You are on page 1of 9

ISSN 0023-1584, Kinetics and Catalysis, 2008, Vol. 49, No. 5, pp. 667675. Pleiades Publishing, Ltd., 2008.

. Published in Russian in Kinetika i Kataliz, 2008, Vol. 49, No. 5, pp. 698706.

Catalytic Reforming of Methane by Carbon Dioxide over Nickel-Exchanged Zeolite Catalysts1


D. Hallichea, O. Cherifia, Y. B. Taaritb, and A. Aurouxb
Laboratoire de Chimie du Gaz Naturel, F.C., USTHB, BP 32, El-Alia, Bab-Ezzowr, Alger, Algeria b Institut de Recherches sur la Catalyse et lEnvironnement de Lyon, UMR5256, CNRS-Universite Claude Bernard Lyon 1, 2 av. Albert Einstein, 69626 Villeurbanne Cedex, France e-mail: dhalliche@yahoo.fr; aline.auroux@ircelyon.univ-lyonl.fr
Received May 25, 2007; in nal form September 16, 2007
a

AbstractA series of nickel-exchanged catalysts based on ZSM-5, USY, and Mordenite zeolites has been prepared by the ionic exchange method. The NiZeol catalysts have been characterized by XRD and BET. The exchange levels and nickel contents of the catalysts have been determined by chemical analysis. The acidity of the zeolite supports has been investigated using NH3 adsorption microcalorimetry. The number of acidic sites was found to decrease according to the following sequence: HUSY > HZSM-5 > HMOR. The temperature programmed reduction studies showed that the most reducible catalyst is NiZSM-5. The Ni-exchanged zeolites presented good catalytic performance in the methane reforming by CO2. At a temperature of 650C, CH4 conversions of 71 and 54% were achieved on NiUSY and NiZSM-5 respectively. At 400C, CO2 FTIR adsorption has shown that CO2 decomposes into CO and oxygen on NiZSM-5 which explains its reactivity at such a low temperature, while no decomposition of this probe molecule was observed on the NiUSY catalyst. The catalytic performance was found to vary in the following sequence at 650C: NiUSY > NiZSM-5 > NiMOR. Moreover, the catalytic performances were found to depend strongly on the CO2/CH4 ratio in the feed and were markedly improved for CO2/CH4 greater than 1. DOI: 10.1134/S002315840805011X
1

1. INTRODUCTION

During the last decade there has been increasing interest in methane reforming with carbon dioxide Eq. (1), not only as an alternative route for syngas production, but also as a method for the recycling or reusing of carbon dioxide, a greenhouse gas, by its use in chemical energy transmission systems [1]. CH 4 + CO 2 2CO + 2H 2 , H = 247.0 kJ/mol. (I)

Recent energy forecasts prepared by the international Energy Agency show that fossil fuels will predominate as an energy source during the 21st century, so that world emissions of CO2 will increase by 50% in 2010 when compared to 1990 levels [2]. The limitation of world carbon dioxide emissions, the main contributor to the greenhouse effect, is currently one of the most ambitious challenges in the eld of catalysis. The high endothermic reaction enthalpy associated with methane reforming by carbon dioxide and its reverse reaction make this process one of the most suitable for application in the storage of renewable energy sources. Literature concerned with the catalysis of this reaction has focused on four specic aspects: the intrinsic activity of the metallic phase, its stability towards car1 The

bon deposition, and the type of support suitable for improving the efciency of the catalysts. Nickel-based catalysts have shown excellent behavior in this reaction with activity comparable to noble metals [3, 4]. Ni has been supported on different materials such as Al2O3 [5], MgO, and TiO2 [6]. Most of these catalysts tend to deactivate by coke formation which is closely related to the catalysts structure and composition [7]. The use of supports with a low concentration of Lewis acid sites such as ZrO2 leads to an enhancement in activities and decrease in carbon deposition [7]. This carbon may be formed by the Boudouard reaction (Eq. (2)) or methane cracking (Eq. (3)) depending on experimental conditions. 2CO CH 4 C + CO 2 , C + 2H 2 , H = 173.0 kJ/mol, H = 75.2 kJ/mol. (II) (III)

text was submitted by the authors in English.

Zeolites are one of the most representative supports and are used widely in academic research and the chemical industry. The catalytic properties of faujasite, mordenite, and ZSM-5 zeolites are well known in hydrocarbon cracking and have been summarized in review papers. It was claimed that zeolites may play a major role as support of Ni catalysts for reforming reactions in view of their micropore structure and high surface area including a high afnity for carbon dioxide [8]. Higher dispersion of active metal species can be

667

668

HALLICHE et al.

Table 1. Overall chemical formulations of the exchanged zeolites Samples NiUSY NiZSM-5 NiMOR Ni, % wt 8.7 3.5 1.8 Lattice Si/Al 4.5 14.0 15.0 Global Si/Al 3.5 14.0 15.0 Exchange level, % 81.2 106.2 60.0 SBET , m2/g 600 330 500

expected also using zeolites supports [9]. Bhat and Sachtler [10] have studied Rh in NaY zeolites and obtained catalysts combining stability with high activity and selectivity. Moreover, Chang et al. [8] have examined a NiZSM-5 catalyst which was prepared by the solidstate reaction of zeolite and nickel nitrate and its applicability to dry reforming of methane, and the effects of addition of alkaline promoters such as K and Ca metals were examined. The results indicated that addition of Ca and K to the catalysts led to decreased coke deposition during the reforming reaction [8]. However, dry reforming of methane catalyzed by transition metalloaded zeolite is not well studied, probably due to their instability under hydrothermal conditions. The present study focuses on the activity in dry reforming of methane using several Ni loaded zeolites HUSY, HZSM-5, and HMordenite with the aim of investigating the inuence of the zeolite support. 2. EXPERIMENTAL 2.1. Catalysts Preparation The zeolite-supported catalysts have been prepared from the commercial HUSY (LZY-82 from Union Carbide), HZSM-5, and Mordenite (from DEGUSSA) zeolites, the characteristics of which are summarized in Table 1. These zeolites have been exchanged by a solution of nickel nitrate at 85C. The resulting Ni-exchanged zeolites have been dried at 110C and calcined at 500C. 2.2. Characterization of Catalysts The nickel content of the calcined catalysts has been determined by ICP ame spectroscopy, and the specic surface areas have been determined by the BET method using a Coultronics apparatus. The results are summarized in Table 1. The X-ray powder diffraction (XRD) for calcined catalysts was carried out in a Philips PW1050/81 automated powder goniometer ( Cu(K) = 1.54184 ). Microcalorimetric studies of the adsorption of NH3 were carried out using a Tian-Calvet heat-ow calorimeter (SETARAM C80) coupled to a volumetric apparatus [11]. The differential heats of adsorption versus adsorbate coverage have been obtained by measuring the heats evolved when doses of gas (NH3) were admitted sequentially onto the zeolite until the surface was

saturated by adsorbed species. The samples were pretreated under vacuum at 400C prior to adsorption. NH3 adsorption was performed at 150C. The calorimetric and volumetric data were used to plot the differential heats of adsorption as a function of coverage and the corresponding adsorption isotherms. The temperature programmed reduction (TPR) proles have been obtained using a conventional ow apparatus with a TCD detector. The sample was placed in a tubular quartz cell and pretreated under an Argon ow for 90 min at a temperature of 400C ( = 3 K min1). The samples were then reduced in an H2 (1%)/Ar mixture at increasing temperatures up to 700C ( = 3 K min1). CO2 adsorption on the samples was examined by FTIR spectroscopy using a Vector 22 spectrometer. Self-supporting wafers of 18 mm diameter were calcined overnight under O2 in the IR cell at 400C and then evacuated (2 h). After cooling to room temperature and adsorption of CO2, the catalyst was heated at 400C (1 h) and then evacuated at 25 and 400C. At each stage the infrared spectrum was recorded at room temperature. 2.3. Catalytic Testing For reaction tests, usually 100 mg of catalyst was pretreated under helium at 450C for 2 h into a tubular U-shaped quartz reactor. The reaction temperature, monitored by a thermocouple placed close to the reactor wall, increased from room temperature to 700C at the heating rate of 4 K min1. Reactant feeding was regulated by mass ow controllers (Brooks 5850 TR). The reaction mixture CH4, CO2 in the proportion 1 : 1 at a ow rate 20 ml min1 was used for reaction tests. This mixture was introduced into the sample by switching a four-way valve located at the reactor inlet. Reactants and products were analyzed in an on-line gas chromatograph (Delsi) equipped with a thermal conductivity detector (TCD) and Carbosieve-B column using helium as the carrier gas and an automatic sampler. 3. RESULTS AND DISCUSSION 3.1. Crystallinity XRD patterns were recorded for all samples in order to evaluate possible changes in the crystallinity of the zeolites. Comparison of XRD proles (Figs. 1a, 1b, and 1c) of zeolite peaks before and after Ni loading and the calcination procedure revealed a very small loss of
KINETICS AND CATALYSIS Vol. 49 No. 5 2008

CATALYTIC REFORMING OF METHANE BY CARBON DIOXIDE (a) Q, kJ/mol 220 200 180 160 140 120 100 80 60 40 20 0
HZSM-5

669

(a)

HUSY

NiUSY

HUSY HZSM-5 HMOR

(b)

200

400

600

800 1000 1200 1400 Amount of NH3, mol/g (b)

Amount of NH3, mol/g 1400 1200 1000 800 600 400 200 0 0.1

NiZSM-5

(c)

HUSY HZSM-5 HMOR

HMOR

0.2

0.3

0.4

0.5 0.6 Pressure, Torr

NiMOR

Fig. 2. (a) Differential heats of NH3 adsorption versus coverage on zeolites. (b) Isotherms of ammonia adsorption on zeolites.

10

20

30 2, deg

40

50

Fig. 1. (a) X-ray spectra for the NiUSY sample and the parent zeolite. (b) X-ray spectra for the NiZSM-5 sample and the parent zeolite. (c) X-ray spectra for the NiMOR sample and the parent zeolite.

crystallinity. However when inspecting Fig. 1a, we can observe a signicant difference in the relative intensities of the peaks in the 1012 2 region. For the HUSY, the signal at the lower degree has a higher intensity while for the NiUSY sample this is reversed. 3.2. Ammonia Chemisorption The relation between acidity and catalytic activity remains one of the most intriguing problems in zeolite catalysis. Adsorption microcalorimetry has been widely used to measure the strength of surface active sites using probe molecules. The method used to study acidity involves interaction with bases. Ammonia is a relatively strong basic probe molecule, with a small diameter, which allows it to titrate all the sites of the studied zeolites. The calorimetry results are illustrated in Fig. 2a which reports the variation of the heats of adsorption as a function of the coverage of the surface by successive doses of NH3.
KINETICS AND CATALYSIS Vol. 49 No. 5 2008

The general features of theses curves are similar to those reported in the literature [12, 13]. These curves show that the zeolites adsorb actively the NH3 probe molecule, which indicates their acidic character. The three zeolites show a plateau of adsorption heat at about 160 kJ mol1 for HMOR and about 140 kJ mol1 for both HZSM-5 and HUSY. It is well known that the strongest acidic sites correspond to particularly high heats [14]. In the case of HUSY, this value can be due to the presence of Lewis acid sites induced by the extraframework Al species (the Si/AI ratio as determined by 28Si and 27Al MAS NMR spectra is around 4.5 while the Si/Al ratio provided by chemical analysis is 3.5) [15]. This study leads to the following sequence of acid sites strength: HMOR > HZSM-5 ~ HUSY. Interestingly the HZSM-5 and HMOR curves display an inection point around 700 mol g1 characteristic of homogeneous Brnsted sites, while the HUSY sample shows a continuously decreasing curve giving rise to a much larger amount of NH3 adsorbed and conrming a signicant heterogeneity of the sites. At higher coverage, the differential heats decrease rapidly to reach values close to 70 kJ mol1, a value corresponding to the limit between physisorption and chemisorption of ammonia at the experiment temperature.

670 Table 2. Reducibility of Ni-exchanged zeolites Samples NiUSY NiZSM-5 NiMOR

HALLICHE et al.

Reduction level, % 48.0 68.0 60.6

Chemisorption isotherms are shown in Fig. 2b. This gure evidences that NH3 adsorption isotherms are of the Langmuir type. The HUSY zeolite seems to adsorb larger amounts of NH3 compared to HZSM-5 and HMOR. Based on the zeolite compositions listed in Table 1, it is possible to calculate the Brnsted acid sites concentration for HUSY, HMOR, and HZSM-5: 3694 mol g1, 1111 mol g1, and 920 mol g1, respectively. Particularly in the case of HUSY, this value is much higher than that obtained by microcalorimetry study (near 1400 mol g1) and could be explained by the presence of a large amount of extraframework aluminum species. Thus, this study leads to the following sequence of concentrations of acid sites as determined by NH3 adsorption: HUSY > HZSM-5 > HMOR. 3.3. Temperature-Programmed Reduction Temperature programmed reduction (TPR) of the catalysts was used to monitor the relative differences in the reducibilities of these catalysts and to identify structural changes induced by reduction. Indirect information about the localization of Ni-species in zeolites was obtained from TPR data [16]. In Table 2, quantitative results of TPR (in the 25700C temperature range) are summarized. Previously, it was determined for the NiY catalyst that the ease of Ni2+ reduction is decreasing in the order: supercages > sodalite cages > hexagonal prisms [17, 18], thus accounting for the observation of several peaks in the TPR proles. Indeed, NiUSY displays two large peaks at about 475 and 535C (Fig. 3). By similarity, with the proles reported for Ni2+ exchanged-Y
H2 consumptions, arb. units 2

zeolites [16], the observed peaks may be attributed to the reduction of Ni ions in different positions: in the supercages (SII sites) and sodalites cages (SII' or SI' sites) respectively. The reduction of Ni2+ in the sodalite cage to Ni0 or Ni+ was observed also for NiNaX zeolites [17]. The formation of Ni+ could arise from the reverse disproportionation reaction as postulated by Kasai et al. 2Ni+. [19]: Ni0 + Ni2+ The NiZSM-5 catalyst shows a TPR prole composed of three peaks with maxima centered around 326, 416, and 510C. The existence of three TPR peaks may suggest that nickel ions have different degrees of interaction with HZSM-5 zeolite [20]. The peak at nearly 326C may be attributed to NiO-like species which are so small as to escape X-ray diffraction detection and were formed at the outer surface of the zeolite as observed by Daza et al. [21]. The TPR peak with maxima at about 416C may be assigned to isolated nickel ions in charge compensation sites of ZSM-5 outside the zeolite channels. The third peak near 510C can be associated to the nickel particles located inside the zeolite channels and are more difcult to reduce. However, the TPR prole of the NiMOR sample shows a single reduction peak centered at about 513C which is assigned to Ni particles located inside the Mordenite main channels and which are also barely reduced. As can be seen in Table 2, the lower reducibility observed for the NiUSY catalyst (reduction level is about 48%) may be due to the presence of extraframework aluminum in the structure of HUSY (as shown in Table 1) which leads to the formation of nonreducible nickel aluminate and thereby decreases the reducibility of nickel as concluded from the studies on the reduction of NiY zeolites [22]. Thus, the reducibility of bivalent nickel species depends on their location in the sites of zeolites and their acidity. 3.4. Reactivity 3.4.1. Inuence of temperature. We have studied the reaction of dry reforming of methane by carbon dioxide in the presence of the various catalysts at temperatures ranging from 400 to 700C. The total ow of the reactant mixture was 20 mL min1, and the CO2/CH4 molar ratio was equal to 1.0. Variation of CO2 and CH4 conversions with temperature is shown in Figs. 4a and 4b for different nickel zeolites. At 400, NiZSM-5 is the only sample which presents signicant CH4 and CO2 conversions (about 14.0 and 13.0%, respectively). With the rise in temperature, the conversions of CO2 and CH4 increased, which conrms the endothermic character of this reaction [23]. The two reactants apparently facilitate each others dissociation, as reported previously by Erdhelyi et al. [24]. At 650C, as our results show, CO2 reforming of methane is efciently catalyzed by Ni-exchanged zeolites samples. For example, CO2 and CH4 conversions
KINETICS AND CATALYSIS Vol. 49 No. 5 2008

1 3

100

200

300

400

500 600 700 Temperature, C

Fig. 3. TPR proles of (1) NiZSM-5, (2) NiUSY, (3) NiMOR.

CATALYTIC REFORMING OF METHANE BY CARBON DIOXIDE

671

were respectively 70.2 vs. 71.2% on NiUSY and 56.3 vs. 54.3% on NiZSM-5, while the NiMOR catalyst showed only 20.2 vs. 22.1%, respectively. The relatively low activity for reforming in the NiMOR case may be explained essentially by the location of Ni ions in the main channels of Mordeniteas already assumed in the TPR studywhere nickel cations are not accessible to the reactant mixture. The low activity observed over this catalyst might be also due to the very strong acidity of the HMOR support (see above) and/or the low exchange level of the NiMOR catalyst (Table 1). However, it was reported [9] that Ni (2.6 wt %) HMOR, prepared by the impregnation method and reduced at 450C, achieved conversion of CO2 within 92% at 750C. This high conversion may be due to the location of nickel particles on the external surface of the Mordenite support, contrary to the case where nickel is introduced by ion exchange. In the same trend, Li et al. [25] have reported a conversion of methane around 74% at 800C over a Ni (4.4 wt %) ZSM-5 catalyst prepared by impregnation and reduced at 350C under H2. In another work [8], the Ni (2.4 wt %) ZSM-5 catalyst (Si/Al > 200) prepared by the solid-state exchange method and reduced by H2 at 700C achieved CO2 and CH4 conversions about 68%. It is therefore clear that the composition and the preparation method of the catalysts together with the pretreatment step play a predominant role. In the case of the reforming of an equimolecular mixture of CH4 and CO2, the CH4 and CO2 conversions should be equal [26]. However, CO2 conversions were higher than CH4 conversions in the case of NiUSY and NiZSM-5. The CO2 and CH4 conversions were respectively 76.0 vs. 74.1% on NiUSY and 61.7 vs. 55.2% on NiZSM-5 catalyst at 700C. The difference in CO2 and CH4 conversions may suggest that the main reaction was accompanied by several secondary reactions under the prevailing reaction conditions. The results could be a clear indication of the occurrence of reverse water gas shift as indicated by Bradford and Vannice [27] in their review article: H2O + CO. Wang and Au [28] attributed CO2 + H2 this result to CO2 complete dissociation: CO2 CO + O and CO C + O. However, with increasing temperatures, the rate of CO formation increased to the point that carbon monoxide formation rates equal 16.3 102 mol g1 h1 and 14.2 102 mol g1 h1 respectively on NiUSY and NiZSM-5 catalysts at 650C as opposed to 20.3 102 mol g1 h1 and 17.0 102 mol g1 h1 at 700C, respectively. The CO yield obtained on NiMOR is only 0.5 102 mol g1 h1 at 650C vs. 3.0 102 mol g1 h1 at 700C. The amounts of carbon in the incoming and outgoing ows are nearly in balance at 650C in the case of NiMOR and NiZSM-5. At higher temperatures, carbon deposition becomes more important, possibly due to methane
KINETICS AND CATALYSIS Vol. 49 No. 5 2008

xCH4, % 90 80 70 60 50 40 30 20 10 0 400 xCO2, % 90 80 70 60 50 40 30 20 10 0 400

NiUSY NiZSM-5 NiMOR

()

450

500

550

600 650 700 Temperature,

NiUSY NiZSM-5 NiMOR

(b)

450

500

550

600 650 700 Temperature,

Fig. 4. (a) Evolution of CH4 conversion vs. temperature for NiZeol catalysts (P = 1 atm, CO2/CH4 = 1, F = 1.2 L h1). (b) Evolution of CO2 conversion vs. temperature for NiZeol catalysts (P = 1 atm, CO2/CH4 = 1, F = 1.2 L h1).

decomposition. In addition, it was reported in the Bradford and Vannice review [27] that CO is the main contributor to the formation of graphitic carbon at low temperature, which agrees with the results obtained by Swaan et al. [29] and Luo et al. [30]. However, the better reducibility of NiZSM-5 compared to NiUSY does not result in improved performance, as observed by Swaan et al. [29]. Some authors even consider that excellent catalytic results can be obtained with catalysts containing few Lewis acid sites, which is the case of NiUSY where the Lewis acid sites are induced by extraframework aluminum atoms [7]. To summarize, the studied catalysts exhibit catalytic performances at 650C, which vary in the following sequence: NiUSY > NiZSM-5 NiMOR. These results, as a whole, suggest that the catalytic performance and the reaction mechanism depend strongly on the nature of the support, and in particular the acidity and the Si/Al ratio of the zeolite, in addition to the Ni loading and to the operating conditions of the dry-reforming reaction. 3.4.2. Effect of the CO/CH4 ratio. As was shown by Chang et al. [8], the conversion of either CO or methane depended strongly on the mixture composition of NiUSY. For example, the CH4 conversions increase

672 (a)
1610 1572

HALLICHE et al.

1 2 3 4 Absorbance (b)
2211 2107 2090

1 2 3 4 2600 2400 2200 2000 1800 1600 1400 Wave number, cm1
Fig. 5 (a) FTIR spectra (1) after admission of CO2, (2) following heating at 400C, (3) outgassing at room temperature and at 400C (4) on the NiUSY catalyst. (b) FTIR spectra (1) after admission of CO2, (2) subsequent heating at 400C, (3) outgassing at room temperature and at 400C (4) on the NiZSM-5 catalyst.

from 71.2 to 81.0% when the CO2/CH4 ratio increases from 1 to 3. Meanwhile, the CO2 conversion decreased sharply, from 70.2 to 28% when the ratio increases from 1 to 3. The carbon balance of the reaction is improved when the reactant mixture is enriched in CO2, with values of 71.7% when CO2/CH4 = 1 and 81.4% when CO2/CH4 = 3. Therefore, in order to avoid deactivation by coke deposition, it is preferable to operate with CO2/CH4 ratios greater than 1, in agreement with the literature [31]. The origin of this carbon deposition could be the reaction of dismutation of CO: 2CO C + CO2, or the decomposition of methane, which is strongly favored at high temperature: CH4 C + 2H2. 3.5. CO2 FTIR Adsorption We examined the adsorption of carbon dioxide on nickel-exchanged HZSM-5 and HUSY zeolites. The most sensitive method for following this process is IR spectroscopy. FTIR studies of CO2 adsorption were carried out at room temperature. A typical spectrum of CO2 adsorbed at room temperature is shown in Figs. 5a and 5b. The spectra of both NiUSY and NiZSM-5 show in Figs. 5a and 5b (1) a strong band at 2350 cm1 which is attributed to CO2 weakly adsorbed on NiZeol [32]. The intensity of this band appeared to be CO2 pressure

dependent. In each case, the reference spectrum is that taken just prior to CO2 admission. It is well known that seven basic types of surface species can be formed by chemisorption of CO2 onto metal oxides, these include carbonates (free carbonates, unidentates, bidendates, and bridged), bicarbonates, carboxylates, and formates [33]. Each of these have characteristic vibrational stretching frequencies in the range from 1200 to 1850 cm1. Thus, after exposure of the NiUSY sample to carbon dioxide at 400C, IR bands within the 1800 to 1300 cm1 range (1572 cm1, 1610 cm1 as a shoulder, 1435 and 1355 cm1) were observed as can be seen in Fig. 5a. According to the literature data, these bands are likely due to the surface carbonates and bicarbonates species [26]. Species like carbonates are usually formed on the alumina supported catalysts under reaction conditions, and they have been proposed in some cases as reaction intermediates [34]. However, no CO2 dissociation occurred to form adsorbed CO on the NiUSY catalyst. This result is in good agreement with those reported in the literature [32]. After NiZSM-5 exposure to carbon dioxide at 400C, CO2 dissociation occurred at 400C as inferred from Fig. 5b (curve 2). It was evidenced by the presence of the bands at 2211, 2136, 2097 cm1. According to the literature [35], the 1R band found in the CO stretching region near 2200 cm1 was assigned to the CO coordinate bound to unsaturated cations and was dependent on the nature of the transition metal cation and the Si/Al ratio of the zeolite. It is well known that CO2 can adsorb on the Ni species from the gas phase and dissociate to oxygen and CO according to the following equation over the NiZSM-5 catalyst [26]: CO2 CO(ads) + O(ads). This nding is in good agreement with the studies carried out by Martin et al. concerned with CO2 adsorption on Ni/SiO2 [36]. The presence of the band at 2211 cm1, which may be assigned to Ni2+-CO, could be explained by the high concentration of nickel cations (overexchange level on HZSM-5 zeolite as shown in Table 1). However, the intensity of this band decreases after evacuation at room temperature. In the same trends, the IR bands at 2136 and 2097 cm1 shown in NiZSM-5 spectra may be attributed to the formation of Ni+(CO)2 gem dicarbonyl which converted to the linear Ni+CO (appears at 2107 cm1) species during evacuation at room temperature. The NiZSM-5 spectra of Fig. 5b (curve 2) present also bands near 1375 and 1445 cm1 which disappear immediately after desorption experiments at room temperature as may be seen from Fig. 5b (curve 3) and a band at 1621 cm1 which remains unaffected after outgassing. Moreover, all bands disappear completely by outgassing at 400C. The band at 1621 cm1 may be
KINETICS AND CATALYSIS Vol. 49 No. 5 2008

1435 1445

2136

1621

1375

1355

CATALYTIC REFORMING OF METHANE BY CARBON DIOXIDE

673

ascribed to carbonate structures (asymmetric stretching mode of HCO 3 ) and/or the formation of water adsorbed, and the 1445 cm1 band may be associated to the symmetric stretching mode of HCO 3 . According to the literature data, the band near 1375 cm1 might be assigned to a surface formate species following the reaction [2] CO 3 + 2H HCO 3 + 2H
2

Reactants adsorption and dissociation: CH4(g) CH4(ads) CO2(ads) CO2(ads) CH4(ads), CO2(ads), CO(ads) + O(ads). 2H2, CO(ads) CO(g). CHx(ads) + (4 x)H(ads),

Recombination and products evolution: 4H(ads) C(ads) + O(ads)

HCO 2 + OH , HCO 2 + H 2 O.

The overall reaction is that of true dry reforming: CH4 + CO2 2CO + 2H2. However, CO2 may adsorb over the basic oxide ion from the support to generate carbonates which in turn may convert into formates in the presence of adsorbed H atoms. The formates would ultimately release CO and generate hydroxide ions. Two of the latter species would then release H2O and regenerate the initial oxide ion. This scheme may be depicted as follows: Formation of carbonates: CO2(ads) + O2
2

In other words, the CO2 dissociation over the NiZSM-5 catalyst at 400C is in good agreement with the obtained results in the methane reforming reaction. This reaction involves CO2 adsorption/dissociation over the catalyst interface. Solymosi et al. [37] proposed that the dissociation of CO2 could be aided by hydrogen species generated in CH4 decomposition to give formate species: CO2(ads) + H(ads) HCOO(ads).

It is well known that the dry reforming reaction on Ni-catalysts can take place through a bifunctional mechanism in which the support participates in the activation of carbon dioxide and has been already proposed by Nakamura et al. [34] for rhodium supported on oxides such as TiO2, Al2O3, or MgO. However, they did not propose any concrete reaction steps. Lercher and Bitter [38] also proposed that the dry reforming of methane over Pt/ZrO2 catalysts proceeds via a bifunctional mechanism with formate-like adsorbed intermediates that may be located on the support surface. They have suggested that catalysts are not efcient for this reaction when the formation of carbonates is not possible. Activation mechanisms that proceed through formate-like intermediates have also been proposed for other reactions such as the reverse-water-gas-shift. Finally, from the literature data and the results above, one may ascribe the formation of the syngas produced in the dry reforming of methane to three distinct schemes occurring more or less simultaneously. The rst would involve, as suggested by Luo et al. [30], the adsorption and dissociation of CH4 on nickel species to give adsorbed H atoms and CHx fragments, which, in turn, would ultimately dissociate to adsorbed carbon and hydrogen atoms. Likewise, concomitantly, CO2 would also dissociate into adsorbed CO and O atoms. Adsorbed CO would thermally evolve, and, further, recombination of adsorbed O and C atoms would release additional CO and recombination of H atoms would give rise to the expected hydrogen. This overall scheme, which may be called true dry reforming, may be represented by the following step sequence:
KINETICS AND CATALYSIS Vol. 49 No. 5 2008

CO 3 (ads).

Reduction of carbonates to formates: CO 3 (ads) + 2H(ads) Evolution of products: HCO 2 2 OH s


HCO 2 (ads) + OH s .

CO + OH, H2O + O2. H2O + CO + H2 ,

The overall scheme may be written as CH4 + CO2 where part of the expected hydrogen would undergo the reverse of the WGS reaction occurring simultaneously with the deposition of carbon originating from methane. The third scheme would involve OH species, either present in the fresh catalyst or formed via the decomposition of adsorbed formates. This basic hydroxide would capture CO2 to hydrogeno-carbonates with the decomposition of methane. The hydrogeno-carbonates would react with adsorbed hydrogen to release water and generate formates. The latter would release CO and regenerate the hydroxide species. This scheme may be represented as follows: Formation of the hydrogeno-carbonates: CO2 + OH HCO 3 (ads) + 2H(ads)

HCO 3 . HCO 2 (ads) + H2O.

Reduction of the hydrogeno-carbonates:

674

HALLICHE et al.

Decomposition of formate species and syngas formation: HCO 2 (ads)

REFERENCES
1. Ashcroft, A.T., Cheetham, A.K., Green, M.L.H., and Vernon, P.D.F., Nature, 1991, vol. 352, p. 225. 2. Ferreira-Aparicio, P., Rodriguez-Ramos, I., Anderson, J.A., and Guerrero-Ruiz, A., Appl. Catal., A, 2000, vol. 202, p. 183. 3. Lemonidou, A.A. and Vasalos, I.A., Appl. Catal., A, 2002, vol. 228, p. 227. 4. Wang, J.B., Wu, Y.S., and Huang, T.J., Appl. Catal., A, 2004, vol. 272, p. 289. 5. Halliche, D., Bouarab, R., Cheri, O., and Bettahar, M.M., Catal. Today, 1996, vol. 29, p. 373. 6. Xu, Z., Li, Y., Zhang, J., Chang, L., Zhou, R., and Duan, Z., Appl. Catal., A, 2001, vol. 210, p. 45. 7. Montoya, J.A., Romero-Pascal, E., Gimon, C., Del Angel, P., and Monzon, A., Catal. Today, 2000, vol. 63, p. 71. 8. Chang, J.-S., Park, S.E., and Chon, H., Appl. Catal., A, 1996, vol. 145, p. 111. 9. Inoue, H., Hatanaka, N., Kidena, K., Murata, S., and Nomura, M., J. Jpn. Pet. Inst., 2002, vol. 45, no. 5, p. 314. 10. Bhat, R.N. and Sachtler, W.M.H., Appl. Catal., A, 1997, vol. 150, p. 279. 11. Halliche, D., Cheri, O., and Auroux, A., J. Therm. Anal. Calorim., 2002, vol. 68, p. 997. 12. Shen, J., Cortright, R.D., Chen, Y., and Dumesic, J.A., J. Phys. Chem., 1994, vol. 98, p. 8067. 13. Auroux, A. and Occelli, M.L., Stud. Surf. Sci. Catal., 1994, vol. 84, p. 693. 14. Auroux, A. and Ben. Tarit, Y., Thermochim. Acta, 1987, vol. 122, p. 63. 15. Halliche, D., Cheri, O., and Auroux, A., Thermochim. Acta, 2005, vol. 434, p. 125. 16. Pawelec, B., Mariscal, R., Navarro, R.M., Martin, J.M.C., and Fierro, J.L.G., Appl. Catal., A, 2004, vol. 262, p. 155. 17. Pawelec, B., Daza, L., Fierro, J.L.G., and Anderson, J.A., Appl. Catal., A, 1996, vol. 145, p. 3007. 18. Bendezu, S., Cid, R., Fierro, J.L.G., and Agudo, A.L., Appl. Catal., A, 2000, vol. 197, p. 47. 19. Kasai, P.H., Bishop, R.J., and McLeod, D., J. Phys. Chem., 1978, vol. 82, p. 279. 20. Romero, M.D., de Lucas, A., Calles, J.A., and Rodriguez, A., Appl. Catal., A, 1996, vol. 146, p. 425. 21. Daza, L., Pawelec, B., Anderson, J.A., and Fierro, J.L.G., Appl. Catal., A, 1992, vol. 87, p. 145. 22. El Solh, T., Jarosch, K., and de Lasa, H.I., Appl. Catal., A, 2001, vol. 210, p. 315. 23. Stagg-Williams, S.M., Noronha, F.B., Fendley, G., and Resasco, D.E., J. Catal., 2000, vol. 194, p. 240. 24. Erdhelyi, A., Csernyi, J., and Solymosi, F., J. Catal., 1993, vol. 141, p. 287. 25. Li, W.Y., Feng, J., and Xie, K.C., Pet. Sci. Technol., 1998, vol. 16, p. 539. 26. Luo, J.Z., Gao, L.Z., Ng, C.F., and Au, C.T., Catal. Lett., 1999, vol. 62, p. 153. 27. Bradford, M.C.J. and Vannice, M.A., Catal. Rev. Sci. Eng., 1999, vol. 41, no. 1, p. 1.
KINETICS AND CATALYSIS Vol. 49 No. 5 2008

CO(ads) + OH.

This overall scheme involves also the reverse reaction of the WGS and could be accompanied by carbon deposition from either methane or CO via the Boudouard reaction. CONCLUSIONS In this study we investigated catalysts based on HZSM-5, HUSY, and HMordenite. XRD experiments did not reveal signicant modication of the zeolite structure after incorporation of nickel. The acidity of the parent zeolites was studied by NH3 adsorption microcalorimetry. The most acidic support is Mordenite, but the HUSY zeolite presents a richer population of Lewis acid sites. Temperature-programmed reduction of these catalysts has shown that reducibility depends on the type of interaction between nickel and the zeolite and on the initial position of nickel in the zeolite; the reducibility of these catalysts was found to vary as follows: NiZSM5 > NiMOR > NiUSY. The catalytic testing of these materials in the reaction of methane reforming by carbon dioxide at various temperatures (400700C) has evidenced differences in catalytic activity, according to the sequence NiUSY > NiZSM-5 > NiMOR at 650C. We have also observed that the NiZSM-5 catalyst gives rise to noticeable catalytic activity at a relatively low temperature (400C), contrary to the other two catalysts. This study has shown the occurrence of many secondary reactions such as methane decomposition C + 2H2), the reversed water gas shift reac(CH4 tion, and CO disputation. The study of the impact of a variation in the CO2/CH4 gas ratio has revealed that a CO2-enriched reactant mixture makes it possible to avoid catalyst deactivation by coke deposition. It has also been shown that better performance is achieved for lower values of the total reactant ow. CO2 FTIR adsorption has shown that CO2 decomposes into CO and O at 400C on NiZSM-5 which explains its reactivity at such a low temperature, while no decomposition of this probe molecule was shown on the NiUSY catalyst. This study also reveals adsorbed intermediates species (formate, bicarbonates, and carbonates) whose formation constitutes a key step in the development of the reaction. ACKNOWLEDGMENTS This work was supported by CMEP project (no. 00 MDU 459). We are grateful for support received.

CATALYTIC REFORMING OF METHANE BY CARBON DIOXIDE 28. Wang, H.Y. and Au, C.T., Appl. Catal., A, 1997, vol. 155, p. 239. 29. Swaan, H.M., Kroll, C.H., Martin, G.A., and Mirodatos, C., Catal. Today, 1994, vol. 21, p. 571. 30. Luo, J.Z., Yu, Z.L., Ng, C.F., and Au, C.T., J. Catal., 2000, vol. 194, p. 198. 31. Mark, M.F. and Maier, W.F., J. Catal., 1996, vol. 164, p. 122. 32. Sigl, M., Bradford, M.C.J., Knozinger, H., and Vannice, M.A., Top. Catal., 1999, vol. 8, p. 211. 33. Rakic, V.M., Hercigonja, R.V., and Dondur, V.T., Microporous Mater., 1999, vol. 27, p. 27.

675

34. Nakamura, J., Aikawa, K., Sato, K., and Uchijima, T., Catal. Lett., vol. 25, p. 265. 35. Chen, Y.G., Tomishige, K., Yokoyama, K., and Fujimoto, K., J. Catal., 1999, vol. 184, p. 479. 36. Turlier, P., Pereira, E.B., and Martin, G.A., Proc. Int. Conf. on CO2 Utilization, Bari, 1993, p. 119. 37. Solymosi, F., Kutsan, Gy., and Erdohelyi, A., Catal. Lett., 1991, vol. 11, p. 149. 38. Bitter, J.H., Seshan, K., and Lercher, J.A., J. Catal., 1997, vol. 171, p. 279.

KINETICS AND CATALYSIS

Vol. 49

No. 5

2008

You might also like