Você está na página 1de 8

Microporous and Mesoporous Materials 102 (2007) 310317 www.elsevier.

com/locate/micromeso

Aluminum-containing SBA-15 as cracking catalyst for the production of biofuel from waste used palm oil
Yean-Sang Ooi, Subhash Bhatia
*
School of Chemical Engineering, University Science of Malaysia, Engineering Campus, 14300 Nibong Tebal, S.P.S Penang, Malaysia Received 10 September 2005; received in revised form 7 November 2006; accepted 30 December 2006 Available online 8 January 2007

Abstract Aluminum-containing SBA-15 mesoporous materials were prepared using two dierent methods in order to compare their cracking activity in gasoline production from waste used palm oil. The catalyst prepared via direct synthesis (AlSBA) possessed disorder pore size distribution whereas the catalyst prepared via post-synthesis (ACSBA) had narrow pore size distribution. Both the catalysts gave comparable activity but regenerated ACSBA exhibited higher activity and yield of gasoline fraction as compared to AlSBA. This could be attributed due to the better thermal stability of ACSBA. 2007 Elsevier Inc. All rights reserved.
Keywords: Aluminum-containing SBA-15; Catalytic cracking; Biofuel; Waste used palm oil

1. Introduction New type of ordered porous materials with combined micro- and mesopores are widely studied by the researchers due to their signicant supplementary advantages [1]. SBA15 is by far the largest mesoporous material with highly ordered hexagonally arranged mesopores, thick wall and thus with better thermal and hydrothermal stability. It has micropores that are created by the penetration of the hydrophilic poly(ethylene oxide) chain from triblock surfactant into the silica walls [1]. The template used for the synthesis of SBA-15 is relatively cheap, nontoxic and biodegradable as compared to others organic directing agents used in the preparation of MCM-41 [2,3]. This has stimulated the researchers to further improve its chemical and physical properties so that it can be used as a catalyst for the industrial processes. However, the pure siliceous SBA-15 does not possess any kind of acidity and therefore its application as a catalyst is limited. The introduction of aluminum in the frame-

Corresponding author. Tel.: +60 4 599 6409; fax: +60 4 594 1013. E-mail address: chbhatia@eng.usm.my (S. Bhatia).

work of SBA-15 to create the desired acidity has received considerable attention [46]. This is due to the increasing importance of the large molecules especially in catalytic upgrading of heavy oils [7]. The strong acidic synthesis conditions generally are not favorable for the direct incorporation of aluminum into SBA-15 [6]; therefore dierent methods have been reported in the literature for the introduction of aluminum in SBA-15 [46]. Yue et al. [4] have utilized aluminum tri-tert-butoxide as aluminum source and successively incorporated aluminum in the framework of SBA-15. Whereas, Vinu et al. [5] used aluminum isopropoxide as aluminum source in direct synthesis of AlSBA-15 with Si/Al ratio up to 7. Han et al. [6] eectively introduced aluminum in SBA-15 by a two-step procedure where the rst precursor contained zeolite nanoclusters and the second preformed precursor assembled with the triblock copolymers in strong acidic media. The aluminum incorporated in the framework of the zeolite nanoclusters was introduced into the mesoporous structure without continuous growth into larger crystals, due to the strong acidic conditions. The cracking activity of direct synthesis AlSBA-15 has been investigated for the production of biofuel from waste fatty acids mixture in our previous work [8]. It will be

1387-1811/$ - see front matter 2007 Elsevier Inc. All rights reserved. doi:10.1016/j.micromeso.2006.12.044

Y.-S. Ooi, S. Bhatia / Microporous and Mesoporous Materials 102 (2007) 310317

311

interesting to compare the catalytic properties of aluminum-containing SBA-15 prepared via post-synthesis method and direct synthesis. The cracking activity of the AlSBA-15 prepared is tested for biofuel production from waste used palm oil. The preparation method using postsynthesis strategy is found to produce catalysts with comparable acidity. In the present research, the catalytic activity of two types of aluminum-containing SBA-15 is compared for the cracking of waste used palm oil. The activity of the regenerated catalyst is also compared with the fresh catalysts activity. 2. Experimental

the maximum peak on the BJH pore size distribution. The samples were degassed for 5 h under vacuum at 573 K (for fresh catalysts) or 393 K (for deactivated catalysts) prior to the analysis. The nature of the acid sites present in the SBA-15 was determined using FTIR technique. The sample was exposed to excess pyridine for 1 h after degassing at 473 K, followed by desorption of physically adsorbed pyridine at 423 K under vacuum. The IR spectra were scanned using a PerkinElmer FTIR (Model 2000) in the range of 14001700 cm1. Transmission electron microscopy (TEM) was performed with a Philips CM12 Transmission Electron Microscope operated at 80 kV. 2.3. Activity test

2.1. Catalyst preparation 2.1.1. SBA-15 and AlSBA-15 SBA-15 and AlSBA-15 were prepared following the procedure reported in the literature [2,4]. 9.8 g of triblock copolymer poly(ethylene glycol)-poly(propylene glycol)poly(ethylene glycol) (Aldrich, average molecular weight = 5800) was dissolved with stirring in 313 ml of deionized water and 40 ml of hydrochloric acid (37 wt.%) for 1 h at 323 K. Subsequently, 21.7 g of tetraethylorthosilicate (Merck, TEOS) was added and stirred for another 10 min. The mixture was heated at 333 K for 24 h and then at 373 K for another 24 h in a Teon container under stirring. The triblock copolymer was removed upon calcination at 823 K for 6 h. Al-containing SBA-15 was synthesized using the same method as mentioned above but an appropriate amount of Al(O-sec-Bu)3 (Merck) was rst dissolved with TEOS in 20 ml of HCl solution for 3 h prior adding in to the triblock copolymer solution. The Alcontaining SBA-15 was denoted as AlSBA(X) where X is the Si/Al ratio (5, 10 and 20). 2.1.2. Post-treatment of SBA-15 Four grams of SBA-15 was dispersed in 100 ml solution of AlCl3 (Fluka) and reuxed at 353 K with stirring for 2 h. The aluminum-containing mesoporous material was ltered and thoroughly washed with deionized water until it is free of Cl ions. The powder was dried at room temperature and calcined at 823 K for 4 h. The materials were designated as ACSBA(X) where C is denoted to the chloride form of the aluminum source, and X is the Si/Al ratio (5, 10 and 20 with the amount of AlCl3 added as 1.76 g, 0.88 g and 0.44 g, respectively). 2.2. Characterization The BET surface area and pore size distribution of the catalysts were measured by nitrogen adsorption using an Autosorb I (Quanta chrome Automated Gas Sorption System). The total pore volume was estimated from the desorbed amount at relative pressure of 0.90. Mesoporosity was calculated from the desorption branch using the BJH method. The BJH pore size was dened as the position of The cracking activity of the catalysts was measured at reaction temperature of 723 K and a feed rate of waste used palm oil (weight hourly space velocity, WHSV) of 2.5 h1 at atmospheric pressure in a xed-bed micro-reactor rig reported elsewhere [9]. 0.5 g of catalyst in the form of powder was loaded over 0.2 g of quartz wool supported by a stainless steel mesh in the micro-reactor (185 mm 10 mm ID) placed in the vertical tube furnace (Model No. MTF 10/25/130, Carbolite, UK). The liquid product was collected in a glass liquid sampler, while the gaseous products were collected in a gas-sampling bulb, once the steady state was reached in the reactor. The unconverted oil was separated from the liquid product by distillation in a micro-distillation unit (Bu chi B850, GKR) at 473 K for 30 min under vacuum (100 Pa) with the pitch as the residual oil. The gaseous products were analyzed over a gas chromatograph (Hewlett Packard, Model 5890 series II) using a HP Plot Q capillary column (divinyl benzene/styrene porous polymer, 30 m long 0.53 mm ID 40 lm lm thickness) equipped with a thermal conductivity detector (TCD) and nitrogen as a carrier gas. The organic liquid product (OLP) was analyzed using a capillary glass column (Petrocol DH 50.2, lm thickness 0.5 lm, 50 m long 0.2 mm ID) at a split ratio of 1:100, using a FID detector. The composition of OLP was dened according to the boiling range of petroleum products in three categories i.e. gasoline fraction (333393 K), kerosene fraction (393453 K) and diesel fraction (453473 K). The spent catalyst was washed with acetone prior to the coke analysis. The amount of coke was determined by the dierence in weight before and after calcination in mue furnace. The catalyst regenerated by calcination in mue furnace was tested also for its cracking activity. The nature of the coke was studied using TGA analysis at a heating rate of 10 K/min, from 373 K to 973 K under a ow of oxygen gas. 3. Results and discussion 3.1. Catalyst characterization The nitrogen adsorption-desorption data of SBA-15 materials are given in Table 1. The BET surface area of

312

Y.-S. Ooi, S. Bhatia / Microporous and Mesoporous Materials 102 (2007) 310317

Table 1 Nitrogen adsorptiondesorption data for as-calcined SBA-15 materials Catalyst SBA-15 AlSBA(5) AlSBA(10) AlSBA(20) ACSBA(5) ACSBA(10) ACSBA(20) BET surface area (m2/g) 676 619 705 728 557 491 529 Pore volume (cm3/g) 1.05 0.60 0.66 0.99 0.73 0.66 0.70 Average pore size (nm) 6.0 8.6 6.1 6.1 6.1

SBA-15 was 676 m2/g with the total pore volume of 1.05 cm3/g and 6.0 nm pore size. By addition of aluminum via direct synthesis, the BET surface area increased to 728 m2/g for AlSBA(20) and 705 m2/g for AlSBA(10), respectively. However, high loading of aluminum was accompanied by a drop in the BET surface area to 619 m2/g for AlSBA(5). This phenomenon was due to the gradual loss of pore network ordering, which could be noticed from the nitrogen adsorption isotherm illustrated
1200

in Fig. 1a. The capillary condensation step on the nitrogen adsorption isotherm was found to be broad, indicating a wide range of pore sizes, as conrmed by the BJH pore size distribution (Fig. 2). It is interesting to note that the pore size of AlSBA(20) increased to 8.6 nm as compared to SBA-15. This may be due to the presence of aluminum in the silica walls [5]. In case of Si/Al ratio lower than 20, the materials showed the loss of the organized structure with the emergence of secondary pore size. The isotherms of AlSBA(10) and AlSBA(5) indicated that these materials have regular cylindrical mesopore system with the blocked openings [1]. This could be observed from the desorption branch of the isotherms and the shape of the hysteresis loop, which corresponds to ink-bottle pores. This may due to the presence of aluminum oxide species. The high loading of aluminum was hindered for direct synthesis due to the pore blockage and disorder pore structure of the material. Introduction of aluminum using post-synthesis produced mesoporous materials with lower BET surface area in the range of 491557 m2/g, which was due to the loading of aluminum into the micropores. This was observed from the decrease in the pore volume but retained the BJH pore

SBA-15
1000 800 600

AlSBA(20) AlSBA(10)
0.1
SBA-15 AlSBA(5)

Volume, cc/g

AlSBA(5)
0.08

AlSBA(10) AlSBA(20)

Dv(d), cc//g
0.2 0.4 0.6 0.8 1.0

0.06

400 200 0

0.04

0.02

0.0

P/Po
1200

10

100

1000

Pore Size,

SBA-15
1000 800 600 400 200 0

ACSBA(5) ACSBA(20) ACSBA(10)


Dv(d), cc//g

0.05
SBA-15 ACSBA(5) ACSBA(20) ACSBA(10)

Volume, cc/g

0.04 0.03

0.02 0.01

0.0

0.2

0.4

0.6

0.8

1.0

0 10 100 1000

P/Po
Fig. 1. Isotherms of nitrogen sorption for aluminum-containing SBA-15 prepared via (a) in situ method and (b) post-treatment.

Pore Size,
Fig. 2. Pore size distribution of SBA-15 compared with Al-containing SBA-15 prepared via (a) in situ method and (b) post-treatment.

Y.-S. Ooi, S. Bhatia / Microporous and Mesoporous Materials 102 (2007) 310317

313

size with identical pore size distribution after post-treatment. It can be observed from Fig. 1b that the isotherms of post-treated SBA-15 exhibited similar shape and hysteresis to that of SBA-15 except with lower adsorbed volume into intra-wall micropores (occurred at low relative pressures). When the material of low Si/Al ratio of 5 was prepared, the coverage of the aluminum started to grow along the mesopore and resulted in a narrower pore size distribution compared to SBA-15 (Fig. 2b). The slightly higher loading of aluminum was also conrmed by the lower Si/ Al content of ACSBA(5) as compared to ACSBA(10) and ACSBA(20). The adsorption in much larger pores where capillary condensation took place at relative pressures above 0.9 corresponds to secondary mesoporosity, such as interparticle voids [7]. The unchanged capillary condensation at relative pressure higher than 0.9, indicating that no aluminum was accommodated at the external surface of post-treated SBA-15. Therefore, post-treatment gave the advantages of retaining the organized pore structure, yet enabling the introduction of high amount of aluminum in SBA-15, which was dicult in the direct synthesis method. Fig. 3 shows the TEM images of the representative samples of SBA-15, AlSBA(5) and ACSBA(5). The mesopores in SBA-15 and post-treated SBA-15 (ACSBA(5)) run smoothly over several micron of length with open mesopore, whereas AlSBA(5) displays some blockage and disorder in the pore openings. A comparatively thicker wall (around 6.010.0 nm) for both the aluminum-containing SBA-15 was observed as compared to SBA-15. The thicker wall gave raise to the volume of micropores in AlSBA(5), which can be further conrmed from the pore size distribution (Fig. 2). The post-synthesis method loaded aluminum in the intra-wall micropores and along the inner surface of the mesopore and micropore; hence ACSBA(5) possesses much dense and uniform pore wall with regular pore structure. In order to identify the nature of the active sites (Lewis and Brnsted acid sites) present in the mesoporous materials prepared, FTIR spectra of the pyridine absorbed region were obtained and the results are shown in Fig. 4. The mesoporous materials show the bands associated with the combination of Lewis and Brnsted acidities (1490 cm1), hydrogen-bonding (1446 cm1; 1600 cm1) and strong Brnsted acidity (1542 cm1 and 1638 cm1) [10]. Strong Brnsted acid site at the band of 1542 cm1 only existed in SBA-15 with Si/Al ratio 5. AlSBA(10) depicted low intensity for hydrogen-bonding band which could be due to the disorder pore structure and therefore prohibited the accessibility of the hydroxyl group. Similarly, the hydrogen-bonding at 1446 cm1 and 1600 cm1 diminished upon high loading of aluminum via post-synthesis, indicating that the loading process occurred at the hydroxyl group. However, it was shown that intensity of the band at 1446 cm1 and 1600 cm1 for ACSBA(20) was retained, suggesting that the high concentration of hydroxyl group in SBA-15.

Fig. 3. TEM images of (a) SBA-15, (b) AlSBA(5) and (c) ACSBA(5).

1.1 1.0

1638 1446 1600 1491

(a)

Absorbance

0.8
1542

(b) (c) (d) (e) (f)

0.6

0.4 (g) 0.2 1700

1650

1600

1550

1500

1450

1400

Wave number, cm-1


Fig. 4. FT-IR spectra for the pyridine-adsorbed: (a) SBA-15, (b) ACSBA(5), (c) AlSBA(10), (d) ACSBA(20), (e) AlSBA(5), (f) ACSBA(10) and (g) AlSBA(20).

314

Y.-S. Ooi, S. Bhatia / Microporous and Mesoporous Materials 102 (2007) 310317

3.2. Catalytic cracking The conversion of waste used palm oil over SBA-15 and AlSBA are presented in Fig. 5a. The cracking products were mainly organic liquid product (OLP), gaseous product, water and coke. The conversion of waste used palm oil and yield of product are dened as: Conversion wt:% gas g OLP g water g coke g 100 used palm oil feed g desired product g 100 used palm oil feed g 1 2

Yield wt:%

The conversion of waste used palm oil decreased with time on stream (TOS) due to the coke formation and its deposition on the catalyst surface and the pore blockage of micropores in the intra-wall. The activity of the aluminum-containing SBA-15 deactivated at a lower rate than SBA-15 except AlSBA(10). It can be explained from the low acidity as discussed in FTIR analysis and inaccessibility of acid sites in AlSBA (10) due to the disorder pore

structure. The conversion of AlSBA(10) in waste used palm oil cracking was the lowest among the catalysts used and it deactivated after 5 h of reaction time and gave a conversion of only 50 wt.%. SBA-15 deactivated at the faster rate after 5 h of TOS due to its weak acidity. Nevertheless, the existence of hydroxyl groups enable SBA-15 to be more active than AlSBA(10) at low TOS before the conversion over both the catalysts dropped to the same level at the end of the rst cycle of reaction. After calcination in air, AlSBA(10) gave a higher conversion (5972 wt.%) as compared to the regenerated SBA-15 but slightly lower than the fresh catalyst (52 wt.% to 75 wt.%). This could be due to the loss of hydroxyl groups in SBA-15 during the calcination. AlSBA(5) and AlSBA(20) show almost identical trend of conversion but with 10 wt.% interval due to the dierence in aluminum content and acidity. AlSBA(5) with higher aluminum loading gave the conversion in the range of 7596 wt.% during 6.4 h TOS whereas AlSBA(20) gave the conversion in the range of 6886 wt%. After regeneration, the catalyst still gave high activity in waste used palm oil cracking but with lower activity as compared to the fresh catalyst. The rate of deactivation was faster for regenerated AlSBA(5) and AlSBA(20) but became constant as the TOS reached 11 h. This behavior indicated that the catalytic properties changed after the regeneration due to the thermal condition during the coke combustion. The BET surface area and pore properties of the regenerated catalyst after second cycle of reaction are given in Table 2. It can be seen from the table that the BET surface area and pore volume decreased accordingly. Secondary pores were also generated in the ACSBA catalysts with an average pore size of 2.3 nm. This could probably due to the destruction of the catalyst structure after thermal treatment and hence affected the catalytic activity. Aluminum-containing SBA-15 prepared via post-synthesis exhibited a more signicant eect of TOS in term of conversion (Fig. 5b). All the catalysts gave high conversion at initial stage of the cracking reaction (more than 90 wt.%); however, the rapid catalyst deactivation was observed. The fast deactivation rate could be attributed to the rapid pore blockage of the catalysts and hence the accessibility to the active sites was prohibited. ACSBA(5) with the growth of aluminum along the mesopore resulted in a narrower pore size distribution gave the fastest deacti-

Table 2 Nitrogen adsorptiondesorption data for regenerated SBA-15 materials Catalyst SBA-15 AlSBA(5) AlSBA(10) AlSBA(20) ACSBA(5) ACSBA(10) ACSBA(20)
a

BET surface area (m2/g) 464 577 560 542 429 374 453

Pore volume (cm3/g) 0.78 0.62 0.58 0.76 0.59 0.56 0.60

Average pore size (nm)a 5.2(2.3) 5.8(2.8) 5.6(2.8) 5.2 4.6(2.3) 4.5(2.3) 5.2(2.3)

Fig. 5. Eect of the time on stream and regeneration for the conversion of waste used palm oil cracking over SBA-15, aluminum-containing SBA-15 prepared via in situ method and post- treatment of SBA-15.

Secondary pore size is given in parentheses.

Y.-S. Ooi, S. Bhatia / Microporous and Mesoporous Materials 102 (2007) 310317

315

vation rate. Nevertheless, the catalysts recovered its activity and showed nearly the same trend of activity as that of the rst cycle of reaction after regeneration. The end of second cycle gave the conversion over ACSBA catalysts in the range of 7080 wt.%, which was higher than AlSBA catalysts (in the range of 6070 wt.%). This shows that the aluminum introduced via post-synthesis method oered a more stable catalyst in terms of cracking activity. In order to determine the deactivation rate of the catalysts, a deactivation model is proposed by assuming that the catalyst activity (u) is dependent on the time on stream (TOS), t. The rate of deactivation is presented as: du k d und dt 3

where u is the catalyst activity, kd is the deactivation rate constant (h1) and nd is the order of catalyst deactivation. The value for kd and nd were estimated from Eq. (3) using non-linear regression analysis method based on LevenbergMarquards algorithm [11]. Table 3 presents the value of kd and nd for dierent catalysts. Pure silica SBA-15 gave the highest value of deactivation rate constant and order of catalyst deactivation. This was due to the lack of acid sites in SBA-15 and deactivated at the fastest rate. AlSBA catalysts deactivated at a lower rate than ACSBA catalysts and the dierence in the values of deactivation rate constant and deactivation order can be observed in Table 3. Despite of the conversion, the yield of desired product, gasoline fraction, was the main concern in determining the most ecient catalyst. Fig. 6 shows the reaction time dependency for the yield of gasoline fraction over various catalysts. There was no correlation between the yield of gasoline fraction with the conversion especially in the rst cycle of the reaction, and the gasoline yield decreased with TOS. This phenomenon is common in cracking process due to the catalyst deactivation and dropping in further cracking towards lighter products. The yield of gasoline fraction was higher over the catalyst with lower Si/Al ratio as more acid sites were available for cracking reaction to occur. AlSBA(5) gave gasoline fraction yield as 40 wt.% at the beginning of the reaction time. The gasoline yield dropped to 23 wt.% after 6.4 h of TOS, since the catalyst was deactivated and more coke was deposited at the active sites. SBA-15 gave a maximum gasoline yield of 28 wt.% at 4.8 h TOS but dropped to the lowest gasoline yield of
Table 3 Deactivation rate constant, kd and deactivation order, n for dierent catalysts Catalyst SBA-15 AlSBA(5) AlSBA(10) AlSBA(20) ACSBA(5) ACSBA(10) ACSBA(20) kd 0.2580 0.0808 0.1750 0.1010 0.1677 0.1825 0.1136 n 2.7 1.1 1.2 1.7 2.2 1.8 1.0

Fig. 6. Eect of the time on stream and regeneration for the gasoline fraction yield of waste used palm oil cracking over SBA-15, aluminumcontaining SBA-15 prepared via in situ method and post- treatment of SBA-15.

12 wt.%. This revealed that the strength of acidity played an important role in gasoline production. This was reported in the cracking of paran where the product chain length was dependent on the strength of the Brnsted sites [12]. After regeneration, the gasoline yield as well as conversion decreased due to the changes in the catalyst structure. Unlike direct synthesis AlSBA catalysts, ACSBA catalysts also gave lower gasoline yield by following the trend of its conversion for both the fresh and regenerated catalysts. On the other hand, the overall yield of gasoline over ACSBA catalysts was higher than AlSBA catalysts. The regenerated catalysts also gave high yield of gasoline especially in the initial stage of reaction. This again conrmed that the post-synthesis catalysts exhibited more reliable activity and selectivity toward biofuel production. Coke cannot be considered as an inert with respect to the cracking reactions, as their formation could strongly inuence the catalyst performance. Two possible deactivation modes were considered (1) active site poisoning by coke coverage; (2) inaccessibility of the reactants due to

316

Y.-S. Ooi, S. Bhatia / Microporous and Mesoporous Materials 102 (2007) 310317

pore blockage [13]. In order to understand the coke deposition and the type of coke formed, the deactivated catalysts were further characterized by TGA analysis and nitrogen adsorption-desorption. Fig. 7 shows the weight loss of the coked catalysts SBA-15, AlSBA-15 and ACSBA(5) after the rst cycle of cracking reaction. The TGA analysis of the aluminum-containing SBA-15 exhibited comparable prole with two distinct weight losses at identical temperature range. The coke content was high for both the catalysts, which comprised almost 60 wt.% of the total weight of the deactivated catalysts. Whereas SBA-15 exhibited lower content of coke formation with only 30 wt.% of the total weight of deactivated catalyst. This was in agreement that cracking reaction generally occurred on the catalyst acid sites and precoke species are made available on the active site [12]. However, the open structure of SBA-15 eased the deposition of coke precursor and resulted in high coke content as compared to the microporous catalyst. The high coking was not only due to low silicaalumina ratio but also could be due to the geometrical freedom which enabled the formation of large polynuclear hydrocarbon. The BET surface area of deactivated catalysts dropped to zero especially for catalysts with low Si/Al ratio, while the BET surface area of SBA-15 was only 98 m2/g at the end of rst reaction cycle. The micropores were blocked with coke for all the catalysts since no microporosity was detected from the nitrogen adsorption-desorption analysis. It is probable that the severe coking of the catalysts started from internal pore mouth plugging. Fig. 8 compares the derivative weight curve of the selected coked catalysts after the rst and second cycle of the cracking reaction. The derivative weight curves gave two peaks in the range of 573673 K and 723873 K that corresponded to the soft and hard coke, respectively. SBA-15 with low acidity exhibited low intensity of the peaks as compared to the aluminum-containing SBA-15. The regenerated SBA-15 gave higher yield of soft coke while regenerated aluminum-containing SBA-15 show weak intensity for both the peaks. This suggests that the catalysts structure or the acidic properties can be partly

a
Derivative Weight, %/min

0 -1 -2 -3 -4 -5 -6 -7 373 473 573 673 773 873 973 SBA-15 ReSBA-15

Temperature, K

b
D erivative Weight, %/min

0 -1 -2 -3 -4 -5 -6 -7 373 473 573 673 773 873 973 AlSBA(5) ReAlSBA(5)

Temperature, K

c
Derivative Weight, %/min

0 -1 -2 -3 -4 -5 -6 -7 -8 373 473 573 673 773 873 973 ACSBA(5) ReACSBA(5)

Temperature, K
Fig. 8. Derivative weight curves for the coke analysis of fresh catalysts and regenerated catalysts after cracking process.

changed due to coking during the cracking reaction or during the thermal treatment of regeneration process. 4. Conclusions Aluminum-containing SBA-15 prepared via post-synthesis enabled the production of catalysts with low silica alumina ratio without altering their pore size distribution. Yet these catalysts have shown comparable activity with the catalysts prepared via direct synthesis in the used palm oil cracking. Furthermore, the regenerated ACSBA catalysts gave better activity and yield of gasoline fraction than AlSBA catalysts, even though the pore structure changed due to coking and regeneration steps.

Fig. 7. TGA analysis of coke for SBA-15, AlSBA(5) and ACSBA (5) after the rst cycle of the cracking reaction.

Y.-S. Ooi, S. Bhatia / Microporous and Mesoporous Materials 102 (2007) 310317

317

Acknowledgments The authors would like to acknowledge the research grant provided by the Ministry of Science, Technology and Environment, Malaysia under long term IRPA grant (Project: 02-02-05-2184 EA005), that has resulted in this article. References
[1] P. van der Voort, P.I. Ravikovitch, K.P. de Jong, M. Benjelloun, E. van Bavel, A.H. Janssen, A.V. Neimark, B.M. Weckhuysen, E.F. Vansant, J. Phys. Chem. B 106 (2002) 5873. [2] D. Zhao, J. Feng, Q. Huo, N. Melosh, G.H. Fredrickson, B.F. Chmelka, G.D. Stucky, Science 279 (1998) 548. [3] D. Zhao, Q. Huo, J. Feng, B.F. Chmelka, G.D. Stucky, J. Am. Chem. Soc. 120 (1998) 6024.

[4] Y.-H. Yue, A. Gedeon, J.-L. Bonardet, J.B. dEspinose, N. Melosh, J. Fraissard, in: A. Sayari, M. Jaroniec, T.J. Pinnavaia (Eds.), Studies on Surface Science Catalysis, vol. 129, Elsevier, Amsterdam, 2000, p. 209. [5] A. Vinu, V. Murugesan, W. Bo hlmann, M. Hartmann, J. Phys. Chem. B 108 (2004) 11496. [6] Y. Han, F.-S. Xiao, S. Wu, Y. Sun, X. Meng, D. Li, S. Lin, F. Deng, X. Ai, J. Phys. Chem. B 105 (2001) 7963. [7] E. Byambajav, Y. Ohtsuka, Appl. Catal. A 252 (2003) 193. [8] Y.-S. Ooi, R. Zakaria, A.R. Mohamed, S. Bhatia, Catal. Commun. 5 (2004) 441. [9] Y.-S. Ooi, R. Zakaria, A.R. Mohamed, S. Bhatia, Biomass and Bioenergy 27 (2004) 477. [10] B. Chakraborty, B. Viswanathan, Catal. Today 49 (1999) 253. [11] J. Ancheyta-Juarez, F. Lopez-Isunza, E. Aguilar-Rodriguez, J.C. Moreno-Mayorga, Ind. Eng. Chem. Res. 36 (1997) 5170. [12] K.A. Cumming, B.W. Wojciechowski, Catal. Rev. Sci. Eng. 38 (1996) 101. [13] M. Guisnet, P. Magnoux, Stud. Surf. Sci. Catal. 88 (1994) 53.

Você também pode gostar