Sei sulla pagina 1di 12

Applied Catalysis A: General 248 (2003) 105116

Propane dehydrogenation over a Cr2 O3/Al2 O3 catalyst: transient kinetic modeling of propene and coke formation
J. Gascn, C. Tllez, J. Herguido, M. Menndez
Department of Chemical and Environmental Engineering, University of Zaragoza, 50009 Zaragoza, Spain Received 17 December 2002; received in revised form 11 February 2003; accepted 11 February 2003

Abstract The kinetics of propane dehydrogenation to produce propene over a Cr2 O3 /Al2 O3 catalyst has been investigated over the temperature range of 525575 C at atmospheric pressure. The reaction rate of coke formation and its inuence over catalyst deactivation has been studied. A LangmuirHinshelwood mechanism provides the best t for the reaction, while a monolayermultilayer mechanism is proposed to model the coke growth. Furthermore, this model was able to predict coke formation under conditions far from those employed in the experiments used to obtain the kinetics. 2003 Elsevier Science B.V. All rights reserved.
Keywords: Reaction kinetics; Reaction mechanism; Propane dehydrogenation; Cr2 O3 /Al2 O3 catalysts; Catalyst deactivation

1. Introduction Forecasts for basic alkenes [1] indicate that the growth rate of propene demand will surpass that of ethene until the year 2005, in spite of the fact that ethene consumption is anticipated to double by 2015. Consequently, naphtha cracker operators are seeking options that enable process exibility and raise propene/ethene production ratios without affecting ethene output. Therefore, it is essential to study now processes involving the production of light alkenes from alkanes. Until the last 20 years [2], the most important ways to obtain propene were the steam cracking of hydrocarbon feedstocks and renery conversion processes (e.g. uid catalytic cracking, visbreaking and coking). The increasing demand for propene derivatives throughout
Corresponding author. Tel.: +34-976-761152; fax: +34-976-762142. E-mail address: qtmiguel@posta.unizar.es (M. Men ndez). e

the 1980s, especially for polypropylene, outstripped the availability from these established sources, and processes for the on-purpose production of propene by the dehydrogenation of propane from natural LPG elds have been developed commercially. Moreover, there are other paths to obtain the alkene like oxidative dehydrogenation [36] or metathesis [7]. Propane dehydrogenation (1) is an equilibrium limited and highly endothermic reaction that is generally carried out at 530630 C and atmospheric pressure using platinum [8] or chromium [9] catalysts. C3 H8 C3 H6 + H2 (1)

The main problems associated with dehydrogenation of alkanes are the formation of coke (coking) and low chain hydrocarbons (cracking). Coking leads to catalyst deactivation and cracking decreases reaction selectivity. Oxides obtained from co-precipitation are widely used in chemical industries as catalysts for the selective oxidation of hydrocarbons and aromatics,

0926-860X/$ see front matter 2003 Elsevier Science B.V. All rights reserved. doi:10.1016/S0926-860X(03)00128-5

106

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o Table 1 Catalyst specications Manufacturer Commercial name Shape Diameter (mm) High (mm) Specic surface (m2 /g) Specic pore volume (cm3 /g) Major components (%) Cr2 O3 Al2 O3 MgO Nikki Chemical N401 AG Pellet 5.0 4.9 214 0.06 14.2 65.4 2.2

dehydrogenations and isomerations due to their acid and dehydrogenating sites. Cr2 O3 /Al2 O3 as a dehydrogenating catalyst has been known since the pioneering work of Frey and Huppke [10,11]. Subsequently, numerous papers have been published in the literature on CrOx /SiO2 , CrOx /Al2 O3 and CrOx /ZrO2 catalysts. It is supposed that active sites for dehydrogenation are Cr3+ , although Cr2+ also has an inuence [12]. Moreover, adding an alkaline metal to the catalyst structure can benet its selectivity, reducing the cracking reactions and the coke formation. Although this reaction has been widely studied, the majority of these works are partial kinetics studies [8,9,13,14] or based on the oxidation state or on the nature of chromium in the reaction [1517]. Only Larsson et al. [18] have combined the kinetics of propane dehydrogenation using a simple power law model with deactivation produced by coke formation. In this work, we present a mechanistic kinetic model for the dehydrogenation, the coke formation and its inuence over catalyst deactivation using a Cr2 O3 /Al2 O3 catalyst, in order to perform this reaction in a two zone uidized bed reactor [19]. To design and optimize the most suitable reactor, a kinetic model that describes accurately the behavior of the system under a variety of conditions is required.

conditions. The thermogravimetric system used was a commercial set from C.I. Electronics Ltd., Model K2 vacuum head, designed for a maximum weight of 5 g. Reproducibility of 10 g is guaranteed. Weight measurement was registered in a computer with a speed of 100 data/min. The reactor feed contained propane diluted in nitrogen. All the streams were mass ow controlled (Brooks). The range of operating conditions was as follows: temperature, 525575 C; catalyst mass, 20 mg; total feed ow rate, 100 cm3 (STP)/min. 2.3. Fixed bed reactor

2. Experimental 2.1. Catalyst In all the experiments performed, a commercial Cr2 O3 /Al2 O3 catalyst, supplied by Nikki Chemical, was used. Table 1 shows the most important specications of this dehydrogenation catalyst. Before the reaction, the catalyst was: ground and sieved to a particle size of 160250 m; calcined in air at 650 C; treated under seven reactionregeneration cycles during 30 min each at 600 C to obtain an aged and stable catalyst [20,21]. 2.2. Thermogravimetric system Coke formation experiments were performed in a thermogravimetric system, under several different

The kinetic reaction study was carried out using an 8 mm internal diameter tubular quartz reactor inside an electrical furnace. The temperature was measured by a thermocouple inside a quartz thermowell, at the center of the catalyst bed, which contained 200 mg of Cr2 O3 /Al2 O3 catalyst. A PID controller maintained temperature variations within 0.5 C of the set point. The reactor feed contained propane diluted in argon. All the streams were mass ow controlled (Brooks). The temperature was varied from 525 to 575 C. The exit gases were analyzed by online gas chromatography (CE Instruments Model GC-8000TOP) equipped with two detectors (TCD and FID) and two columns (Chromosorb P AW 23% SP-1700 80/100 and molecular sieve 10 80/100). 2.4. Fluidized bed reactor Experiments using a uidized bed reactor were performed to check the ability of the proposed

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o

107

model to predict coke formation. A 2.5 cm internal diameter tubular quartz reactor, with a porous quartz plate as gas distributor, was used. The temperature was measured by a thermocouple inside a quartz thermowell, at the center of the catalyst bed, which contained 25 g of catalyst. The reactor feed contained propane diluted in argon. All the streams were mass ow controlled (Brooks). The exit gases were analyzed by online gas spectrometry (Balzer, Model Omnistar 200) equipped with an ionic detector (Faraday Cup) and a secondary electron multiplier (SEM). The operating temperature was between 575 and 600 C, and the total feed ow rate was 375 ml (STP)/min. 2.5. Fitting Once the kinetics have been calculated, the suitability of the models to represent the kinetic data can be

assessed. To choose the best kinetic model between those proposed, the model selection criteria (MSC) [22] has been used for each of the next described models. The MSC is computed as: MSC = ln
l j=1 (Yobsj l j=1 (Yobsj

Yobs )2 Ycalj )2

2p j

(2)

where j is the number of experimental points, p the number of parameters and Yobs the weighted mean of the experimental observations. The MSC is useful because it takes into account the number of parameters of a given model and, therefore, allows a comparison of different models. Fittings were performed with commercial software Scientist that employs the LevenbergMardquardt algorithm. The plug ow reactor was approached by a nite difference approach.

Fig. 1. Molar ow rate of products in the reactor outlet stream for different temperatures. W/F = 10 g h/mol; QT = 150 cm3 (STP)/min; propane/argon ratio = 1.

108

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o

3. Experimental results 3.1. Dehydrogenation of propane Several sets of experiments have been performed in the xed bed reactor, varying the feed composition and temperature. Fig. 1 shows the outlet molar ow of all species involved in the reaction. The main products are propene and hydrogen, but ethene and methane appear also as cracking products. Propane conversion and selectivity to propene (Fig. 2) decreases with time due to deactivation by coke deposition that apparently only affects the dehydrogenation reaction, but not the cracking. The effect of temperature upon propane conversion and selectivity is very marked. Conversion increases quickly with temperature, dehydrogenation and cracking are both very endothermic reactions and, as a result, a small rise in temperature produces a large increase in conversion. The effect of temperature on catalyst deactivation is also very striking; activity decreases initially faster at high temperatures. Selectivity to propene (Fig. 2) decreases with temperature

probably because its effect is greater over cracking than over the dehydrogenation reaction. Fig. 3 shows the effect of the feed composition for a given spatial time. The propane conversion decreases slightly when propane concentration increases, which could suggest an order of propane reaction slightly lower than 1. The selectivity to propene also decreases slightly with propane concentration. 3.2. Coke formation Thermogravimetric experiments have been carried out, varying the feed composition and temperature. Fig. 4 shows the variation with time of coke formation at different temperatures. The trend of coke content with time has two zones, more or less differentiated depending on the temperature. At 525 and 550 C, in the rst initial zone, the rate of coke formation decreases with time and is higher than in the second one, where the rate of formation remains nearly constant. At 575 C, the rst zone almost disappears. These results suggest that in the initial zone coke formation deactivates itself when the catalyst surface becomes coked, after this a residual coking

Fig. 2. Conversion of propane and selectivity to propene as a function of time for different temperatures. W/F = 14 g h/mol; QT = 150 cm3 (STP)/min; propane/argon ratio = 0.5.

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o

109

Fig. 3. Conversion of propane and selectivity to propene at different propane partial pressures. T = 550 C; QT = 150 cm3 (STP)/min.

Fig. 4. Coke content in catalyst as a function of time for different temperatures. W = 20 mg; propane/argon ratio = 3; QT = 100 cm3 (STP)/min.

110

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o

Fig. 5. Coke content in catalyst as a function of time for different propane partial pressures. W = 20 mg; T = 575 C; QT = 100 cm3 (STP)/min; pressure = 1 bar.

activity remains constant in the second zone. Similar trends have been pointed out by Pea et al. [20,21] for butane dehydrogenation on Cr2 O3 /Al2 O3 catalyst and by van Sint et al. [8] for propane dehydrogenation on Pt/ -Al2 O3 catalyst. A very strong inuence of the reaction temperature on coke formation can also be observed in Fig. 4. In the second zone, the coke formation rate increases markedly with temperature. At the nal time shown in Fig. 4, ve times more coke is formed at the highest temperature studied (575 C) than at the lowest (525 C). In Fig. 5, the inuence of the propane concentration in propaneargon mixtures is shown. There are no appreciable effects in the conditions studied. The apparent reaction order of the coking process with respect to the propane concentration is very low. This effect has also been observed by van Sint et al. [8]. This is a counter-intuitive result, not easy to explain. It may be hypothesized that due to the strong propene adsorption (as will be shown later) the surface coverage is close to unity in all the experimental conditions, and thus the coke formation by ethane polymerization results independent of the propane concentration in the gas phase.

4. Kinetic modeling The kinetic modeling has been developed following several steps. In the rst step, the data obtained in the xed bed reactor were tted to a power law model and different mechanistic models assuming a simple deactivation model. In the second step, thermogravimetric experiments were tted to different coke formation models. Statistical parameters of the tting were used to discriminate between the proposed models in both the steps. Finally, in the third step, the best kinetic models obtained previously were used to calculate the effect of coke formation over deactivation and the kinetic parameters. 4.1. Dehydrogenation of propane Different models have been tested in order to obtain the best t for the dehydrogenation kinetics: a power law model and LangmuirHinshelwood mechanisms. In all cases, the cracking reaction has been tted to a power law rate expression with rst order relative to propane.

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o Table 2 Kinetic equations for the proposed reaction mechanisms Model Power-law LHHW 1 LHHW 2 Kinetic equation

111

PC 3 H 6 P H 2 Keq KC3 H8 k1 [PC3 H8 (KC3 H6 /KC3 H8 )(PC3 H6 PH2 /Keq )] (rC3 H8 ) = 1 + K C 3 H 8 P C 3 H 8 + KC 3 H 6 P C 3 H 6 k1 [PC3 H8 (PC3 H6 PH2 /Keq )] (rC3 H8 ) = 1 + KC3 H6 PC3 H6 (rC3 H8 ) = k1 PC3 H8

In this work, several rate equations have been tested, as shown in Table 2. The rst corresponds to a power-law equation. The others are derived from LangmuirHinshelwood mechanisms. The rst, LHWW-1, supposes that adsorption equilibrium constants for propane and propene have the same order. The second, LHHW-2, assumes that propane adsorption is negligible. After combining reaction and adsorption equilibrium and performing the active sites balance, the equations shown in Table 2 for each mechanism are obtained. To choose between the reaction kinetics proposed, it was decided, as a rst approach, to t all kinetic data using a simple and empirical deactivation model. This avoids potential errors resulting from extrapolating data to initial time to obtain initial rates. The model combines exponential activitytime obtained by Dumez and Froment [23] (a = exp(Cc )) and the coketime relationship proposed by Voorhies [24] (Cc = Atn ). Thus, the activitytime relationship is given by: = exp(t n ) (3)

value obtained for propene adsorption is very high. Furthermore, the activation energy of thermal cracking is greater than for the propane dehydrogenation (Table 5), in agreement with what is expected for a gas phase phenomenon versus a catalytic process. A small standard deviation of kinetic parameters is obtained due to the large quantity of experimental data, which contributes to more precise results from a statistical point of view. In all the temperature range, propane conversion and selectivity to propene was well predicted with the model. 4.2. Coke formation kinetics The models used in this work to describe coke formation were previously developed by Pea et al. [20,21], Romero et al. [26], Rodrguez et al. [27] and van Sint et al. [8]. This model, called the monolayer multilayer coke growth model (MMCGM) was rst formulated by Nam and Kittrell [28] and generalized
Table 3 Model selection criteria for studied kinetics Model Reaction kinetics (Table 2) Potential LHHW 1 LHHW 2 MSC 5.03 5.45 5.59 R2 0.98 0.99 0.99 0.99 0.99 0.99 0.87 0.99

where is given by = 0 eEa (1/T 1/T0 ) and n a xed value which does not depend on temperature. The use of this activitytime relationship allows us to employ all the conversionselectivity data and not only those from the extrapolation to 0 time. As may be seen in Table 3, the best tting obtained (best MSC value) was for the LangmuirHinshelwood mechanism (LHHW-2) that accounts for inhibition by propene due to adsorption. This is the chosen model and agrees with the equation proposed by Wan and Chu [25] for a ZnO/silicalite catalyst. Mechanistic model is better than the potential model. In agreement with the chosen model, the equilibrium constant

Coke formation kinetics (Table 4) 1 6.67 2 6.50 3 6.53 Deactivation model Potential Exponential 3.05 5.23

112

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o

by Corella and Monzn [29]. In the generalized model, coke content in the catalyst versus time can be described taking into account a simultaneous formation of coke over the surface of the catalyst and a multilayer coke deposition. Therefore, the total coke formation velocity must be described as the addition of monolayer and multilayer coke growth. dC dCm dCM = + dt dt dt with rCm = and rCM = dCM n = k2 Cm (Cmax Cm )m , dt dCm = k1 (Cmax Cm )h dt (4)

where Cm and CM are the coke concentration in monolayer and multilayer, Cmax is the maximum coke concentration in monolayer, k1 and k2 are the kinetic co efcients, given by ki = ki0 eEai (1/T 1/T0 ) , with T0 being the reference temperature, h the kinetic order for monolayer coke formation, and n and m the kinetic orders for coke formation in multilayer. The same values tested by Monzn et al. [30] for the kinetic orders (h, n and m) in previous kinetic studies on coke formation have been used. Table 4 shows the integrated equations for those values. Model 1, which
Table 4 Models proposed to describe coke formation Model 1 h =1 n =1 m =0 Model 2 h =2 n =0 m =0 Model 3 h =2 n =1 m =2 dC = k1 (Cmax Cm )2 + k2 Cm dt k1 t k2 2 C = Cmax ln(1 + k1 Cmax t) + k2 Cmax t 1 + Cmax k1 t k1 dC = k1 (Cmax Cm )2 + k2 dt k1 t 2 C = Cmax + k2 t 1 + Cmax k1 t dC = k1 (Cmax Cm ) + k2 Cm dt k 1 k2 C = Cmax [1 ek1 t ] + k2 Cmax t k1

shows the best MSC value (Table 3), has been chosen for the third step of the tting. The tting results show that the activation energy for the monolayer formation is smaller than that for multilayer, which agrees with the fact that monolayer growth over the catalyst surface is catalyzed. In agreement with the postulated model, most of the monolayer coke is deposited in the rst reaction instants (Fig. 4). Simultaneously, multilayer coke is deposited over the existing coke, the rate depending on the monolayer coke concentration. The physical meaning of the model is also coherent with the experimental observations: it forecasts a constant growth of coke from the moment when all the monolayer coke has been deposited. It should be noted that experiments at different temperatures can be tted with the same value of Cmax , which means that nal covering of the catalyst surface with monolayer coke is the same for each temperature. This model does not consider the effect of the reaction atmosphere in the coke formation. As will be shown later, the prediction of the model are satisfactory, in spite of this simplication. It may be expected that the model prediction will fail if a large amount of hydrogen is feed to the reactor, as is usually done in several industrial processes. Under such operating conditions the coke formation will be decreased. Therefore, the above equation are only useful when no hydrogen is fed with the propane. 4.3. Deactivation model After having chosen the best kinetic mechanism, the next objective was to nd the relationship between the coke content in the catalyst and the propane dehydrogenation activity. According to the mechanistic model, only monolayer coke would promote deactivation. This involves that after all the monolayer coke was formed, the catalyst had a remaining activity. However, experimental activitytime results prove that activity decreases with time, even after the monolayer has been formed. This suggests that multilayer coke also deactivates. The existence of catalytic activity even after all the surface has been covered by coke can be explained by some catalytic activity of the coke itself, as has been found by several authors [31]. Two deactivation equations have been tested, a potential equation and an exponential equation.

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o

113

Potential equation : a = (1 1 Cm 2 CM )m (5) Exponential equation : a = exp(1 Cm 2 CM ) (6) where 1 , 2 and m are three constants independent of the temperature. The exponential equation shows clearly the best MSC value (Table 4) and has been chosen. Fig. 6 shows the nal complete kinetics chosen and Table 5 shows the values of the parameters obtained for this model and the statistics of the tting. All the experiments can be tted with the same values for 1 and 2 which means that the effect of coke over deactivation is independent of temperature; multilayer coke also causes deactivation but of a smaller order than monolayer coke (1 > 2 ). In Figs. 1

Table 5 Kinetic parameter values for the kinetic model chosen (Fig. 6) Parameter k01 K02 Ea1 H 1 2 k02 k03 Ea2 Ea3 Cmax k04 Ea4 Value (units) 0.0516 (mmol/(g s)) 3450 (mmol/l) 35.5 (kJ/mol) 595 (kJ/mol) 813 (g catalyst/g coke) 289 (g catalyst/g coke) 0.00242 (s1 ) 0.000357 (s1 ) 221 (kJ/mol) 325.8 (kJ/mol) 0.000682 (mg coke/mg catalyst) 10 E-5 (mmol/(g s)) 308 (kJ/mol) S.D. 0.0006 1177 13.8 23.4 101 36 4.9 E-5 4.5 E-6 1.12 0.53 7.4 E-6 1.6 E-5 13.8

Fig. 6. Proposed kinetic scheme.

114

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o

and 4, the model predictions have been plotted (product molar ow and coke content in catalyst, respectively, as a function of time) and the good agreement with the experimental data can be veried.

5. Model checking The model resulting from the t of the data obtained under different conditions should have predictive capabilities. Three experiments were designed with the following aims: To check the model applicability at a different temperature and in a different reaction system to those employed in obtaining the kinetic data. To check if the model predicts how much activity is recovered after a partial catalyst regeneration. A third experiment was done, to obtain additional proofs of the strong propene adsorption. All the experiments were carried out in the uidized bed reactor in order to have an isothermal system. In the rst experiment, propane dehydrogenation was performed at 600, 25 C higher than the maximum temperature employed to obtain the kinetic model. After reaction during 30 min, the coke was

oxidized with O2 while the exit gases were analyzed (Fig. 7) with a mass spectrometer until no COx were detected. The total coke formed from this experiment was 0.029 g coke/g catalyst, while the model prediction was 0.0306 g coke/g catalyst. Thus, the prediction error was approximately 5%. This conrms the reliability of the model to predict the coke content, even in a different reaction system from that employed to obtain the kinetic data. In the second experiment, propane dehydrogenation was carried out at 575 C during 60 min. Coke content in the catalyst was calculated using the kinetic model. Then, the catalyst was partially regenerated, using an amount of O2 smaller than the stoichiometric to consume all the coke. The reaction time corresponding to this remaining amount of coke was calculated and the reaction was performed again with this partially regenerated catalyst. The conversion and yield to propene as a function of time coincide in both the experiments (Fig. 8), taking into account the time displacement calculated previously. The third experiment aimed to experimentally check the propene adsorption. To this end, the reaction was performed in a uidized bed reactor (T = 550 C, total ow = 350 cm3 (STP)/min), and the gas feed was suddenly changed to pure argon. The concentration

Fig. 7. Coke combustion in the uidized bed reactor: product distribution in the exit stream as a function of time.

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o

115

Fig. 8. Yield to propene and propane conversion before and after partial catalyst regeneration. W/F = 12.5 g h/mol.

of propane and propene at the reactor exit was followed by mass spectrometry. Fig. 9 shows the normalized concentration values (mass spectrometer signal divided by the signal in the reaction conditions) of

both the compounds. It may be seen that, whereas, propane is quickly removed from the reactor, the signal corresponding to propene decreases much more slowly, which suggests that propene is being strongly

Fig. 9. Variation of the propane and propene concentrations, relative to the values in steady state after a sudden change to pure argon feed.

116

J. Gasc n et al. / Applied Catalysis A: General 248 (2003) 105116 o [3] H.H. Kung, Adv. Catal. 40 (1994) 1. [4] E.A. Mamedov, V. Corts-Corbern, Appl. Catal. 127 (1995) 1. [5] F. Cavani, F. Trir, in: J.J. Spivey (Ed.), Specialist Periodical Reports: Catalysis, vol. 11, Royal Society of Chemistry, London, 1994, p. 246. [6] T. Blasco, J.M. Lpez-Nieto, Appl. Catal. 157 (1997) 117. [7] P. Amigues, Y. Chauvin, D. Commereuc, J. Mol. Catal. 65 (1991) 39. [8] M. van Sint, J.A.M. Kuipers, W.P.M. van Swaij, Catal. Today 66 (2001) 427. [9] I. Suzuki, Y. Kaneko, J. Catal. 47 (1977) 239. [10] F.E. Frey, H.F. Huppke, Ind. Eng. Chem. 25 (1933) 54. [11] F.E. Frey, H.F. Huppke, US Pat. 2098959 (1937). [12] S. De Rossi, G. Ferraris, S. Fremiotti, J. Catal. 148 (1994) 36. [13] F. Ashmawy, C. McHuliffe, J. Chem. Soc., Faraday Trans. 80 (1984) 221. [14] S.D. Jackson, J. Grenfell, I.M. Matheson, G. Webb, Stud. Surf. Sci. Catal. 122 (1999) 149. [15] N.E. Fouad, S.A. Halawy, M.A. Mohamed, M.I. Zaki, Thermochim. Acta 329 (1999) 23. [16] M.I. Zaki, M.A. Hasan, N.E. Fouad, Appl. Catal. Part A: Gen. 171 (1998) 315. [17] A. Cimino, D. Cordischi, S. De Rossi, J. Catal. 127 (1991) 744. [18] M. Larsson, N. Henriksson, B. Andersson, Appl. Catal. Part A: Gen. 166 (1998) 9. [19] C. Callejas, J. Soler, J. Herguido, M. Menndez, J. Santamara, Stud. Surf. Sci. Catal. 130 (2000) 2717. [20] J.A. Pea, A. Monzn, J. Santamara, Appl. Catal. Part A: Gen. 101 (1993) 185. [21] J.A. Pea, A. Monzn, J. Santamara, J. Catal. 142 (1993) 59. [22] M. Akaike, Math. Sci. 14 (1976) 5. [23] F.J. Dumez, G.F. Froment, Ind. Eng. Chem. Process. Des. Dev. 15 (1976) 291. [24] A. Voorhies, Ind. Eng. Chem. 4 (1945) 318. [25] B.Z. Wan, H.M. Chu, J. Chem. Soc., Faraday Trans. 88 (1992) 2943. [26] E. Romero, J.C. Rodrguez, J.A. Pea, A. Monzn, Can. J. Chem. Eng. 74 (1996) 1034. [27] J.C. Rodrguez, C. Guimon, A. Borgna, A. Monzn, J. Catal. 171 (1997) 268. [28] I.S. Nam, J.R. Kittrell, Ind. Eng. Chem. Process. Des. Dev. 23 (1984) 237. [29] J. Corella, A. Monzn, Ann. Quim. 84 (1988) 205. [30] A. Monzn, E. Romeo, C. Royo, R. Trujillano, F.M. Labajos, V. Ribes, Appl. Catal. Part A: Gen. 185 (1999) 53. [31] G. Webb, Catal. Today 7 (1990) 139.

adsorbed on the catalyst. This result provides an independent check of the reaction mechanism deduced from the kinetic analysis.

6. Conclusions A kinetic model for the process of propane dehydrogenation over a Cr2 O3 /Al2 O3 catalyst has been developed. Extensive experimental work has been performed to obtain the kinetic parameters involved in the reaction, in the coke formation and in the activitycoke content relationship. A LangmuirHinshelwood model with strong adsorption of propene has been chosen as it provides the best t to describe the reaction kinetics. A monolayermultilayer coke growth model has been proposed to describe coke content in the catalyst as a function of time. The relation between coke content in the catalyst and catalyst activity has also been studied and determined with the activity being related with both monolayer and multilayer coke. This complete model explains well the qualitative and quantitative observations presented in this work. Also, employing the kinetic parameters obtained gives a satisfactory prediction of coke formation and catalyst regeneration in additional experiments. Independent experiments conrm the strong propene adsorption predicted by the kinetic model.

Acknowledgements The authors thank DGI (Spain) for nancial support in the project PPQ-2001-2519-CO2-01. References
[1] J. Cosyns, J. Chodorge, D. Commereuc, B. Torck, Hydrocarbon Process (March 1998). [2] P. Eisele, R. Killpack, Ullmanns Encyclopedia of Industrial Chemistry, 6th ed., Villey-CCH, Weinheim, 1998.

Potrebbero piacerti anche