed the internal pore structure of Rh/alumina. We combined the results of the TPD, FT-IR and activity studies in order to predict the reaction mechanism for the NO reduction over Rh/alumina using an SO2 -free feed. In the temperature range of 250350 C we detected NO+ , NO , NCO and CN species on the Rh sites. Cyanides are strongly-bound to Rh and most probably do not participate in the SCR of NO [18]. We also used the X-ray Photoelectron Spectroscopy (XPS) to determine the oxidation states of Rh on a catalyst reacted under the standard feed (1000 ppmv NO, 1000 ppmv C3 H6 , 5% O2 , in He) at different temperatures. On the fresh sample Rh was in the Rh2 O3 form, while about 10% of the initial Rh was reduced as a result of the SCR of NO with C3 H6 . Therefore, we propose the following reaction mechanism for the NO reduction (absence of SO2 in the feed). NO is adsorbed either associatively on fullyoxidized Rh sites (RhNO+ ) or dissociatively on reduced Rh sites (RhN) [16]. Products of the partial oxidation of propene are adsorbed on Rh/alumina

104

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

Fig. 13. Dependence of the rate of NO reduction and C3 H6 oxidation on the NO concentration. Feed: 2502000 ppmv NO, 1000 ppmv C3 H6 , 5% O2 in He.

(FT-IR studies) and interact with RhN: RhN + RhCO RhNCO + Rh Formation of isocyanate RhNCO + RhNO+ N2 + CO2 + 2Rh (1) (2)

A similar reaction scheme was also proposed by Bamwenda et al. [18], Ukisu et al. [29], Haneda et al. [31] and Sumiya et al. [9]. Given that the order of reaction with respect to NO is 0.5 we postulate that the formation of the isocyanate (Eq. (1)) is the controlling step of the overall reaction under the reaction conditions of this study. This is in accordance with our previous remark that similar apparent activation energies with respect to NO and C3 H6 imply the existence of a common reaction intermediate for NO reduction and C3 H6 oxidation. If an apparent order of reaction higher than one were estimated, then Eq. (2) would be the rate determining step [35]. The oxidation of NO to NO2 or NO3 over a precious metal (e.g. Pt or Rh) is the initial step of proposed oxidation schemes for the NO reduction [17,36]. Takahashi et al. [37] introduced the concept of a compound where NO3 is stored before its reduction to N2 . In our TPD, FT-IR and activity studies we concluded that the presence of SO2 in the feed inhibited the oxidation of NO. Obviously, this implies that the reduction of NO to N2 does not proceed via its oxidation to NO2 . Burch

et al. [10] proposed an NO decomposition mechanism over Pt/alumina. The key step in this mechanism is the reduction of the oxidized Pt sites by C3 H6 and thus, the generation of reduced Pt sites where NO adsorbs dissociatively and recombines to release N2 or N2 O. It was noticed in previous studies [16,19] that the dissociative adsorption of NO takes place on partiallyreduced Rh sites. However, our XPS measurements showed that Rh in the sulfated Rh/alumina is fully oxidized, though C3 H6 exists in the feed. Therefore, we postulate that a different NO reduction mechanism applies over Rh/alumina when there is SO2 in the feed. This does not necessarily apply for an SO2 -free feed, where the proposed reaction mechanism involves the dissociative adsorption of NO and the recombination of two N atoms may lead to dinitrogen in addition to that formed according to mechanism proposed in this study. Finally, no RhNCO band was observed in the FT-IR experiments over the sulfated sample, implying that a reaction scheme different from that described by Eqs. (1) and (2) should be considered for the SO2 containing experiments. The reaction of partially oxidized hydrocarbons with NO has been proposed [3840] as a possible NO reduction mechanism over Cu-ZSM5 and Ptalumina. When the sulfated sample was used in our IR experiment we noticed the RhNO+ band along with those of carbonyl species formed on the support. A higher apparent activation energy was estimated in the SO2 -

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106

105

containing experiments with respect to the SO2 -free experiments. This implies that a less complicated reaction scheme than the one expressed by Eqs. (1) and (2) applies in the former reaction conditions. Based on the above remarks we propose the following reaction mechanism for the NO reduction with C3 H6 in the presence of SO2 and excess O2 : RhNO+ +Ca Hb Oc [N]+COx +H2 O +Rh (3)

In the above discussions we have not referred to the contribution of the support on the SCR of NO. Xue et al. [11] emphasized the role of the support on the oxidation of NO and SO2 and they reported that alumina, an amphoteric support, tends to react with acidic molecules like NOx and SO2 . On the other hand, Amiridis et al. [41] attributed the deNOx activity of Rh/alumina to the ability of alumina to convert C3 H6 into partially-oxidized species and subsequently to stabilize them on the surface. Hamada [42], as well, postulated that there is some cooperation between noble metals and alumina. In the same study an enhancement in NO reduction was noticed over some sulfatepromoted metal oxide catalysts, while inhibition was noticed over sulfate-promoted alumina. The high activity of the above catalyst was linked with a change in the acid strength and amount of the active sites. Our FT-IR results indicated an inhibition of the NO and C3 H6 oxidation on alumina over the sulfated samples (Figs. 8 and 10). Our previous activity studies showed that SO2 affects the NO reduction reversibly and the NO oxidation irreversibly [4]. Therefore, we postulate that the Rh atoms act as the reduction sites of the Rh/alumina catalyst, while the interface between Rh and alumina catalyzes the oxidation of NO to NO2 . Exposure of the catalyst to an SO2 environment could apparently modify this interface and thus deactivate the above oxidation sites. This can lead to higher reduction rates over the sulfated catalyst than those over the fresh catalyst.

SO2 to SO3 and, probably, the formation of aluminum sulfate. When an NO/SO2 /O2 feed was used in the TPD experiment we did not observe the NOx peak that was linked with the Rh presence in the catalyst nor that of the SO2 adsorbed on the support. We attributed this behavior to the preferential adsorption of SO2 on Rh and the subsequent oxidation of SO2 to SO3 that spills over the support. We used our FT-IR experiments to identify surface species on Rh/alumina that participate in the NO reduction. The proposed reaction mechanism for an SO2 -free feed is based on the reaction of isocyanates with nitrosonium species. Both species are formed on the Rh sites. The role of propene in the proposed reaction scheme is double, i.e., to form CO on Rh and to reduce the oxidation state of Rh. On the other hand, in the presence of SO2 in the feed Rh maintained its initial oxidation state and our FT-IR showed the formation of RhNO+ and that of products of the partial oxidation of propene on the support. The proposed reaction mechanism for an SO2 -containing feed is the reaction of RhNO+ and Ca Hb Oc derived from the partial oxidation of propene. Our activity studies showed that the catalytic reduction of nitric oxide with propene in excess oxygen was enhanced when we added SO2 in the feed, while the NO oxidation was inhibited. We believe that these two effects are interconnected since the inhibition of the oxidation reaction leads to higher concentration of the compounds involved in the NO reduction and to a less complicated reaction mechanism for NO reduction. We postulate that the oxidation reactions take place on the Rhalumina interface, while the reduction of NO on the noble metal. The addition of the SO2 in the feed causes the sulfation of the support, and thus deactivates these oxidation sites.

Acknowledgements The authors wish to thank prof. X.E. Verykios for the opportunity to carry out the FT-IR experiments in the laboratory of Heterogeneous Catalysis of the University of Patras and Dr. D. Kontaridis and Ph.D. student K. Elmasidis for the helpful discussions. This work was funded by the Commission of the European Community, under Contracts EV5V-CT094-0535 and ENV4-CT97-0658.

5. Conclusions TPD studies showed that in the presence of O2 in the feed the formation of NO2 and nitrates from adsorbed NO was enhanced. The overall amount of SO2 adsorbed on the alumina sample is smaller than that adsorbed on Rh/alumina, implying the oxidation of

106

E.A. Efthimiadis et al. / Applied Catalysis B: Environmental 22 (1999) 91106 [24] S.D. Worley, J.P. Wey, W.C. Neely, in: D.J. Dwyer, F.M. Hoffmann (Eds.), Surface Science of Catalysis, ACS Symp. Ser. 482, 1416 April 1991, ACS, Washington, 1991, p. 251. [25] Z. Zhang, A. Kladi, X.E. Verykios, J. Phys. Chem. 98 (1994) 6804. [26] J.A. Anderson, C.H. Rochester, J. Chem. Soc., Faraday Trans. 1 85 (1989) 1117. [27] V.A. Matyshak, O.V. Krylov, Catal. Today 25 (1995) 1. [28] F. Solymosi, J. Rasko, Appl. Catal. 10 (1984) 19. [29] Y. Ukisu, S. Sato, A. Abe, K. Yoshida, Appl. Catal. B 2 (1993) 147. [30] T. Chak, A.M. Efstathiou, X.E. Verykios, J. Phys. Chem. B 101 (1997) 7968. [31] M. Haneda, Y. Kintaichi, M. Inada, H. Hamada, Proc. JECAT, Tsukuba, Paper No. PII-7, 1997, p. 407. [32] C.C. Chang, J. Catal. 53 (1978) 374. [33] C.R. Apesteguia, C.R. Brema, T.F. Garetto, A. Borgna, J.M. Parera, J. Catal. 89 (1984) 52. [34] H.C. Yao, H.K. Stepien, H.S. Gandhi, J. Catal. 67 (1981) 231. [35] G.F. Froment, K.B. Bischoff, Chemical Reactor Analysis and Design, Wiley, New York, 1979. [36] T. Tanaka, T. Okuhara, M. Misono, Appl. Catal. B 4 (1994) L1. [37] N. Takahashi, H. Shinjoh, T. Iijima, T. Suzuki, K. Yamazaki, K. Yokota, H. Suzuki, N. Miyoshi, S. Matsumoto, T. Tanizawa, T. Tanaka, S. Tateichi, K. Kasahara, Catal. Today 27 (1996) 63. [38] C.J. Bennett, P.S. Bennett, S.E. Golunski, J.W. Hayes, A.P. Walker, Appl. Catal. 86 (1992) L1. [39] J.L. dItri, W.M.H. Sachtler, Catal. Lett. 15 (1992) 289. [40] M. Sasaki, H. Hamada, Y. Kintaichi, T. Ito, Catal. Lett. 15 (1992) 297. [41] M.D. Amiridis, T. Zhang, R.J. Farrauto, Appl. Catal. B 10 (1996) 203. [42] H. Hamada, Catal. Today 22 (1994) 21.

References
[1] Y. Li, J.N. Armor, Appl. Catal. B 5 (1995) L257. [2] X. Feng, W.K. Hall, J. Catal. 166 (1997) 368. [3] T. Tabata, M. Kokitsu, H. Ohtsuka, O. Okada, L.M.F. Sabatino, G. Bellussi, Catal. Today 27 (1996) 91. [4] E.A. Efthimiadis, G.D. Lionta, S.C. Christoforou, I.A. Vasalos, Catal. Today 40 (1998) 15. [5] N. Okazaki, S. Tsuda, Y. Igarashi, A. Tada, Chem. Lett. (1997) 635. [6] R. Burch, T.C. Watling, Appl. Catal. B 17 (1998) 131. [7] G. Zhang, T. Yamaguchi, H. Kawakami, T. Suzuki, Appl. Catal. B 1 (1992) L15. [8] R. Burch, E. Halpin, J.A. Sullivan, Appl. Catal B 17 (1998) 115. [9] S. Sumiya, M. Saito, H. He, Q.-.C. Feng, N. Takezawa, K. Yoshida, Catal. Lett. 50 (1998) 87. [10] R. Burch, P.J. Millington, A.P. Walker, Appl. Catal. B 4 (1994) 65. [11] E. Xue, K. Seshan, J.R.H. Ross, Appl. Catal. B 11 (1996) 65. [12] J.C. Summers, Environ. Sci. Technol. 13 (1979) 321. [13] A. Abe, N. Aoyama, S. Sumiya, N. Kakuta, K. Yoshida, Catal. Lett. 51 (1998) 5. [14] E. Xue, K. Seshan, J.G. van Ommen, Appl. Catal. B 2 (1993) 183. [15] E.A. Hyde, R. Rudham, C.H. Rochester, J. Chem. Soc., Faraday Trans. 1 80 (1984) 531. [16] F. Solymosi, J. Sarkany, Appl. Surf. Sci. 3 (1979) 68. [17] S. Naito, M. Tanimoto, Chem. Lett. (1983) 1935. [18] G.R. Bamwenda, A. Ogata, A. Obuchi, J. Oi, K. Mizuno, J. Skrzypek, Appl. Catal. B 6 (1995) 311. [19] H. Arai, H. Tominaga, J. Catal. 43 (1976) 131. [20] G. R. Banwenda, A. Obuchi, A. Ogata, K. Mizuno, Chem. Lett. (1994) 2109. [21] S.C. Chuang, C.-.D. Tan, J. Catal. 173 (1988) 95. [22] W.C. Hecker, A.T. Bell, J. Catal. 92 (1985) 247. [23] A.A. Davydov, in: C.H. Rochester (Ed.), Infrared Spectroscopy of Adsorbed Species on the Surface of Transition Metal Oxides, Wiley, Chichester, 1990.

Vous aimerez peut-être aussi