You are on page 1of 13

Applied Catalysis A: General 234 (2002) 221233

Catalytic oxidation of methane over CeO2 -ZrO2 mixed oxide solid solution catalysts prepared via urea hydrolysis
Sitthiphong Pengpanich a,c , Vissanu Meeyoo a, , Thirasak Rirksomboon b , Kunchana Bunyakiat b,c
a

Centre for Advanced Materials and Environmental Research, Mahanakorn University of Technology, Bangkok 10530, Thailand b Petroleum and Petrochemical College, Chulalongkorn University, Bangkok 10330, Thailand c Department of Chemical Technology, Chulalongkorn University, Bangkok 10330, Thailand Received 22 October 2001; received in revised form 28 December 2001; accepted 28 March 2002

Abstract In this study, CeO2 -ZrO2 mixed oxide catalysts were prepared via urea hydrolysis and tested for methane oxidation. Highly uniform solid solution particles of ceria-zirconia were obtained under the conditions of this study. The incorporation of Zr into the CeO2 lattice was found to promote the redox properties. The methane oxidation activity of the mixed oxides was found to be dependent on the Ce:Zr ratio, which relates to the degree of reducibility. It was postulated that the cubic phase, uorite structure, which is mainly found in Ce1 x Zrx O2 (where x < 0.5) can be reduced more easily than the tetragonal phase found in Ce1 x Zrx O2 (where x > 0.5). The catalytic activity decreased with an increasing Zr content. The mixed oxide catalyst, Ce0.75 Zr0.25 O2 solid solution, was reported to exhibit the highest activity for methane oxidation. Kinetic studies of methane oxidation over such a mixed oxide catalyst (Ce0.75 Zr0.25 O2 ) showed that the methane oxidation rates strongly depend on methane concentration, but only slightly on the oxygen concentrations. The LangmuirHinshelwood mechanism (oxygen dissociative chemisorption on the active sites and non-dissociative chemisorption of methane) can satisfactorily t the experimental results obtained from the kinetic studies for this catalyst. The activation energy of methane oxidation is calculated based on this surface reaction mechanism as being 100.8 kJ/mol. 2002 Elsevier Science B.V. All rights reserved.
Keywords: CeO2 -ZrO2 ; Solid solution; Methane oxidation; Urea hydrolysis

1. Introduction The production of energy by the combustion of methane or natural gas is well established. However, gas-phase combustion can only occur within the ammability limits and the temperatures produced during the combustion process can rise to above

Corresponding author. Tel.: +66-2-988-4039; fax: +66-2-988-4039. E-mail address: vissanu@mut.ac.th (V. Meeyoo).

ca. 1600 C, whereby unwanted nitrogen oxides are formed [1]. Catalytic combustion offers the possibility of being an alternative means of energy production. A wide range of concentrations of hydrocarbons can be oxidized over a suitable catalyst, and it is possible to operate outside the ammability limits of fuels. In addition, the reaction conditions can usually be controlled more precisely, with reaction temperatures being maintained well below 1600 C. This may be a promising way both to decrease NOx emission and to avoid thermal sintering of the catalyst [1,2].

0926-860X/02/$ see front matter 2002 Elsevier Science B.V. All rights reserved. PII: S 0 9 2 6 - 8 6 0 X ( 0 2 ) 0 0 2 3 0 - 2

222

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

Catalysts used for catalytic combustion are classied in two groups: noble metals and transition metal oxides. The use of noble metals such as Pt, Pd or Rh on a supporting substrate like alumina in combustion catalysis is well known. However, due to the severe operating conditions, many catalysts and supports are not highly thermally resistant [3]. Transition metal oxides are the other class of catalysts used for the catalytic oxidation of hydrocarbons. Although they are less active than noble metals, they have the advantage that numerous active solid state phases can be obtained [4]. Various transition metal oxides have been investigated as catalysts and supports for catalytic combustion of methane [2,4,5]. Ceria (CeO2 ) has been widely used as a promoter and an oxidation catalyst because of its unique redox properties and high oxygen storage capacity [6,7]. It has been reported that CeO2 has potential uses for the removal of post-combustion pollutants, for the removal of organics from wastewater (catalytic wet oxidation) and in fuel cell technology [7]. In addition, CeO2 is a key component in the three-way catalysts (TWCs) which are extensively used for automotive emission control [7,8]. There are multiple reasons for depositing CeO2 onto TWC surface. It is known to stabilize the well-dispersed noble metals and to promote the water-gas shift and hydrocarbon-reforming reactions. Moreover, it has been suggested for uses to store and release oxygen under lean and rich conditions, respectively [8]. However, pure CeO2 has poor thermal stability properties [9]. Recently, it has been reported that the addition of zirconia (ZrO2 ) to CeO2 leads to improvements in its oxygen storage capacity, redox properties and thermal resistance, and to better catalytic activity at lower temperatures [1012]. This was found to be due to the partial substitution of Ce4+ with Zr4+ in the lattice of CeO2 , which consequently resulted in a solid solution formation [6,10,11]. Hori et al. [10] reported that the benecial effects of ZrO2 were pronounced in solid solutions which had the oxygen storage capacity values three to ve times higher than that of pure CeO2 . When the optimum Zr dopant concentration was 25 mol%, solid solution materials aged at 1000 C showed a higher oxygen storage capacity than that of pure CeO2 . This result could be ascribed to an occurrence of Zr promotion via an alteration of bulk properties of the CeO2 . Furthermore,

the redox properties of CeO2 have been enhanced by the addition of Zr4+ into the lattice of CeO2 by the formation of solid solutions [9]. Nevertheless, it was also reported that CeO2 -ZrO2 solid solutions showed a high catalytic activity, particularly for oxidation of CO, CH4 and C3 H6 [6,12,13]. The aforementioned reasons led us to begin a study for developing an appropriate solid solution preparation method of such mixed oxides and a better understanding of the role of its activity. Many preparation methods have been applied for the preparation of CeO2 -ZrO2 solid solution for catalytic applications. These included the high-temperature ring or high-energy milling of a mixture of the oxides [14], co-precipitation [6,10,15] and solgel techniques [16]. Among these methods, the solgel technique is found to be very benecial since it yields products with high purity, homogeneity and well-controlled properties; it is a low temperature process as well. The properties of the nal products were dependent on the temperature and on the hydrolysis catalysts [17,18]. In this study, we attempted to prepare the solid solutions of CeO2 -ZrO2 via urea hydrolysis. The catalytic activities of these materials were tested for the oxidation of methane and the kinetics of methane oxidation over the selected catalysts was also studied.

2. Experimental 2.1. Catalyst preparation In this study, mixed oxide solid solutions of Ce-Zr were prepared via urea hydrolysis. The Ce-Zr mixed oxide samples were prepared from Ce(NO3 )3 6H2 O (99.0%, Fluka) and ZrOCl2 8H2 O (99.0%, Fluka). The starting metal salts were dissolved in distilled water to the desired concentration (0.1 M). The ratio between the metal salts was altered depending on the desired solid solution concentration: Ce1 x Zrx O2 in which x = 0, 0.25, 0.50, 0.75 and 1.0. Then, the mixed metal salt solution was added with a 0.4 M of urea (99.0%, Fluka) solution with the salt to urea solution ratio of 2:1 (v/v), and the mixture was kept at 100 C for 50 h. The sample was then allowed to cool to room temperature prior to being centrifuged to separate a gel product from the solution. The gel product was

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

223

washed with ethanol and dried overnight in an oven at 110 C. The product was then calcined at either 500 or 900 C for 4 h. 2.2. Characterizations BrunauerEmmettTeller (BET) surface area was determined by N2 adsorption at 77 K (ve point BET method using a Quantachrome Corporation Autosorb). Prior to the analysis, the samples were outgassed to eliminate volatile adsorbents on the surface at 250 C for 4 h. A Rigagu X-ray diffractometer system equipped with a RINT 2000 wide-angle goniometer using Cu K radiation and a power of 40 kV 30 mA was used for examination of the crystalline structure. The intensity data were collected at 25 C over a 2 range of 2080 with a scan speed of 5 (2 )/min and a scan step of 0.02 (2 ). FT-Raman spectra were obtained using a PerkinElmer 2000 FT-Raman spectrometer with a diode pumped YAG laser and a room temperature super InGaAs detector. The laser power was ca. 460 mW. A frequency range of 1004000 cm1 was observed. Temperature programmed reduction (TPR) measurements were carried out to investigate the redox properties over the resultant materials. About 50 mg of catalyst was placed in a quartz tube and pretreated in a 20 ml/min He atmosphere at 400 C for 1 h prior to running the TPR experiment, and then cooled down to room temperature in He. The feed of 1% CO in He at a ow rate of 50 ml/min was used as a reducing gas. The temperature of the sample was raised at a constant rate of 10 C/min. The amount of CO consumption during the increasing temperature period was measured using a mass spectrometer (Balzer Instruments modeled Thermostar GSD 300T).
Table 1 BET surface areas (m2 /g) of catalysts with the aging time = 50 h Calcination temperature ( C) Ce:Zr ratio 100:0 500 900 101.6 4.6 75:25 108.4 9.2

2.3. Activity tests and kinetic studies Catalytic activity tests for methane oxidation were carried out in a differential quartz tube microreactor (i.d. 10 mm). Typically, 0.1 g of catalyst was packed between the layers of quartz wool. The reactor was placed in an electric furnace equipped with K-type thermocouples. The temperature of catalyst bed was monitored and controlled by Shinko temperature controllers. Gas mixtures containing 2% methane, 21% oxygen and balance of helium were used. The total ow rate of feed gas into the reactor was kept at 100 ml/min by the use of Brooks mass ow controllers. Exit gases were chromatographically analyzed (Shimadzu GC 17A tted with a thermal conductivity detector) after separation on a 30 m GS-CarbonPLOT column (30 C). The kinetic studies were carried out in the same system as mentioned earlier at temperatures of 400, 450, 500, 550 and 600 C. The concentrations of the reactant gases were varied between 0.5 and 2.0% for methane, 5.0 and 21.0% for oxygen, and balance of helium for a total ow rate of 200 ml/min. The kinetic data were obtained using the initial rate method.

3. Results and discussion 3.1. BET surface area As shown in Table 1, the BET surface areas of the mixed oxides prepared via urea hydrolysis are reasonably high. It was found that the BET surface areas of CeO2 -ZrO2 mixed oxide catalysts are slightly higher than that of pure CeO2 . It was noticed that the BET surface areas of CeO2 -ZrO2 mixed oxide catalysts increase with increasing Zr content. This might be due to the substitution of a Zr4+ ion which has a smaller cationic radius, in the Ce4+ lattice location.

50:50 116.0 12.5

25:75 120.1 21.3

0:100 79.0 12.2

224

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

Fig. 1. XRD patterns for Ce1 x Zrx O2 mixed oxide catalysts with the aging time = 50 h and calcined at 500 C: ( ) cubic phase; ( tetragonal phase; ( ) monoclinic phase.

Interestingly, at high temperatures (ca. 900 C), the surface areas of all the oxide samples were found to decrease drastically, but the loss of surface area of the mixed oxide samples was lower than that of pure CeO2 . It is promising that the stabilization of the surface area of CeO2 is probably due to inhibition of the sintering process by ZrO2 doping. 3.2. XRD and FT-Raman Fig. 1 shows the XRD patterns for Ce1 x Zrx O2 mixed oxides. It was found that, after calcination at 500 C, the XRD patterns of CeO2 and Ce0.75 Zr0.25 O2 show visible tailing at about 29, 33, 48, and 60 , (2 ) which represent the indices of (1 1 1), (2 0 0), (2 2 0) and (2 2 2) planes, respectively. This indicates a cubic uorite structure. No evidence for extra peaks due to non-incorporated ZrO2 was observed in any XRD patterns of Ce0.75 Zr0.25 O2 . This suggests that ZrO2 will
Table 2 Values of lattice parameter (a, ) of catalysts Calcination temperature ( C) Ce:Zr ratio 100:0 500 900 5.4201 6.0452 75:25 5.4126 5.9118

be incorporated into the CeO2 lattice to form a solid solution while maintaining the uorite structure. The lack of free ZrO2 was also conrmed by FT-Raman spectroscopy. Conversely, Ce1 x Zrx O2 samples with x 0.5 indicate the presence of a mixture of two phases: the cubic and the tetragonal. Some tetragonality of the attained phase was suggested by the splitting of the (3 1 1) and (4 0 0) reection at about 79 and 70 (2 ), respectively. The presence of a monoclinic phase is only observed in the pure ZrO2 sample. It should be noticed that the diffraction peaks were shifted to higher degrees with the increasing amounts of ZrO2 . This observation was attributed to shrinkage of lattice due to the replacement of Ce4+ (1.09 ) with a smaller cation radius Zr4+ (0.86 ) [6]. Calculation of the cell parameters (a) was carried out using the (1 1 1) plane. The values of a are given in Table 2. It can be seen that the values of a decrease slightly with the increasing

50:50 5.4088 5.8718

25:75 5.4052 5.8339

0:100 5.4804 5.8727

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

225

Fig. 2. Raman spectra of Ce1 x Zrx O2 mixed oxide catalysts with the aging time = 50 h and calcined at 500 C.

Fig. 3. TPR prole of Ce1 x Zrx O2 mixed oxide catalysts calcined at 500 C with aging time = 50 h. The reducing gas contains 1% carbon monoxide in helium and a heating rate of 10 C/min was used. The gas ow rate was 50 ml/min.

226

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

Zr content. The lattice parameters are in the range of 56 . These results are in good agreement with those obtained from other conventional methods [15,19]. The FT-Raman spectra of CeO2 and Ce0.75 Zr0.25 O2 , as shown in Fig. 2, present only the adsorption peak ca. 456 cm1 , which is a typical of the F2g Raman active mode of the uorite-structured materials [9,20,21]. No characteristic band of pure ZrO2 was detected. These results are in agreement with the XRD results. 3.3. CO-TPR Figs. 3 and 4 show the CO-TPR proles of Ce1 x Zrx O2 mixed oxides calcined at 500 and 900 C. For CeO2 and Ce1 x Zrx O2 mixed oxides calcined at 500 C (Fig. 3), two peaks are observed at lower and higher temperature regions. The rst peak in the low temperature region is the reduction of

surface oxygen and the other peak is the reduction of bulk oxygen [6,9,22]. In contrast, ZrO2 does not show the reduction behavior signicantly. The CO-TPR proles for Ce1 x Zrx O2 with x = 0.25, 0.5 and 0.75 are similar to that for pure CeO2 . However, it should be noted that reduction temperatures of the mixed oxides were shifted to higher temperatures when the Zr content was greater than 0.5. It might be postulated that this is a result of the irreducible structure of the material caused by a high Zr content. It should be pointed out that the structure of Ce0.25 Zr0.75 O2 is dominated by the tetragonal phase, whilst the structures of Ce0.75 Zr0.25 O2 and Ce0.5 Zr0.5 O2 are mainly in the cubic phase. This suggests that the tetragonal phase may be more difcult to reduce. This attribution is in good agreement with that reported by Fornasiero et al. [11]. The TPR proles for the samples calcined at 900 C are shown in Fig. 4. We found that reduction

Fig. 4. TPR prole of Ce1 x Zrx O2 mixed oxide catalysts calcined at 900 C with aging time = 50 h. The reducing gas contains 1% carbon monoxide in helium and a heating rate of 10 C/min was used. The gas ow rate was 50 ml/min.

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

227

Fig. 5. Light-off curves over Ce1 x Zrx O2 mixed oxide catalysts prepared via solgel technique with aging time = 50 h calcined at 500 C. Gas mixture contains 2.0% CH4 , 21.0% O2 and balance He. Total ow 100 ml/min.

of ceria was insignicant, since Ce1 x Zrx O2 with 0.25 x 0.5 samples still show two peaks of reduction. It is clear that the incorporation of Zr into CeO2 lattice promotes the redox properties in spite of a loss in surface area after calcination at high temperatures. 3.4. Catalytic activities for methane oxidation Figs. 5 and 6 show the light-off temperatures for methane oxidation over Ce1 x Zrx O2 mixed oxide catalysts calcined at 500 and 900 C, respectively. A higher activity is indicated by a lower temperature for 50% conversion of methane. The catalytic activity of the mixed oxide catalysts is higher than those of both pure ceria and pure zirconia. One can postulate that Zr which has been incorporated into CeO2 lattices to form a solid solution inuences the catalytic activity for methane oxidation. The experiments were carried out for comparison of the methane oxidation activity of the mixed oxides calcined at either 500 or 900 C. The results show that the activities of the mixed oxides calcined at 500 C are higher than those of the mixed oxides calcined at 900 C. This might be due to the fact that surface

areas and reducibility of the mixed oxides calcined at 900 C are low compared to those of the mixed oxides calcined at 500 C, as discussed earlier. Thus, it is our intention not to further investigate catalytic activities of the mixed oxides calcined at 900 C. The catalytic activity of the mixed oxides decreases with an increase in Zr loading. We found that Ce0.75 Zr0.25 O2 has higher activity than Ce0.5 Zr0.5 O2 and Ce0.25 Zr0.75 O2 . Although Ce0.75 Zr0.25 O2 has lower surface area than Ce0.5 Zr0.5 O2 and Ce0.25 Zr0.75 O2 , it shows the highest activity. The result suggests that the catalytic activity is virtually independent of surface area. The trend of catalytic activity altering with Zr loading is analogous to that of CO-TPR proles, as shown in Figs. 3 and 4. It was found that the Ce0.75 Zr0.25 O2 mixed oxide samples are more reducible than the Ce0.5 Zr0.5 O2 and Ce0.25 Zr0.75 O2 , respectively. Hence, it is apparent that catalytic activity is related to the redox properties of the mixed oxide. In addition, the catalytic activity of mixed oxide samples depends on the structural properties of the mixed oxide samples. It should be noted that the cubic-uorite-structure mixed oxide catalysts are more active than tetragonal and monoclinic structures [11]. In this work, we found that Ce0.75 Zr0.25 O2

228

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

Fig. 6. Light-off curves over Ce1 x Zrx O2 mixed oxide catalysts prepared via solgel technique with aging time = 50 h calcined at 900 C. Gas mixture contains 2.0% CH4 , 21.0% O2 and balance He. Total ow 100 ml/min.

Fig. 7. Plot of the linear range for the kinetic study.

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

229

Fig. 8. Calculated and experimental reaction rate (rCH4 ) vs. PCH4 at different PO2 feed values at 400 C for the Ce0.75 Zr0.25 O2 calcined at 500 C mixed oxide catalyst.

calcined at 500 C has the highest catalytic activity for methane oxidation. Similar ndings were reported elsewhere [12,20,21] for the oxidation of CO and C 3 H6 . 3.5. Kinetic studies The kinetic studies of methane oxidation were carried out on the highest catalytic activity material, Ce0.75 Zr0.25 O2 calcined at 500 C, at ve different temperatures between 400 and 600 C. The methane and oxygen feed concentrations were varied in the ranges 0.52.0 and 5.021.0% (balance helium), respectively. The kinetic data were obtained by using the initial rate method. As shown in Fig. 7, the linear range conversion of methane was attained when the conversion is less than 30%. As a result, the initial rate data were collected at conversions of less than 20% in all cases. The effect of the variation of the methane partial pressure at a constant oxygen partial pressure on the rate of reaction of methane oxidation is shown in

Fig. 8. It is apparent that the reaction rate is strongly inuenced by the partial pressure of methane. In contrast, the rate of reaction is slightly increased with increasing oxygen partial pressure, as shown in Fig. 9. The reaction orders of the catalyst were determined by using a power law model. It was found that the reaction order for methane concentration becomes 0.9 (ca. 1.0) and that for oxygen concentration becomes 0.1 (ca. 0). This result conforms with the dependence of the partial pressures of the reactants on the reaction rate, generally about a rst-order in methane and about a zero order in oxygen as reported elsewhere [2,5,2327]. The calculated apparent activation energy for this pseudo-rst-order model is 93.5 kJ/mol. This value falls in the range of activation energies reported for methane oxidation over Pd, Rh and Pt catalysts (71100 kJ/mol) [1]. Surface reaction mechanisms, classical Langmuir Hinshelwood models and EleyRideal model, were proposed for the catalytic methane oxidation in our study, as summarized in Table 3.

230

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

Fig. 9. Calculated and experimental reaction rate (rCH4 ) vs. PO2 at different PCH4 feed values at 400 C for the Ce0.75 Zr0.25 O2 calcined at 500 C mixed oxide catalyst.

Fig. 10. Arrhenius plot of kR constants for the Ce0.75 Zr0.25 O2 calcined at 500 C mixed oxide catalyst.

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233 Table 3 Kinetic models tested in tting the methane oxidation process Model 1 Reaction mechanism
1 k1 2 O2 + X O X; CH4 +Xk2 CH4 X; k3

231

Rate determining step


1 2 O2

Rate expression rCH4 = k1 PO2


1/2

Rate parameter Negative parameter

R2

+X OX

1 + k2 PCH4 k2 PCH4 1 + k1 PO2


1/2

CH4 X + O

Xproduct X + X; product Xproduct + X


k4

CH4 + X CH4 X

rCH4 =

Negative parameter

CH4 X + O X product X + X

rCH4 =

k3 k1 k2 PO2 PCH4 (1 + k1 PO2 + k2 PCH4 )2


1/2

1/2

k3 = 3.62 102 exp

O2 + Xk1 O2 X; CH4 +Xk2 CH4 X; CH4 X + O2 Xproduct X + X; product Xproduct + X


k4 k3

O2 + X O2 X

rCH4 =

k1 PO2 1 + k2 PCH4

12067 ; T 705.82 k1 = 7.74 102 exp ; T 291.25 k2 = 5.70 102 exp T Negative parameter

0.994

CH4 + X CH4 X CH4 X + O2 X product X + X

rCH4 = rCH4

k2 PCH4 1 + k1 PO2 k3 k1 k2 PO2 PCH4 = (1 + k1 PO2 + k2 PCH4 )2

Negative parameter k3 = 2.26 102 exp k1 k2 11819 ; T 332.65 = 4.99102 exp ; T 166.96 = 7.88 102 exp T 12053 = 8.39 exp ; T 1611.4 = 1.0 101 exp T

0.989

1 K1 2 O2 + X O X; CH4 + O

Xproduct X; product Xproduct + X


k3

k2

CH4 + O X product X

rCH4 =

1/2 k2 k1 PO2 PCH4 1/2 1 + k1 PO2

k2 k1

0.929

Note: units vary from model to model. The units must then be determined from the rate equation with rates expressed in mol/g/s and all pressures in kPa.

Models 1 and 2 are derived using the Langmuir Hinshelwood approach but with a difference in the adsorption step of oxygen. While model 1 assumes a dissociative adsorption of oxygen or monoatomic adsorption of oxygen, a non-dissociative adsorption of oxygen is proposed in model 2. The EleyRideal reaction mechanism (model 3), on the other hand, assumes the reaction between a dissociately adsorbed oxygen and gaseous methane. By performing multi-linear regression analysis, it was found that only LangmuirHinshelwood rate expressions, which assume surface reaction as a rate-determining step, and EleyRideal rate expression yielded positive parameters, as shown in Table 3.

However, it was found that the LangmuirHinshelwood kinetic rate expression which involved the surface reaction between dissociative adsorption of oxygen and non-dissociative adsorbed methane yielded the best correlation (R 2 = 0.994) for the conditions of our experiments, as shown in Figs. 8 and 9. The rate parameters of such a rate expression, as shown in Table 3, can be interpreted to show that oxygen adsorption is faster than methane adsorption, subsequently resulting in chemisorption of methane onto the catalyst surface. This result is in agreement with reports elsewhere [28,29]. The values of kinetic parameters on the kinetic rate expression versus temperature are expressed in Figs. 1012. In this study,

232

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

Fig. 11. Relation of KO2 constants vs. temperature for the Ce0.75 Zr0.25 O2 calcined at 500 C mixed oxide catalyst.

Fig. 12. Relation of KCH4 constants vs. temperature for the Ce0.75 Zr0.25 O2 calcined at 500 C mixed oxide catalyst.

the activation energy based on this surface reaction mechanism of methane oxidation over this catalyst is 100.8 kJ/mol, which is well inside the range reported for other transition mixed oxides [2,5,23,24].

4. Conclusions It can be concluded that this preparation method can give a satisfactory solid solution of ceria-zirconia

S. Pengpanich et al. / Applied Catalysis A: General 234 (2002) 221233

233

at low temperatures. The ceria-zirconia solid solutions still possess redox properties after loss of surface area from calcination at high temperatures. The incorporation of Zr into the CeO2 forming a solid solution could modify the redox behavior by not changing the structure of Ce1 x Zrx O2 with x 0.5 samples. Furthermore, the catalytic activity for methane oxidation of the catalysts was enhanced by the incorporation of Zr. We found that the catalytic activity for methane oxidation is more dependent on structure and redox properties than the BET surface area. The optimal formula that shows the highest activity is achieved with Ce0.75 Zr0.25 O2 . The kinetic study shows that the reaction rate depends on the methane concentration. The reaction mechanism can be expressed by the following LangmuirHinshelwood model: r = kR KO2 KCH4 PO2 PCH4 (1 + KO2 PO2 + KCH4 PCH4 )2
1/2 1/2

where kR is the reaction rate constant, and KO2 and KCH4 denote the adsorptiondesorption equilibrium constants of oxygen and methane, respectively, which the surface reaction of dissociated oxygen with methane chemisorption on the catalyst surface is the rate-determining step.

Acknowledgements The authors would like to thank Thailand Research Fund (grant number PDF/38/2541) for the nancial support. References
[1] J.H. Lee, D.L. Trimm, Fuel Process. Technol. 42 (1995) 339. [2] G. Saracco, G. Scibilia, A. Iannibello, G. Baldi, Appl. Catal. B 8 (1996) 229. [3] V. Labalme, E. Garbowski, N. Guilhaume, M. Primet, Appl. Catal. A 138 (1996) 93. [4] P. Artizzu, E. Garbowski, M. Primet, Y. Brulle, J. Saint-Just, Catal. Today 47 (1999) 83.

[5] B.W.-L. Jang, R.M. Nelson, J.J. Spivey, M. Ocal, R. Oukaci, G. Marcelin, Catal. Today 47 (1999) 103. [6] K. Otsuka, W. Ye, M. Nakamura, Appl. Catal. A 183 (1999) 317. [7] A. Trovarelli, C. de Leitenburg, M. Boaro, G. Dolcetti, Catal. Today 50 (1999) 353. [8] J. Kaspar, P. Fornasiero, M. Graziani, Catal. Today 50 (1999) 285. [9] P. Fornasiero, G. Balducci, R. Di Monte, J. Kaspar, V. Sergo, G. Gubitosa, A. Ferrero, M. Graziani, J. Catal. 164 (1996) 173. [10] C.E. Hori, H. Permana, K.Y.S. Ng, A. Brenner, K. More, K.M. Rahmoeller, D.N. Belton, Appl. Catal. B 16 (1998) 105. [11] P. Fornasiero, R. Di Monte, G. Ranga Rao, J. Kaspar, S. Meriani, A. Trovarelli, M. Graziani, J. Catal. 151 (1995) 168. [12] J.R. Gonzalez-Velasco, M.A. Gutierrez-Ortiz, J. Marc, J.A. Botas, M.P. Gonzalez-Marcos, G. Blanchard, Appl. Catal. B 22 (1999) 167. [13] C. Bozo, N. Guilhaume, E. Garbowski, M. Primet, Catal. Today 59 (2000) 33. [14] A. Trovarelli, F. Zamar, J. Llorca, C. de Leitenburg, G. Dolcetti, J.T. Kiss, J. Catal. 169 (1997) 490. [15] D. Terribile, A. Trovarelli, J. Liorca, C. de Leitenburg, G. Dolcetti, Catal. Today 43 (1998) 79. [16] S. Rossignol, Y. Madier, D. Duprez, Catal. Today 50 (1999) 261. [17] X. Bokhimi, A. Morales, O. Navaro, M. Portilla, T. Lopez, F. Tzompantzi, R. Gomez, J. Solid State Chem. 135 (1998) 28. [18] V. Meeyoo, P. Wright, in: Proceedings of the Regional Symposium on Chemical Engineering 1999, Songkhla, Thailand, 1999, p. C2-1. [19] D.J. Kim, J. Am. Ceram. Soc. 72 (1989) 1415. [20] M. Thammachart, V. Meeyoo, T. Risksomboon, in: Proceedings of the Chemeca 2000: The 28th Australian and New Zealand Chemical Engineering and Industrial Conference, Perth, WA, Australia, 2000, p. E-5.1. [21] M. Thammachart, V. Meeyoo, T. Rirksomboon, S. Osuwan, Catal. Today 68 (2001) 53. [22] M.-F. Luo, X.-M. Zheng, Appl. Catal. A 189 (1999) 15. [23] L.A. Isupova, G.M. Alikina, O.I. Snegurenko, V.A. Sadykov, S.V. Tsybulya, Appl. Catal. B 21 (1999) 171. [24] P. Ciambelli, V. Palma, S.F. Tikhov, V.A. Sadykov, L.A. Isupova, L. List, Catal. Today 47 (1999) 199. [25] R.F. Hicks, H. Qi, M.L. Young, R.G. Lee, J. Catal. 122 (1990) 280. [26] D.L. Trimm, C. Lam, Chem. Eng. Sci. 35 (1980) 1405. [27] A.V. Kucherov, N.V. Nekrasov, A.A. Slinkin, E.A. Katsman, S.L. Kiperman, Stud. Surf. Sci. Catal. 105 (1997) 1655. [28] A. Frennet, Catal. Rev. Sci. Eng. 10 (1974) 37. [29] T. Engel, G. Ertl, Adv. Catal. 28 (1979) 1.

You might also like