Anda di halaman 1dari 12

Fuel 111 (2013) 449460

Contents lists available at SciVerse ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

1-Butene oligomerization over ZSM-5 zeolite: Part 1 Effect of reaction conditions


A. Coelho a, G. Caeiro b, M.A.N.D.A. Lemos a,, F. Lemos a, F. Rama Ribeiro a
a b

IBB Institute for Biotechnology and Bioengineering, Centre for Biological and Chemical Engineering, Instituto Superior Tcnico, UTL, Av. Rovisco Pais, 1049-001 Lisboa, Portugal Galp Energia, SGPS, S.A., Rua Toms da Fonseca, 1600-209 Lisboa, Portugal

h i g h l i g h t s
 Systematic study of the conversion of

g r a p h i c a l a b s t r a c t

1-butene into fuels by oligomerization over HZSM-5 catalysts.  Optimum conditions for selectivity for C 8 hydrocarbons (86 wt.%) was obtained with H-ZSM-5, at 200 C and 0.5 bar.  HZSM-5 is a suitable catalyst and allows almost steady-state operation.  Deactivation, with consequent decrease in acidity, improves the performance of the catalyst under some conditions.  Low temperature performance of the catalyst seems to be hindered by the volatility of the products.

a r t i c l e

i n f o

a b s t r a c t
The demand for middle distillates (kerosene and diesel) in comparison to gasoline fractions is increasing constantly, particularly in the majority of European countries. Thus, maximizing the production of middle distillates in the rening process is of immediate interest to reners. However, the production of liquid fuels, for instances using a catalytic cracker, always generates a signicant amount of lighter fractions, in particular of olenic nature. Thus, the use of oligomerization reactions to convert the lighter olen cuts into middle distillates to incorporate in the diesel pool is a promising process for the production of clean diesel fractions. In this work, 1-butene oligomerization over H-ZSM-5 zeolite has been investigated in a differential reactor operating at ambient pressure. The effect of the reaction conditions, such as reaction temperature, contact time and partial pressure was studied on the activity, selectivity and stability of the catalyst. The results show that an increase in the reaction temperature and/or partial pressure and in the contact time produces an improved catalyst activity. The data also show that high partial pressure improves the selectivity to C 8 . Moreover, when increasing the temperature from 150 C to 200 C the C 8 hydrocarbons selectivity increases, whereas above this temperature it decreases as expected, due to the competition of cracking reactions. Furthermore, a decrease in contact time between the reaction mixture and the acid sites of the catalyst caused the C 8 hydrocarbons fraction in the product to increase. The highest selectivity towards C 8 hydrocarbons (86 wt.%) was obtained with H-ZSM-5, at 200 C and 50 kPa of partial pressure and for the lowest value of contact time analyzed (12.5 103 h). In this way, the operating conditions must be tuned in order to achieve a signicant conversion and selectivity in desired fraction (C 8 ). 2013 Elsevier Ltd. All rights reserved.

Article history: Received 7 June 2012 Received in revised form 24 March 2013 Accepted 25 March 2013 Available online 9 April 2013 Keywords: 1-Butene oligomerization H-ZSM-5 zeolite Catalyst deactivation Heterogeneous catalysis Light olens

Corresponding author. Tel.: +351 21 841 78 90; fax: +351 21 841 91 98.
E-mail address: mandal@ist.utl.pt (M.A.N.D.A. Lemos). 0016-2361/$ - see front matter 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.fuel.2013.03.066

450

A. Coelho et al. / Fuel 111 (2013) 449460

1. Introduction Olen reactivity over acid catalysts (mainly zeolites) has been extensively studied, in order to understand the mechanism of catalytic cracking which has been explained in terms of a carbenium ion mechanism [19]. However, the oligomerization reaction, taken as the reverse of the cracking, is also a relevant process and it has also been receiving considerable attention by researchers [1016] since it provides a way to convert surplus lighter olens (such as propene and butene fractions which are generally obtained from the FCC-Fluid Catalytic Cracking efuent) into heavier hydrocarbons which are useful as sulfur-free fuels, mainly gasoline (C5AC10) [13,14]. At present, from an industrial viewpoint, the most widely used process for producing gasoline from oligomerization of light olens is the Catpoly process developed by Universal Oil Products Company (UOP) in the 1930s, which uses the phosphoric acid supported over Kieselguhr silica clay (SPA) [1719] Despite, their extensive commercial application, this catalyst has several disadvantages, namely the catalyst life time is relatively short, there is also environmental concern associated with its disposal and corrosion problems, and there is no possibility of reusing the spent catalyst. Furthermore, the products obtained are nearly exclusively in the gasoline fractions [17,20]. In this way, the search for a possible replacement of phosphoric acid with other acid catalysts, easier to use and more environmentally friendly, led to the emergence of many different processes. A lot of development has been made in the eld of homogeneous catalysts [21] and, in the past decades, a signicant research effort has been also directed towards the development of heterogeneous catalysts [19,22,23]. From a practical point of view, the latter option should always be the rst choice both because of the ease of the catalyst separation and due to their lower environmental impact [19,24]. Among the various processes developed are the Selectopol and Polynaphta processes (from IFP/Axens) which are based in silica alumina amorphous [25,26], Shells SGPK process based in the Ni-mordenite catalyst [2729] and one of the most referenced processes, which dates from the 1980s, MOGD process-Mobil olen to gasoline and distillate developed by Mobil Research and Development Corporation [3033]. This latter process uses ZSM-5 zeolite, to convert light olens (C3AC4) from Fluid Catalytic Cracking (FCC) into higher molecular weight products such as gasoline or diesel fuels. This process assumes special importance in the present since most of other processes were often oriented towards the production of gasoline [34] or to produce high-octane gasoline blending components [29]. However, as mentioned earlier, recently, the fuel market is shifting towards an increase in diesel consumption in detriment of gasoline, especially in the majority of European countries [35]. At the same time, environmental legislation has led to more strict specications on petroleum derivates [36]. In this way, the rening industry needs to adjust their production scheme in order to meet both the growth of the overall market and the shifting of the demand as efciently as possible. ZSM-5 is a promising candidate for diesel production from olen oligomerization and more studies using this zeolite seemed to be appropriate. ZSM-5 zeolite (MFI topology) is a well-known crystalline microporous aluminosilicate which possess a tree-dimensional structure with both straight and sinusoidal pores and it may have strong acidity with acid sites with different strengths [37]. The shape selectivity properties of the ZSM-5 favor the formation of linear hydrocarbons [31,37,38]; this is particularly suitable for the production of good quality distillate fuels (diesel) [39]. Moreover, this zeolite exhibits slower deactivation rates than other zeolites with

similar acidity [9,10,40]. For these reasons, ZSM-5 zeolite has been widely used in a variety of industrial processes. For example, ZSM5 is used as an additive in the FCC catalyst to improve octane number of FCC naphtha [1,27,41], it is also used in the disproportionation of toluene, xylenes isomerization, alkylation [42] and aromatization reactions [8,9]. It has also been found to be particularly suitable for converting methanol to gasoline (MTG) and olens (MTO) [79]. It is widely accepted that the oligomerization reaction of olens over acid catalysts occurs via a carbocation mechanism [10,11,38,43] and its activity and selectivity strongly depend on the operating conditions, through thermodynamics, kinetic, diffusional limitations and shape-selectivity effects [12,38]. In this way, an understanding of the fundamental aspects of this reaction is still needed. With this in mind, in the present work, we investigated the effect of operating conditions, such as reaction temperature, partial pressure and contact time between the reaction mixture and catalyst acid sites on the activity, selectivity and product distribution as well as on the stability of H-ZSM-5 zeolite catalyst. This work has been focused on a range of temperatures which favors oligomerization, which is lower than the one used in other studies on the transformation of light olens over ZSM5, typically in a range of temperatures more related to catalytic cracking [44]. 2. Experimental 2.1. Materials Prior to use, the zeolite NH4-ZSM-5 (Zeolyst, Si/Al = 15, CBV 3024E) was calcined in a tubular reactor at 500 C for 8 h under a 1 1 ow of dry air of 0.5 Ldry air g , in order to remove any organic zeo h compounds that could be adsorbed on the zeolite and to obtain its protonic form (H-form). After calcination, the zeolite was always maintained in a container with a constant and high humidity in order to ensure that the surface of the zeolite is fully saturated, protecting it from adsorbing any contaminant, and at the same time, to keep the amount of water adsorbed on the catalyst constant. 1-Butene and the other gases used were high-purity products: 1-butene (AirLiquide, 99%) nitrogen (AirLiquide, 99+%), air (AirLiquide, 99+%), hydrogen (AirLiquide, 99+%) and ammonia (Linde, 99.5%) and were dried by molecular sieves before use. 2.2. Catalytic test Prior to each experiment, the catalyst was pre-treated under a ow of dry nitrogen of 60 mL min1, to clean the surface of the catalyst, by heating it at 10 C min1 up to the temperature of 450 C, where it was keep for 8 h. The reactor was then allowed to cool down to the reaction temperature under the ow of N2. The catalytic tests were performed using 1-butene as reactant, diluted with dry nitrogen in a xed bed continuous ow reactor, in the gas phase, at atmospheric pressure and within a temperature range from 150 C to 250 C. The partial pressure of the reactant was modied, from 12.5 kPa to 50 kPa, by changing the nitrogen ow for the same olen ow (25 mL min1), which results in the same contact time between the olen and the catalyst. The amount of H-ZSM-5 catalyst used in these tests was always 50 mg. However, at 200 C and with a partial pressure of 50 kPa some tests with 150 mg and 300 mg of the catalyst were also carried out, keeping the ow rates constant, in order to study the effect of contact time. The reaction conditions used in this work are given in Table 1. To avoid the deactivation of the catalyst that would occur during the chromatographic analysis, the reaction was interrupted by

A. Coelho et al. / Fuel 111 (2013) 449460 Table 1 Operating conditions for the catalytic transformation of 1-butene. Reaction temperature (C) 150 200 225 250 200 Partial pressure (kPa) 12.5 25 50 50 Total ow (nitrogen plus olen) (mL min1) 200 100 50 50 Catalyst mass (mg) 50 Contact timea 1/ WHSV (h) 12.5 103

451

50 150 300

12.5 103 37.1 103 74.2 103

2.4.3. Nitrogen adsorption Nitrogen adsorptiondesorption isotherms at 196 C of the calcined catalysts were obtained using a Micromeritics ASAP 2010 apparatus to determine the micropore volume and external surface area. The sample was previously degassed under vacuum at 150 C for 24 h. The micropore volume (Vmicro) and the external surface area (Sext) were calculated using the t-plot method, with the HarkinsJura equation for determining the thickness of the adsorbed layer. The nitrogen adsorption isotherm at 196 C on the fresh H-ZSM-5 sample is presented in Fig. 1a. 2.4.4. Ammonia temperature programed desorption (NH3-TPD) The acid properties of the catalyst were obtained by ammonia temperature desorption (TPD). Previously, the zeolite was placed in a quartz reactor and it was pre-treated, during 8 h at 450 C under a ow 60 mL min1 of dry nitrogen. After cooling to 90 C, ammonia adsorption was performed by injection of the successive pulses of ammonia, until the catalyst surface was saturated with ammonia. The amount of ammonia in the efuent He stream was measured by a thermal conductivity detector (TCD). The desorption experiments of the ammonia saturated zeolites, were carried out in the TA Instruments SDT 2960 simultaneous DSCTGA apparatus using an alumina pan. The sample was heated up to 700 C with a heating rate of 10 C min1, and about 23 mg of each saturated catalyst were used. Prior to this desorption step, the samples were outgassed under N2 ow (80 mL min1) by heating at a rate of 10 C min1 up to 150 C, and this temperature was maintained for 30 min to remove physisorbed water and NH3. Blank experiments, where no ammonia was adsorbed, were also performed to account for baseline drifts. The TGDSC apparatus was calibrated, according to the manufacturers specications, in relation to weight, temperature and DSC calibrations. The acid strength distribution was obtained from the NH3-TPD experiments by numerical deconvolution, as described elsewhere [4749]. 2.5. Analyses of the catalysts after reaction The amount of coke, formed after 21 min of reaction, was determined by TG/DSC analyses in a TA Instruments SDT 2960 simultaneous DSCTGA apparatus, under an air ow of 75 mL min1; about 2022 mg of coked catalyst were used. The sample is heated up to 210 C at 10 C min1; this temperature is maintained for 30 min to remove any water that might be adsorbed in the zeolite. After this pre-treatment the burning of the coke is performed using the same air ow, by increasing the temperature, at 10 C min1, up to 800 C; this temperature was maintained for an additional 25 min, before cooling down to the room temperature. Therefore, the amount of coke was estimated according to the following formula:

a Contact time values were determined with the mass of dry catalyst. (WHSV = weight hourly space velocity).

cutting off the olen feed. During these periods, the sample was kept only under a ow of pure nitrogen at the working temperature. Thus, the experiment consists of a sequence of feeding cycles of the reactant, with 3 min each at the end of which the outlet stream is sampled into the gas chromatograph. 2.3. Analysis of the products All products were analyzed on-line with a Shimadzu GC-9A gas chromatograph equipped with a PLOT (KCl/Al2Cl3) capillary column (50 m) and a Flame Ionization Detector (FID). The chromatogram was integrated with a Shimadzu C-R3A integrator. The products were lumped into several fractions based on their respective GC retention times, identied by the injection of pure hydrocarbons calibration mixtures. The activity, which measures the rate of consumption of the reagent consumed per unit mass of dry catalyst, was dened accordingly to the following equation:

Activity

F 0 xfinal W

}where F0 is the molar ow rate of reagent (mol min1), xnal the conversion at the exit stream and W is the mass of dry catalyst (g). Note that all the thermodynamic equilibrated butene isomers were considered as reactants [45,46]. 2.4. Catalyst physicochemical characterization 2.4.1. Elemental analyses The concentration of silicon and aluminium in the zeolite was obtained by elemental analyses. ICP (Inductively Coupled Plasma) was used to determine aluminium, while silicon was measured by atomic emission spectrometry. The results of elemental analysis are presented in Table 2. 2.4.2. Scanning electron microscopy Scanning electron microscopy (SEM) was performed to evaluate crystallite size and morphology of the H-ZSM-5 used. This technique was carried out in a JEOL JSM-7001F Field Emission Gun (FEG-SEM).

Coke amount wt:%

mi mf 100 mi

Table 2 Physico-chemical properties of H-ZSM-5 zeolite. Al (%w/w)a 2.2 (0.81)


a b c d

Si (%w/w)a 39

(Si/Al)molara 17.0

Total acidity (mmol NH3 g1)b 1.02

Vmicro (cm3 g1)c 0.186

Sext (m2 g1)d 50

From elemental chemical analysis. Value in mmol Al g1 given in parenthesis. From ammonia TPD experiments. Vmicro, micropore volume, determined by the t-plot method. Sext, external surface area, determined applying the t-plot method.

452

A. Coelho et al. / Fuel 111 (2013) 449460

(a)
Vads (cm3/g) (STP)

200 160

(b)
dq/dt(mmol.s -1)

3.0E-05 2.5E-05 2.0E-05 1.5E-05 1.0E-05

(I)

120 80 40 0 0 0.2 0.4 0.6 0.8 1

(II)

(IV)
5.0E-06 0.0E+00 150

(III)

250

350

450

550

650

750

Relative pressure (P/P0)

T(C)

Fig. 1. H-ZSM-5 zeolite nitrogen adsorption isotherm at 196 C (a), NH3-TPD curves for the H-ZSM-5 zeolite and (b) (I simulated curve; II experimental curve; IV decomposition curves; and III blank curve).

where mi represents the weight of coked catalyst after desorption of water and mf represents the weight of coked catalyst after burningoff the coke. The catalyst deactivation was computed as:

Catalyst deactivation

ai af 100 ai

obtained for this zeolite (Table 2). This value is in good accordance with the expected value based on aluminium content using the Si/ 1 Al ratio from the supplier (1.04 mmol g zeo ) although it is a little higher than the value obtained from elemental analysis reported 1 in Table 2 (0.81 mmol g zeo ), in line with the difference already discussed on the Si/Al ratio obtained by elemental analysis. 3.2. Catalytic transformation of 1-butene We will now present and discuss the main results obtained in the transformation of 1-butene over the H-ZSM-5 catalyst used in this work. The oligomerization reaction is a relatively straightforward one: the reaction is started by the dimerization of two olen molecules, forming a dimer which can then react with further monomers, to produce higher oligomers, or crack, producing lighter products. The global activity is mainly related to the ability of the catalyst to produce the rst dimers; this reaction is bimolecular, so it is expected to increase with pressure, and exothermic, so the equilibrium displacement towards the products is expected to be favored by lower temperatures; in contrast the secondary reaction, cracking, is the reverse of the oligomerization reaction and will be, in principle, favored by higher temperatures. The main difculty in this reaction is, thus, to be able to perform the reaction at a rate as high as possible, which would imply using higher temperatures, while keeping the selectivity towards the oligomers high, which will be disfavored by increasing temperatures [16]. It is also expected that some deactivation occurs, in particular by deposition of what is known as coke, on the surface of the catalyst. This deactivation, however, is expected to be very limited since the catalyst has relatively narrow pores that prevent coke formation. 3.2.1. Effect of time-on-stream and contact time Fig. 2a presents some examples of the time-course evolution of the reaction. As expected the catalyst suffers some deactivation and the conversion decreases rapidly in the beginning and tends to stabilize. Also, as it would be expected, the conversion increases (although the activity decreases somewhat see Fig. 3) with increasing contact time. Fig. 2b shows the evolution of C 8 selectivity with time-onstream (t-o-s) and contact time. Again, as expected, the selectivity tends to decrease with an increase in contact time due to the extension of secondary reactions. Selectivity, however, shows a relatively lower sensitivity towards time-on stream it tends to be relatively constant except for the lowest contact time where it

where ai is the initial activity determined for 3 min and af is the activity determined after 21 min of reaction. 2.6. Testing for mass transport limitations The existence of external mass transfer limitations was investigated experimentally for the catalyst H-ZSM-5 at 200 C by measuring the conversion of 1-butene as function of the reactant ow, while maintaining the contact time between reactant and catalyst constant, by increasing the catalyst mass. Three ow rates (25, 50 and 100 mL min1) were tested with different amount of catalysts (50, 100 and 200 mg), respectively. Since the conversion values for all experiments were rather similar (8.0%, 8.6% and 7.5%), indicating that the mass transport did not inuence the reaction rate, it was concluded that there are no external diffusion limitations. Internal limitations were checked, using the WeiszPrater criterion [50] and, as it has been reported elsewhere [49], there are no signicant diffusion limitations. 3. Results and discussion 3.1. Physico-chemical characteristics of H-ZSM-5 zeolite The chemical composition of the calcined material presented a Si/Al ratio of 17. The small difference, observed between this value and the value provided by the manufacturer (Si/Al = 15) can be ascribed to the differences in the analytical methods. The values in Table 2 show that the H-ZSM-5 zeolite presents an external surface area of 50 m2 g1, and a micropore volume of 0.186 cm3 g1 (see also Fig. 1a) both values are in agreement to those reported in literature [51]. Scanning electron microscopy (SEM) showed that the H-ZSM-5 presented crystallites of ca. 2 lm. The acidity of the H-ZSM-5 zeolite was determined by NH3-TPD, as shown in Fig. 1b) and as expected it has acid sites with different strengths. Two desorption peaks are present, the low-temperature peak at 150300 C corresponds to relatively weak acid sites, while the high-temperature peak at 350480 C corresponds to stronger 1 acid sites. A total amount of acid sites of 1.02 mmol g zeo was

A. Coelho et al. / Fuel 111 (2013) 449460

453

(a) 18
1-Butene conversion (%)
16 14 12 10 8 6 4 2 0
0 3 6 9

-3 1/1/WHSV=74.2 WHSV=0.0742 h x 10 h -3 1/1/WHSV=37.1 WHSV=0.0370 h h x 10

(b) 100
90

-3

Selectivity C 8+ (wt. %)

1/1/WHSV=12.5 WHSV=0.0125 h h x 10

80 70 60 50 40 30 20 10 0
-3 1/1/WHSV=74.2 WHSV=0.0742 h x 10 h -3 h x 10 1/1/WHSV=37.1 WHSV=0.0370 h -3 x 10 1/1/WHSV=12.5 WHSV=0.0125 h h

12

15

18

21

12

15

18

21

Time-on-stream (min)

Time-on-stream (min)

Fig. 2. Effect of time-on-stream on the conversion of 1-butene (a) and selectivity to C 8 products (b) for 1-butene transformation over H-ZSM-5 zeolite for different contact times at 200 C and 50 kPa of partial pressure.

increases slightly with time-on stream. This increase is probably due to the fact that deactivation will occur rst on the more acidic sites which are the ones that most favor cracking reactions [44,52 55]. The changes in selectivity are also depicted in Fig. 3b were it can also be seen that the extent of oligomerization, as measured by the average number of carbon atoms in the products, steadily decreases with increasing contact time (from 8.24 to 7.34 and 6.98 as contact time increases from 12.5 103 h to 37.1 103 h and 74.2 103 h). In Fig. 4a, a further detail on product distribution shows that the increase in contact time also decreases sharply the formation of trimers, clearly favoring cracking products, Scracking (C3AC7) increases from 12 wt.% to 32 wt.% when increasing the contact time from 12.5 103 h to 74.2 103 h. The aromatics selectivity (mainly benzene, xylenes, ethylbenzene, toluene and C9 aromatics) was always less than 10 wt.%. We will now look at the effect of contact time on deactivation; deactivation is mainly due to the formation of coke, a general denomination to the non-desorbed products that prevent access to the acid sites decreasing the number and average acidity of the accessible sites [56,57]. In Table 3 we can see that, although coking is usually associated with secondary reactions, coke deposition decreases with increasing contact time. This may indicate that the deactivation is associated with the formation of the primary products, which also decrease with increasing contact time

This interpretation is also consistent with the results obtained for the coke deposited on the catalysts, which are shown in Fig. 4b. We can observe that, apart from a water desorption peak (not shown) that occurs around 100 C, the TG analysis of the hydrocarbons deposited on the catalyst shows two main peaks, one around 300 C and another one around 500 C. The rst one is assigned mostly to the desorption of heavy products (reversible coke), probably in the range from C12 to C20 in view of the evaporation temperature, and is associated with an endothermic DSC signal, while the second one, which can be attributed to the burning of heavier molecules, is associated with an exothermic DSC signal. From these results we can see that, although the amount of heavier coke is mostly unaffected by the increase in contact time, there is a clear decrease in the rst one. This was further veried by the observation that, when performing the reaction at lower temperatures, if the used catalyst was washed with a light solvent (such as hexane) at the end of the reaction a signicant amount of oligomers that were retained inside the porous system were recovered. 3.2.2. Effect of reaction temperature and partial pressure 1-Butene conversions and initial activities obtained as a function of reaction temperatures and partial pressures are depicted in Fig. 5, accordingly to the operating conditions listed in Table 1. The initial activities were determined using the conversion values obtained for the t-o-s = 3 min.

(a)

18 16 14

0.0020 0.0018 0.0016

(b) 100
95 90

8.4 8.2 8.0

Activity (mol.min -1.g -1)

Conversion (%)

12 10

Selectivity C 8+ (wt %)

0.0014 0.0012 0.0010

85 80 75 70 65 7.2 60 55 50 0.00 0.02 0.04 0.06 7.0 6.8 0.08 7.8 7.6 7.4

8 6 4 2 0 0.00 0.02 0.04 0.06

0.0008 0.0006 0.0004 0.0002 0.0000 0.08

1/ WHSV (h)

1/ WHSV (h)

Fig. 3. Effect of contact time (reciprocal of WHSV) for the initial (t-o-s = 3 min) on (a) conversion of 1-butene (open symbols) and catalyst activity (closed symbols) (b) and selectivity to C 8 (open symbols), and average carbon number (closed symbols) at 200 C and for 50 kPa of partial pressure.

Average Carbon Number

454

A. Coelho et al. / Fuel 111 (2013) 449460

(a)
Product selectivities (wt %)

100 90 80 70 60 50 40 30 20 10 0

-3 1/ WHSV=0.0125 1/WHSV=12.5 x 10h h -3 h 1/WHSV=37.1 x 10h 1/ WHSV=0.0370

(b)

0.18 0.16 0.14

12.5 x 10-3 h 37.1 x10-3 h 74.2x10-3 h

200 C

1/WHSV=74.2 x 10h h 1/ WHSV=0.0741

-3

DTG (mg min-1)


er s C9 -C 11 A ro m at ic s g ki n er s

0.12 0.10 0.08 0.06 0.04 0.02 0.00 200

C8 +

cr ac

di m

t ri

400

600

800

T (C)

Fig. 4. Effect of contact time on (a) the selectivities in cracking (C3AC7), trimers (C12), C9AC11 products and dimers (C8) and on (b) coke deposited on the catalyst for 1-butene transformation at atmospheric pressure over H-ZSM-5 zeolite at 200 C and 50 kPa.

Table 3 Coke content obtained from TG/DSC experiments and deactivation percentage, for 1butene conversion over H-ZSM-5 at 200 C and 50 kPa for three different contact times. Contact time (1/WHSV) (h) 12.5 103 37.1 103 74.2 103 Deactivation (wt.%) 52.48 40.80 38.81 Coke (wt.%) 7.83 7.58 6.70

The results clearly show that 1-butene conversion is dependent on the reaction temperature and partial pressure of the olen. As expected, by thermodynamic and kinetic considerations, 1-butene conversion increases when temperature and/or partial pressure increases. Note that the rst step of this transformation is the oligomerization reaction, a second order exothermic reaction which consequently is thermodynamically favored at low temperature. Nevertheless, as temperature increases the reaction rate also increases, although the cracking reaction rate has a larger increase [37]. The inuence of the time-on-stream (t-o-s) can also be seen in Fig. 5b and as seen above, the conversion decreases gradually with reaction time. The results of the coke content deposited on the catalysts and deactivation %, at the end of each 1-butene experiment,

after 21 min of time-on-stream, are shown in Fig. 5. It is interesting to note that, in all cases, the deactivation occurs from early values of time-on-stream (t-o-s) and that the catalyst usually tends to a quasi-steady-state activity after a period of initial deactivation, which can be useful for a practical point of view. It was also observed, that the amount of the coke deposited on the zeolites has a general tendency to increase with partial pressure (see Fig. 6a), in line with the changes in activity. This is also probably due to the production of higher molecular weight oligomers inside the zeolite pores being fully consistent with the results obtained for selectivity to C 8 products. The inuence of temperature is less clear: the amount of coke deposited increases up to 225 C and then decreases when the temperature increases up to 250 C. It is likely that at lower temperatures the catalysts deactivation is mainly due to the formation of heavy oligomers that do not evaporate (which is consistent with the results presented above see Fig. 4b) and at sufciently high temperature (above 225 C), these heavy products can evaporate or be cracked into smaller molecules that can be diffuse more easily out of the catalyst pores, reducing coke deposition. However, to explain the fact that, the deactivation of the catalyst decreased or remain practically constant when there is an increase in the coke percentage (see Fig. 6), it is necessary to consider that, the deactivation of zeolite catalysts by coke is caused either

Fig. 5. Effect of partial pressure or reaction temperature on the initial activity (after 3 min of reaction) (a) and inuence of time-on-stream, for various temperatures (N 250 C, r 225 C, s 200 C and j 150 C) (b) on the 1-butene conversion over H-ZSM-5 zeolite.

A. Coelho et al. / Fuel 111 (2013) 449460

455

(a) 10
9 8 7

(b)
Deactivation (%)

100 90 80 70 60 50 40 30 20 10 0 150 175 200 225 250

Coke (wt %)

6 5 4 3 2 1 0 150 175 200 225 250

Temperature (C)

Temperature (C)

Fig. 6. Coke content obtained from TG/DSC experiments (a) and deactivation percentage (b), for 1-butene conversion over H-ZSM-5 zeolite at (N) 50 kPa, (j) 25 kPa and (r) 12.5 kPa.

by poisoning of acid sites or by pore blockage. In the rst case, one coke molecule blocks one active site, in the second case, one coke molecule blocking the access of reactants to, on average, more than one active site of the catalysts [5658]. Therefore, the deactivating effect is much more pronounced in the case of pore blockage. Both deactivation modes can occur simultaneously although one of the two mechanisms is usually predominant. The inuence of temperature and partial pressure on selectivity is depicted in Fig. 7.

The results indicate that the selectivity towards C 8 increases when the partial pressure of the reagent in the feed was increased from 12.5 kPa to 50 kPa while keeping the temperature constant. As discussed above this result can be attributed to the fact that oligomerization reaction is a second order reaction. In this way, although the oligomerization reaction can be conducted at a wide range of olen partial pressures, higher olen partial pressures are preferred since they promote the bi-molecular reactions leading to oligomerization [59,60]. This effect is more noticeable at lower temperatures.

(a)
Selectivity C 8+ (wt %)

100 90 80 70 60 50 40 30 20 10 0 150 175 200 225 250 2 0 125

(b) 14
12

Yield C 8+ (wt %)

10 8 6 4

150

175

200

225

250

Temperature (C)

Temperature (C)

(c)
Yield C 8+ (wt %)

14 12 10 8 6 4 2 0 0 5 10 15 20 25 30

1-Butene conversion (%)


Fig. 7. Effect of temperature or partial pressure ((N) 50 kPa, (j) 25 kPa and (r) 12.5 kPa) on the selectivity in C 8 (a) or yield in C8 (b) and yield in C8 vs. initial conversion (t-o-s = 3 min) (c), for the 1-butene transformation over H-ZSM-5 zeolite.

456

A. Coelho et al. / Fuel 111 (2013) 449460

Changes in the reaction temperature have a more complex effect in the selectivity in C 8 . Fig. 7a shows that the temperature rise has two distinct effects; it was found that by increasing the temperature from 150 C to 200 C the C 8 hydrocarbons in the product increased, whereas above this latter temperature they decreased. The selectivity to the oligomerized products (C 8 ) goes through a maximum of approximately 86 wt.% at 200 C and 50 kPa of partial pressure. The decrease of selectivity for higher temperatures was expected and can be explained by the fact that cracking reactions, having higher activation energy, are favored by the increasing in temperature. However, the appearance of a maximum is not an expected behavior if one takes into account the kinetic and thermodynamic aspects of the reaction. Since the C 8 products result from the initial oligomerization reaction while products with a number of carbon atoms lower than eight carbon atoms result from cracking reactions that occur on these initial oligomers, it would be expected that at low temperatures the C 8 selectivity should be higher and decreased steadily with temperature. However, the experimental observation of a low selectivity for oligomerization products at the lowest temperature studied (at 150 C), can be explained by the fact that some of the oligomerization products formed have boiling points above the operating temperature. Thus, they may be retained (through condensation/ trapping) inside the pores of the zeolites, staying longer in contact with acid sites of the catalyst given the greater difculty in leaving the pores owing to the strong adsorption [8] of these products on the acid sites. For example, C 9 hydrocarbons have

boiling point temperatures above 150 C [61]. This leads to an increase of the probability of successive cracking reactions and consequently to an increase in the cracking products (see also Fig. 8a), consistent with the observed selectivity towards cracking products when the contact time between catalyst and product mixture was increased (see discussion above, in particular in relation to Fig. 3). The results also show that the C 8 yield increases both with reaction temperature and partial pressure although when working with lower temperatures or lower butene conversions the C 8 yield is unaffected by partial pressure (see Fig. 7b and c). These results indicate that when in practical operation of a reactor a compromise between selectivity and activity has to be considered, and if the goal is to achieve a good selectivity on the oligomers production, lower activity values are certainly to be expected. A more detailed view of the inuence of temperature on the product selectivities obtained for 1-butene transformation at 50 kPa of partial pressure is shown in Figs. 8a and 9. It can be seen that, with exception of 150 C, the selectivity toward light hydrocarbons, C3AC7, (Scracking), increases with reaction temperature while, the dimers product-C8 hydrocarbons (Sdimers), trimers product-C12 hydrocarbons (Strimers) and C9AC11 products (SC9AC11) selectivities show a general trend to decrease with temperature, in accordance with the fact that the lighter hydrocarbons (<C8) are the result of the cracking of the heavier products formed by the oligomerization reaction. These results indicate that at 200 C, oligomerization of olens is the dominant reaction while

(a)
Product Selectivities ( wt %)

70

150 C

50 kPa
60

(b)
Product Selectivities (wt %)

70

200 C; 50 kPa
60 50 40 30 20 10 0

t-o-s=3 min t-o-s=9 min t-o-s=21 min

200 C 225 C

50 40 30 20 10 0

250 C

cracking

dimers

trimers

C9-C11
70

cracking

dimers

trimers

C9-C11

(c)
Product Selectivities ( wt %)

70

200 C
60 50 40 30 20 10 0

Product Selectivities ( wt %)

50 kPa 25 kPa 12.5 kPa

(d)

250 C
60 50 40 30 20 10 0

50 kPa 25 kPa 12.5 kPa

cracking

dimers

trimers

C9-C11

cracking

dimers

trimers

C9-C11

Fig. 8. Product distribution (wt.%) for 1-butene oligomerization at atmospheric pressure: effect of temperature for t-o-s = 3 min and partial pressure 50 kPa (a), effect of timeon-stream at 200 C and partial pressure 50 kPa (b) and effect of partial pressure at 200 C and 250 C (t-o-s = 3 min).

A. Coelho et al. / Fuel 111 (2013) 449460

457

at higher temperature cracking is predominant leading to a higher cracking selectivity (C3AC7 hydrocarbons). As discussed above, it would be expected that at 150 C the selectivity towards cracking products would be lower than those obtained at other temperatures studied, but this is not the case. Partial pressure, on the other hand, has a much smaller inuence on the selectivities (see Fig. 8c and d), although an increase in dimers selectivity can be observed upon increase of the partial pressure, as it would be expected. In general we can say that, as the partial pressure is increased, the selectivity towards cracking products decreases while the selectivity towards hydrocarbons with eight or more carbon atoms increases. Another interesting aspect is the change of the selectivity with time-on-stream, depicted in Fig. 8b) for 200 C and 50 kPa. While the selectivity to cracking products does not change signicantly there is an enhancement of selectivity for the dimers, accompanied by a decrease in the selectivity to trimers and C9AC11 products. In fact these observations are to be expected since secondary products are the most affected with the decrease of conversion due to catalyst deactivation. Additional information on product distribution, including parameters like olens/parafns, parafns/aromatics, oligomerization/cracking, aromatization/cracking and aromatization/oligomerization ratios, at different temperatures and partial pressures are depicted in Figs. 9 and 10. It was observed that substantial amounts of parafns are formed, resulting from hydrogen transfer reactions involving reac-

tant or product olens. Under all the reaction conditions investigated, 80 wt.% or more of the products observed in the gas phase were parafns and olens, in the range from C3 to C12. The remaining products (20 wt.% or less) were aromatics (mainly benzene, xylenes, ethylbenzene, toluene and C9 aromatics see Table 3), probably formed by dehydrocyclization of the olens. The aromatics selectivity increases with increasing temperature and partial pressure (except at 150 C), as it can be seen in Table 4. This is in-line with what is thermodynamically expected [62] and in view of the fact that hydrogen transfer involves high activation energy bimolecular reactions. The aromatization/cracking ratio, however, was found to decrease with increasing reaction temperature for the three partial pressures studied (Fig. 10a) whereas aromatization/oligomerization tends to increase with reaction temperature, except at 150 C. For a given temperature both ratios are higher at higher partial pressures. Higher aromatic selectivity observed at 150 C (see Table 4) may be attributed to the same effect that was discussed above, where the low volatility of the oligomerization products induces a larger contact time between the reaction mixture and the active sites of the catalyst. In fact the thermodynamics for the secondary reactions (hydrogen transfer, cracking and aromatization) indicates that all are favored with increasing temperature [12]. The experimental olen to parafn ratio obtained as a function of reactor temperature and olen pressure is illustrated in

(a)
Product Distribution (wt %)

35 30 25 20 15 10 5 0 1 2 3 4 5

150 C
Paraffins Olefins Aromatics

(b)
Product Distribution (wt %)

35 30 25 20 15 10 5 0

200 C

Paraffins Olefins Aromatics

10 11 12

10 11 12

Number of Carbon Atoms 225 C

Number of Carbon Atoms 250 C


Paraffins Olefins Aromatics 25 20 15 10 5 0

(c)
Product Distribution ( wt %)

35 30 25 20 15 10 5 0 1 2 3 4

Paraffins Olefins Aromatics

(d)
Product Distribution (wt %)

35 30

10 11 12

10 11 12

Number of Carbon Atoms

Number of Carbon Atoms

Fig. 9. Product distribution for the transformation of 1-butene at different reaction temperature at 50 kPa with the fresh H-ZSM-5 catalyst.

458

A. Coelho et al. / Fuel 111 (2013) 449460

Fig. 10. Effect of temperature or partial pressure ((N) 50 kPa, (j) 25 kPa and (r) 12.5 kPa) on oligomerization/cracking (open symbols), aromatization/cracking (lled symbols) and aromatization/oligomerization (grayed symbols) mass ratios (a) and for olens/parafns (b) and parafns/aromatics (c) molar ratios for the 1-butene transformation over fresh H-ZSM-5 zeolite.

Table 4 @ Olens/parafns (O/P), parafns/aromatics (P/A) C@ 5 and C3 olens and hydrogen/ carbon (H/C) molar ratios obtained for the 1-butene transformation over H-ZSM-5 zeolite at different temperatures and partial pressures of reactant. Partial pressure (kPa) Reaction temperature (C) 150 200 225 250 150 200 225 250 150 200 225 250 Aromatics (wt.%) Molar ratio O/P 5.00 1.55 2.08 3.69 2.88 1.32 1.90 2.33 1.02 0.73 1.79 2.25 P/ A 0.6 8.8 6.9 5.1 1.5 6.7 5.4 5.6 4.4 6.2 4.2 3.3
@ C@ 5 =C3

H/C 1.87 2.06 2.06 2.04 1.93 2.06 2.05 2.05 2.04 2.07 2.03 2.01

12.5

25

50

19.3 4.9 5.5 5.9 15.4 6.8 7.3 6.4 11.1 9.1 9.4 10.8

5.03 3.39 2.77 2.05 4.87 3.63 2.78 2.28 3.63 3.93 2.80 1.80

Fig. 10b). It is possible to see that, in most cases and as expected, this ratio presents values greater than one [63,64]. This ratio is an indication of the extent of secondary reactions; in fact, should none of these secondary hydrogen transfer reactions occur we would expect that the olen/parafn ratio to be innite. It is possible to see, in Fig. 9, that increasing the temperature from 200 C to 250 C clearly increases the production of C3 and C5 olens, while C8 parafns production decreases due to the cracking secondary reactions. However it can also be seen that C1 and C2 products (methane and ethane/ethylene, respectively) are

not observed, which is an indication that, as expected, the protolytic cracking mechanism is negligible for these operating conditions. The same is true for the two other partial pressures studied (results not shown). A parafn/aromatic ratio of three would be expected since for the formation of an aromatic hydrocarbon from an olen, three other olen molecules would have to be hydrogenated, producing three parafn molecules. However, in most cases, the molar parafns/aromatics ratios have values greater than 3 and in some cases with relatively high values (see Table 4 and also Fig. 10b). This fact indicates that a signicant part of the hydrogen responsible for parafn production is associated with coke formation. Also, it can be observed from Table 4 that, with few exceptions, the hydrogen/carbon ratio obtained in the products has a similar H/C ratio of the reactant (1-butene), presenting a value close to 2. The values observed for H/C ratio are consistent with the parafn to aromatic ratio discussed above and the fact that little coke is, in fact, deposited on the catalyst. @ The ratio between C@ 5 =C3 products was also calculated for all runs (see Table 4) since, apart from butane itself, these would be the most probable products of dimer cracking. The results showed that, for the same partial pressure, there is a general trend for the @ C@ 5 =C3 molar ratio to decrease with increasing temperature and always present values higher than 1, probably due to further transformation of the cracking products, which may be involve oligomerization reactions. 4. Conclusions The results obtained in this systematic study showed that olen oligomerization is a reaction for which the activity and selectivity

A. Coelho et al. / Fuel 111 (2013) 449460

459

strongly depend on the operating conditions and where secondary reactions, such as cracking, constitute a serious limitation. Temperature, reactant partial pressure and contact time between the reaction mixture and the acid sites of the catalyst have great effect in controlling activity and selectivity of the catalysts. The results showed an increase in the catalytic activity with increasing temperature, partial pressure and or contact time. However it was found that oligomerization selectivity is maximal at 200 C. At higher temperatures cracking increases rapidly, resulting in a product distribution that shifts towards lighter hydrocarbons. Although it would be reasonable to assume that at 150 C the selectivity towards oligomerization should be still higher than at 200 C, in view of thermodynamic considerations, an unexpected higher cracking selectivity was obtained which was attributed to the fact that some of the oligomerization products formed will remain condensed/trapped within catalyst due to their low volatility. It was also observed that the higher partial pressures favored the formation of heavier products while long contact time favors the cracking products. The results also show that the deactivation, which is accompanied by a lowering of the acidity of the catalyst, improved selectivity for dimer (C8) products. According to the results previously described, the best conditions, in terms of the selectivity toward oligomerization products (C 8 ), when using this HZSM-5 (Si/Al = 15) zeolite, seem to be the combination of higher partial pressures (50 kPa), lower value of contact time (12.5 103 h) and a temperature of 200 C. Furthermore, although the catalyst presents some initial deactivation, its activity stabilizes after a while and, under the right conditions, this limited deactivation is accompanied by an increase in oligomerization selectivity. Under these conditions it was possible to obtain the highest value of selectivity for C 8 (ca. 86 wt.%). As a sequence for this work the inuence of the acidity of the catalyst should be investigated since it seems clear that very acidic catalysts tend to favor cracking reactions in relation to oligomerization. It can also be suggested that the oligomerization of olens would benet from a process design that allowed the continuous removal of the oligomerization products. Acknowledgments A. Coelho wishes to thank to Fundao para a Cincia e Tecnologia for nancial support for the PhD Grant (Ref. SFRH/BD/66744/ 2009). The authors also would like to thank Jos Roque (from GALP Energia, Portugal) for fruitful discussions. References
[1] Sadeghbeigi R. Fluid catalytic cracking handbook, design, operation and troubleshooting of FCC facilities. 2nd ed. Gulf Professional Publishing; 2000. p. 1316. [2] Corma A. Inorganic solid acids and their use in acid-catalyzed hydrocarbon reactions. Chem Rev 1995;95:559614. [3] Corma A, Wojciechowski BW. Catalytic cracking: catalysts, chemistry and kinetics (chemical industries). New York: M. Dekker; 1986. [4] Buchanan JS. The chemistry of olens production by ZSM-5 addition to catalytic cracking units. Catal Today 2000;55:20712. [5] Tiong Sie S. Acid-catalyzed cracking of parafnic hydrocarbons. 1. Discussion of existing mechanisms and proposal of a new mechanism. Ind Eng Chem Res 1992;31:18819. [6] Jacobs PA. Carboniogenic activity of zeolites. Amsterdam: Elsevier; 1977. [7] Buchanan JS, Santiesteban JG, Haag WO. Mechanistic considerations in acidcatalyzed cracking of olens. J Catal 1996;158(1):27987. [8] Stocker M. Gas phase catalysis by zeolites. Micropor Mesopor Mater 2005;82:25792. [9] Caeiro G, Carvalho RH, Wang X, Lemos MANDA, Lemos F, Guisnet M, et al. Review activation of C2C4 alkanes over acid and bifunctional zeolite catalysts. J Mol Catal A: Chem 2006;255:13158. [10] Sanati M, Hornell C, Jaras S. The oligomerization of alkenes by heterogeneous catalysts. Catalysis 1999;14:23687. [11] Kissin YV. Chemical mechanisms of catalytic cracking over solid acidic catalysts: alkanes and alkenes. Catal Rev 2001;43(12):85146.

[12] Quann RJ, Green LA, Tabak SA, Krambeck FJ. Chemistry of olen oligomerization over ZSM-5 catalyst. Ind Eng Chem Res 1988;27(4):56570. [13] Peratello S, Molinari M, Bellussi G, Perego C. Olen oligomerization: thermodynamics and kinetics over a mesoporous silicaalumina. Catal Today 1999;52:2717. [14] Van Grieken R, Escola JM, Moreno J, Rodrguez R. Direct synthesis of mesoporous M-SBA-15 (M = Al, Fe, B, Cr) and application to 1-hexene oligomerization. Chem Eng J 2009;155:44250. [15] Van Grieken R, Escola JM, Moreno J, Rodriguez R. Liquid phase oligomerization of 1-hexene over different mesoporous aluminosilicates (Al-MTS, Al-MCM-41 and Al-SBA-15) and micrometer/nanometer HZSM-5 zeolites. Appl Catal A 2006;305:17688. [16] de Klerk A. Oligomerization of 1-hexene and 1-octene over solid acid catalysts. Ind Eng Chem Res 2005;44(11):388793. [17] Perego C, Peratello S. Process for producing gasolines and jet fuel from nbutane. U.S. patent 5498811; 1996. [18] McMahon JF, Bednars C, Solomon E. Polymerization of olens as a renery process. In: Kobe KA, Mcketta JJ, editors. Advances in petroleum chemistry and rening, vol. 7. New York: Wiley; 1963. p. 285. [19] Cheng JC, Miseo S, Soled SL, Buchanan JS, Feeley JS. Lightly branched higher olen oligomerization with surface modied zeolite catalyst, U.S. patent 7759533 B2; 2010. [20] Martens LR, Verduijn JP, Mathy GM. The development of an environmental friendly catalytic system for the conversion of olens. Catal Today 1997;36:45160. [21] Skupinska J. Oligomerization of a-Olens to higher oligomers. Chem Rev 1991;91:61348. [22] Kashagulgova NS, Konovalchikov LD, Nefedov BK, Kolesnik SO. Heterogeneous catalysts for oligomerization of lower olens. Chem Technol Fuels Oils 1992;28(12):65962. [23] Schmidt R, Bruce Welch M, Randolph BB. Oligomerization of C5 olens in light catalytic naphtha. Energy Fuels 2008;22(2):114855. [24] Catani R, Mandreoli M, Rossini S, Vaccari A. Mesoporous catalysts for the synthesis of clean diesel fuels by oligomerisation of olens. Catal Today 2002;75:12531. [25] Klerk A. FischerTropsch rening: technology selection to match molecules. Royal Soc Chem, Green Chem 2008;10:124979. [26] Marcilly C. Present status and future trends in catalysis for rening and petrochemicals. J Catal 2003;216:4762. [27] Degnan Jr TF. Applications of zeolites in petroleum rening. Topics Catal 2000;13:34956. [28] Berg VD, Petrus J. Process for preparing normally liquid hydrocarbons, EP0439865A1; 1991. [29] Golombok M, de Bruijn J. Dimerization of n-butenes for high octane gasoline components. Ind Eng Chem Res 2000;39(2):26771. [30] Garwood WE, Caesar PD, Brennan JA. Light olen processing. US patent 4 150 062; 1972. [31] de Klerk A. Properties of synthetic Fuels from H-ZSM-5 oligomerization of FischerTropsch type feed materials. Energy Fuels 2007;21:30849. [32] Garwood WE, Lee W. Process for separating ethylene from light olen mixtures while producing both gasoline and fuel oil. US patent 4 227 992; 1980. [33] Tabak SA, Krambeck FJ, Garwood WE. Conversion of propylene and butylene over ZSM-5 catalyst. AIChE J 1986;32(9):152631. [34] Golombok M, Bruijn J. Dimerization of n-butenes for high octane gasoline components. Ind Eng Chem Res 2000;39:26771. [35] Europia European Industry association Annual Report. Alain Mathuren, Frederique huca, Copryrigh Europia, Brussels, Belgium; 2009. [36] Jarullah AT, Mujtaba IM, wood AS. Improvement of the middle distillate yields during crude oil hydrotreatment in a trickle-bed reactor. Energy Fuels 2011;25:77381. [37] Conor CTO. In: Ertl G, Knzinger H, Weikamp J, editors. Handbook of heterogeneous catalysis, vol. 5. Weinheim: VHC Verlagsgesellschaft mbH; 1997. p. 2380. [38] Du Toit F. Process and apparatus for the production of diesel fuels by Oligomerisation of olenic feed streams. WO0204575A2; 2002. [39] Chen CSH, Bridger RF. Shape-selective oligomerization of alkenes to nearlinear hydrocarbons by zeolite catalysis. J Catal 1996;161:68793. [40] Flego C, Marchionna M, Perego C. In: Molecular sieves: from basic research to industrial applications. Proceedings of the 3rd international zeolite symposium (3rd FEZA), Elsevier; 2005. p. 12718. [41] Liua C, Denga Y, Panb Y, Gub Y, Qiao B. Effect of ZSM-5 on the aromatization performance in cracking catalyst. J Mol Catal A: Chem 2004;215(12):1959. [42] Hartford RW, Kojima M, OConnor CT. Lanthanum ion exchange on H-ZSM-5. Ind Eng Chem Res 1989;28:174852. [43] Derouane EG, editor. Catalysts for ne chemical synthesis microporous and mesoporous solid catalysts, vol. 4. JohnWiley & Sons, Inc.; 2006. p. 126. [44] Oliveira P, Borges P, Ramos Pinto R, Lemos MANDA, Lemos F, Vdrine JC, et al. Light olen transformation over ZSM-5 zeolites with different acid strength a kinetic model. Appl Catal A Gen 2010;384(1-2):17785. [45] Villegas JI, Kumar N, Heikkila T, Lehto V-P, Salmi T, Murzin DYu. A study on the dimerization of 1-butene over beta zeolite. Topics Catal 2007;45(14):18790. [46] Rutenbeck D, Papp H, Freude D, Schwieger W. Investigations on the reaction mechanism of the skeletal isomerization of n-butenes to isobutene: Part I. Reaction mechanism on H-ZSM-5 zeolites. Appl Catal A: Gen 2001;206: 5766.

460

A. Coelho et al. / Fuel 111 (2013) 449460 [55] Wang B, Manos G. A novel thermogravimetric method for coke precursor characterisation. J Catal 2007;250(1):1217. [56] Guisnet M, Magnoux P. Organic chemistry of coke formation. Appl Catal A: Gen 2001;212:8396. [57] Cerqueira HS, Caeiro G, Costa L, Rama Ribeiro F. Review, deactivation of FCC catalysts. J Mol Catal A: Chem 2008;292:113. [58] Hopkins PD, Miller JT, Meyers BL, Ray GJ, Roginski RT, Kuehne MA, et al. Acidity and cracking activity changes during coke deactivation of ultrastable Y zeolite. Appl Catal A: Gen 1996;136:2948. [59] Kuechler KH, Brown SH, Verberckmoes AA, Puttemans MP, Silverberg SE. Hydrocarbons compositions useful for producing fuels and methods of producing the same. U.S. patent 7692049 B2; 2010. [60] Kuechler KH, Brown SH, Jaensch H, Mathys GM, Luo S, Cheng JC. Olen oligomerization to produce hydrocarbon compositions useful as fuels. U.S. patent 7678954 B2; 2010. [61] Lide DR. In: Boca Raton: CRC, Editor-in-Chief (Ed.). CRC Handbook of Chemistry and Physics, Ready-Reference book of Chemical and Physical, 85th Edn. CRC Press, London, New York, Washington, 20042005, section 3, p. 246564. [62] Choudhary VR, Devdas P. Product selectivity and aromatics distribution in aromatization of propane over H-GaMFI zeolite: inuence of temperature. Micropor Mesopor Mater 1998;23:2318. [63] Guisnet M, Andy P, Gnep NS, Benazzi E, Travers C. Skeletal isomerization of nbutenes: I. Mechanism of n-butene transformation on a non deactivated Hferrierite, catalyst. J Catal 1996;158:55160. [64] Guisnet M, Ribeiro F Rama. Deactivation and regeneration of solid catalysts, catalytic science series, vol. 9. London: Imperial College Press; 2001. ch. 1.

[47] Costa C, Lopes JM, Lemos F, Rama Ribeiro F. Activityacidity relationship in zeolite Y. Part 2. determination of the acid strength distribution by temperature programmed desorption of ammonia. J Mol Catal 1999;144:22131. [48] Costa C, Dzikh IP, Lopes JM, Lemos F, Rama Ribeiro F. Activityacidity relationship in ZSM-5. Application of Brnsted type equations. J Mol Catal 2000;154:193201. [49] Borges P, Ramos Pinto R, Lemos MANDA, Lemos F, Vedrine JC, Derouane EG, et al. Light olen transformation over ZSM-5 zeolites A kinetic model for olen consumption. Appl Catal A: Gen 2007;324:209. [50] Froment GF, Bischoff KB. Chemical reactor analysis and design. 2nd ed. Wiley; 1990. [51] Henry M, Bulut M, Vermandel W, Sels B, Jacobs P, Minoux D, et al. Low temperature conversion of linear C4 olens with acid ZSM-5 zeolites of homogeneous composition. Appl Catal A: Gen 2012;413414:6277. [52] Guisnet M, Magnoux P. In: Delmon B, Froment GF (Eds.). Catalyst eactivation, Fundamental description of deactivation and regeneration of acid zeolites, Stud. Surf. Sci. Catal, vol. 88, Proceedings of the 6th international symposium, Elsevier, Amsterdam; 1994. p. 5368. [53] Guisnet M, Magnoux P, Martin D. Roles of acidity and pore structure in the deactivation of zeolites by carbonaceous deposits. In: Bartholomew CH, Fuentes GA, editors. Catalyst deactivation, proceedings of the 7th international symposium. Amsterdam: Elsevier; 1997. p. 120. [54] Wang B, Manos G. Acid site characterization of coked USHY zeolite using temperature programmed desorption with a component-nonspecic detector. Ind Eng Chem Res 2007;46(24):797783.

Anda mungkin juga menyukai