Sei sulla pagina 1di 9

Chemical Engineering Journal 176177 (2011) 202210

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

The synthesis and characterization of aluminum loaded SBA-type materials as catalyst for polypropylene degradation reaction
Zeynep Obal, Naime Asl Sezgi , Timur Do gu
Chemical Engineering Department, Middle East Technical University, Inonu Boulevard, 06531 Ankara, Turkey

a r t i c l e

i n f o

a b s t r a c t
The performance of pure and aluminum containing SBA-type catalysts prepared using different aluminum sources and different Al/Si ratios were investigated in the polypropylene degradation reaction using a thermogravimetric analyzer. For the synthesis of catalysts, aluminum isopropoxide and aluminum sulphate were used as the aluminum sources. Synthesized materials had high surface areas and exhibited nitrogen adsorption isotherms of type IV. EDS results showed that the aluminum incorporation into the structure was very effective. 27 Al MAS NMR spectra of the catalysts exhibited tetrahedrally and octahedrally coordinated aluminum species in the structure. TEM images of synthesized catalysts showed well-ordered hexagonal arrays of uniform mesopores with cylindrical channels. With the aluminum loading, DRIFTS analysis of the pyridine adsorbed materials revealed the existence of Brnsted acid sites in the synthesized catalysts in addition to Lewis acid sites. Thermogravimetric analysis results showed a marked reduction in the degradation temperature in the presence of aluminum containing SBA-type catalysts. The activation energy value of the degradation reaction decreased to about 5170 kJ/mole in the presence of catalysts synthesized using aluminum sulphate as the aluminum source and when aluminum isopropoxide was used as the aluminum source activation energy value of the reaction decreased to 8289 kJ/mol. 2011 Elsevier B.V. All rights reserved.

Article history: Received 30 November 2010 Received in revised form 5 April 2011 Accepted 13 April 2011 Keywords: Catalytic degradation Polypropylene Mesoporous Catalyst-SBA15-TGA

1. Introduction Consumption of plastic products has been signicantly increasing for the last century. This has a negative effect on the environment due to the generation of huge amount of waste plastics. To overcome this environmental problem, several disposal methods are being offered, namely landlling, incineration and recycling. Most of these waste plastics are non-biodegradable; therefore they are being felt unsuitable for landll disposal. Additionally, plastic waste is more voluminous than the other waste types, so providing suitable and safe depots for them is more difcult and expensive. Incineration of these waste plastics can be another solution, but it often generates harmful emissions like nitrous and sulphur oxides, dusts, dioxins and other toxins, depending on the nature of the waste polymers. By these methods, the potential energy content of waste plastics is also wasted [14]. Therefore, conversion of these plastic wastes to valuable and useful chemicals can be a good solution to the pollution problem. Additionally, products obtained by conversion of these waste plastics can be a good alternative to current petroleum products.

Corresponding author. Tel.: +90 312 210 26 08; fax: +90 312 210 26 00. E-mail addresses: z obali@metu.edu.tr (Z. Obal), sezgi@metu.edu.tr (N.A. Sezgi), gu). tdogu@metu.edu.tr (T. Do 1385-8947/$ see front matter 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.cej.2011.04.063

Several methods are offered for the conversion of waste plastics. Non-catalytic and catalytic thermal degradation of waste plastics are two of these methods [35]. Non-catalytic thermal degradation of polymers is attractive, but unfortunately this method leads towards a broad distribution of hydrocarbons up to waxy products. Additionally, in this method, higher temperatures greater than 500 C are needed to receive light hydrocarbons [6]. In contrast, catalytic thermal degradation process allows the plastic degradation to be performed at lower temperatures while the product distribution can be controlled by a right selection of the catalyst being used [7]. The most commonly used catalysts in the catalytic thermal degradation of polymers are zeolites (HZSM-5, HY, HUSY, etc.) [6,8,9], amorphous SiO2 Al2 O3 [10] and activated carbon [11]. Zeolites show excellent properties as catalysts for the catalytic thermal degradation of polymers due to their strong acidity for carboncarbon bond scission. However, the pore size of these catalysts is limited to maximum value of about 1.0 nm which hinders the access of bulky molecules to the acid sites located inside the channels [12]. Also, coking causes pore closure and hence catalyst deactivation. Therefore, catalysts with larger pore diameters (between 2 and 50 nm) have been developed in recent years. Mesoporous molecular sieves were rst discovered in 1992 [13], which were designated as M41S family. These materials are mainly made up of SiO2 and allow faster diffusion of large organic

Z. Obal et al. / Chemical Engineering Journal 176177 (2011) 202210

203

molecules than zeolites. The most extensively studied member of these mesoporous materials is MCM-41, which has exible synthesis conditions and advantageous properties such as high surface area and narrow pore size distribution [14]. M41S materials are advantageous in many reactions, as mentioned above, but they are limited to a pore diameter of Due to their low pore wall thickness (12 nm) approximately 80 A. thermal stability of these materials is also reported as low. These characteristics limit the usage of these materials in size-selective reactions [15,16]. To overcome this situation, Zhao et al. extended the family of highly ordered mesoporous silicates by synthesizing SBA-type materials in 1998 [17,18]. These materials have pore sizes and non-ionic block copolymers as strucbetween 20 and 300 A ture directing agents were used. These materials are synthesized in highly acidic media. SBA-15 is the most popular member of this family. It is synthesized using triblock copolymer poly(ethylene oxide)-poly(propylene oxide)-poly(ethylene oxide) as the surfactant. These materials possess large surface area (>700 m2 /g) with large pore diameter and large pore wall thickness. Large pore wall thickness of SBA-type materials results in high hydrothermal stability. In other words, these materials were reported to have higher hydrothermal stability and thicker pore wall than M41S materials [16]. Use of these mesoporous materials as catalysts in purely siliceous form is restricted due to their relative low acidity, in comparison with those of microporous zeolites such as HZSM-5 [12]. In order to improve the catalytic performance of these materials, incorporation of aluminum and other metal ions (Fe+3 , Pd+4 , V+5 , etc.) into the framework of these materials is needed [5,1921]. Incorporation of aluminum ions into such mesoporous materials creates Brnsted acid sites, solving the problem of the low acidity of silicate structured mesoporous materials [12]. The most commonly used aluminum sources are sodium aluminate, aluminum sulphate, aluminum nitrate, aluminum isopropoxide and aluminum hydroxide [2224]. Much less is known about the catalytic activity of mesoporous SBA catalysts in cracking of polymers especially in polypropylene. Therefore, the objective of this study is to synthesize aluminum containing SBA-15 catalysts using different aluminum sources and Al/Si ratios and to examine their performances in the polypropylene degradation reaction.

1.1 g of aluminum source, depending on the Al/Si ratio, was added dropwisely and the resulting solution was stirred at room temperature. Then, the temperature was raised above room temperature to evaporate the solvent completely. After complete evaporation of the solvent, the solid product was dried under vacuum for 24 h. The samples prepared from different aluminum sources were designated as ALS-xA and ALS-xB, where x represents the Al/Si molar ratio of the synthesis mixture and A and B denote aluminum sulphate (Al2 (SO4 )3 18H2 O, Merck) and aluminum isopropoxide (C9 H21 AlO3 , Merck) as the aluminum sources, respectively.

2.2. Catalyst characterization The synthesized materials were characterized using several techniques. The X-ray diffraction (XRD) analysis of the synthesized materials was performed using Rigaku D/MAX2200 diffractometer with a CuK radiation source. The scanning range of 2 was set between 0.8 and 10 with a scanning speed of 1 per min. surface area, average pore diameter, and The adsorptiondesorption isotherms of the synthesized material were measured by the nitrogen adsorption technique using Micromeritics Gemini V1 apparatus. Scanning electron microscopy (SEM) experiments were performed on QUANTA 400F Field Emission SEM apparatus operating at 30 kV to determine the morphology of the synthesized materials. The elemental composition of the synthesized materials was determined using EDS analysis equipment coupled with JEOL 6400 apparatus. Transmission electron microscopy (TEM) images of the synthesized materials were obtained at the Electron Microscopy Center of the Faculty of Biology and Geology of Babes-Bolyai University in Romania, using a JEOL Model JEM 1010 transmission electron microscope equipped with SIS MegaView CCD camera. The coordination environment of aluminum atoms in the materials was determined using 27 Al MAS NMR. Their spectra were recorded at room temperature on a BRUKER Avance 300 Spectrometer with a resonance frequency of 78.1 MHz. The magnetic eld, the spin rate of the samples and the number of scans were 7.05 T, 5 kHz, and 12,000, respectively. The pulse length was adjusted to 3.86 s with a relaxation delay of 1 s. External Al(H2 O)6 3+ was used as a reference. Diffuse reectance FT-IR spectroscopy (DRIFTS) analyses of synthesized materials were carried out using a Perkin-Elmer Spectrum1 FT-IR instrument. DRIFTS spectra of the samples were obtained at room temperature and by placing the sample into the sample pan of the DRIFTS cell. DRIFTS spectra of these samples were obtained at a wavenumber range of 4004000 cm1 with the resolution of 4 cm1 . The synthesized material was mixed with KBr with a weight ratio of 1/20. Additionally, to determine the nature of acid sites in the synthesized materials, DRIFTS spectra of the pyridine adsorbed samples were obtained under same conditions using the same FT-IR instrument.

2. Experimental 2.1. Catalyst preparation Pure SBA-15 was synthesized following a direct hydrothermal synthesis method [25,26]. Firstly, a certain amount of triblock copolymer, Pluronic P123 (Aldrich) was dispersed in 2 M of HCl solution and stirred for 4 h at 40 C. Then, tetraethyl orthosilicate (TEOS, Merck) solution was added to this solution by continuously mixing. The resulting gel was aged at 40 C for 2 h. After that, it was transferred into a Teon-lined stainless steel autoclave for hydrothermal synthesis at 100 C for 48 h. The solid product was ltered and washed with deionized water to remove excess HCl and dried at 80 C. Finally, the solid product was calcined at 540 C for 8 h in owing air to decompose triblock copolymer. This sample was named as S2 . Aluminum loaded SBA-15 catalysts were prepared by impregnation method. The procedure used in this method was a modied version of the procedure described in literature [27,28]. In this method, 1 g of the pure SBA-15 silica, prepared according to the procedure given above, was dried under vacuum for 2 h to remove its humidity. After that, 30 ml of deionized water was added and stirred for 2 h. At the end of these 2 h period, in the range of 0.12 to

2.3. Thermogravimetric analysis(TGA) The performance of the synthesized SBA-type catalysts in the polypropylene degradation reaction was examined using PerkinElmer Thermal Analyzer. The polypropylene used in this work was provided from Aldrich and had the following features: isotactic, Mw : 250,000, :0.9 g/ml, melt index: 12 g/10 min, mp :160165 C. Thermogravimetric analysis was performed under nitrogen atmosphere in the temperature range of 35550 C with a constant heating rate of 5 C/min. In catalytic cracking experiments, samples were prepared with a polymer to catalyst weight ratio of 2.

204

Z. Obal et al. / Chemical Engineering Journal 176177 (2011) 202210 Table 1 Al/Si atomic ratios of the synthesized Al-SBA-type catalysts. Sample ID ALS-0.03A ALS-0.1A ALS-0.2A ALS-0.03B ALS-0.1B ALS-0.2B Al/Si (EDS) 0.03 0.07 0.17 0.026 0.08 0.14 Al/Si (in the initial gel solution) 0.03 0.1 0.2 0.03 0.1 0.2

Fig. 1. XRD pattern of pure SBA-15 (S2 ).

3. Results and discussion 3.1. Characterization of catalysts XRD patterns of pure and aluminum containing SBA-type materials are given in Figs. 1 and 2, respectively. For these materials, the diffraction patterns have reection peaks only in the low Bragg angle range (less than 10). A typical SBA-15 XRD pattern shows three reections corresponding to Bragg angle values of 0.92 , 1.60 and 1.85 [17]. For the sample S2 (Fig. 1), the major peak was at a 2 value of 0.98 and the second and the third peaks were at 1.64 and 1.86 , respectively. When aluminum sulphate was used as the aluminum source, the intensity of the peaks decreased and

the well-ordered structure of SBA-15 deteriorated with an increase in aluminum loading (Fig. 2(a)). On the other hand, when aluminum isopropoxide was used as the aluminum source, some deformations in the main peaks of SBA-15 were observed which may be due to the structural irregularity (Fig. 2(b)). The Al/Si atomic ratios of aluminum loaded SBA-type materials evaluated from EDS analyses are reported in Table 1. These results showed that aluminum was successfully incorporated into the structure of SBA-15 both at low and high aluminum loadings for both aluminum sources. It could be said that impregnation method is generally more successful than hydrothermal synthesis method in the incorporation of aluminum into SBA-15 structure since a highly acidic synthesis gel is not favorable for aluminum incorporation [29]. The nitrogen adsorptiondesorption isotherms of synthesized SBA-type catalysts are shown in Fig. 3. According to the IUPAC clas-

Fig. 2. XRD patterns of Al-SBA-type catalysts synthesized using different aluminum sources; (a) Al2 (SO4 )3 18H2 O, and (b) C9 H21 AlO3 .

Fig. 3. N2 adsorptiondesorption isotherms of Al-SBA-type catalysts synthesized using different aluminum sources; (a) Al2 (SO4 )3 18H2 O and (b) C9 H21 AlO3 (solid boxes: adsorption branch; blank boxes: desorption branch).

Z. Obal et al. / Chemical Engineering Journal 176177 (2011) 202210

205

Fig. 4. Pore size distributions of Al-SBA-type catalysts synthesized using different aluminum sources; (a) Al2 (SO4 )3 18H2 O, and (b) C9 H21 AlO3 .

sication, all the samples showed Type IV adsorption/desorption isotherms with a H1 hysteresis loop, which is characteristic of mesoporous materials [30]. This hysteresis loop was broad and the adsorption and desorption branches were parallel in most of the synthesized materials (Fig. 3). A well-dened capillary condensation step was observed at a relative pressure range of 0.550.84, indicating the formation of uniform pore size distributions. Additionally, the hysteresis opening decreased with an increase in aluminum loading. The adsorbed N2 volumes in the materials synthesized using Al2 (SO4 )3 signicantly decreased with an increase in aluminum loading and pore volumes of these materials(Table 2) also decreased. This might be due to the blocking of the mesopores of SBA-15 by aluminum species (Fig. 3(a)). However, the adsorbed N2 volume in the materials synthesized using C9 H21 AlO3 did not decrease signicantly (Fig. 3(b)) and their pore volume values approximately remained constant. Results reported in our work are quite similar to the results of Zukal
Table 2 The physical properties of the synthesized Al-SBA-type catalysts. Sample ID Surface area BET (m2 /g) Pore volume BJH desorption (cc/g)a 1 03 0.87 0.61 0.31 0.86 0.84 0.85 Average pore diameter dp(nm) (4V/AbyBET) 6.0 6.5 6.4 6.2 6.0 6.4 6.5 Microporosity (%)

Fig. 5. SEM images of the materials synthesized using different aluminum sources; (a) aluminum sulphate ALS-0.2A, and (b) aluminum isopropoxide ALS0.2B.

S2 ALS-0.03A ALS-0.1A ALS-0.2A ALS-0.03B ALS-0.1 B ALS-0.2B

822 560 368 200 730 575 617

24.6 23.3 22.3 24.7 24.3 23.1 21.6

a These BJH pore volumes were measured in the range between 1.7 nm and 300 nm.

et al. [28]. In Zukals study, with an increase in aluminum loading, the adsorbed N2 volumes in the materials synthesized using aluminum chlorhydrol did not decrease much. In our study, the same trend was observed for the materials synthesized using aluminum isopropoxide. However, the adsorbed N2 volumes in the materials synthesized using aluminum sulphate decreased signicantly with an increase in aluminum loading. Aluminum source used in the synthesized material affects the textural properties of the synthesized materials. Use of different aluminum source in synthesized material also affects pore size distribution. Physical properties of the synthesized materials are given in Table 2. The BET surface areas of the materials synthesized using aluminum isopropoxide as the aluminum source decreased with the amount of aluminum loading. This could be an indication of the aluminum incorporation into pore surface by modifying its pore diameter. On the other hand, when aluminum sulphate was used as the aluminum source, the surface area of the materials decreased signicantly. This might be due to the blocking of materials micropores and the lling of the corona that surrounds the mesopores of SBA-15 by Al species. The successive incorporation of aluminum into the SBA-15 material may smooth the mesopore surface [28]. All the samples contained considerable amount of micropores, approximately 24% of the pores. These micropores interconnect mesopores to each other. In fact, a microporous corona region surrounding the ordered mesopores of SBA-15 is expected. The pore size distributions of synthesized SBA-type catalysts, calculated by BJH model from desorption branch, are given in

206

Z. Obal et al. / Chemical Engineering Journal 176177 (2011) 202210

Fig. 6. TEM images of the materials synthesized using aluminum sulphate (a) ALS-0.03A, (b) ALS-0.2A, and aluminum isopropoxide (c) ALS-0.03B, (d) ALS0.2B as the aluminum source.

Fig. 4. All the catalysts showed unimodal pore size distributions in the mesopore range and these pore size distribution curves supported the ordered mesoporosity of these materials. The average pore diameters of all the synthesized materials are in the range of 6.06.5 nm, which is an indication of mesoporous structure. Although the same Al/Si molar ratios for different aluminum sources were used in the synthesis of catalysts, difference in their textural properties was observed. Most probably, this difference was due to the presence of sulphate in the catalysts synthesized using aluminum sulphate as aluminum source. Sulphur was observed in the EDS analysis of the catalysts synthesized using aluminum sulphate. In other words, use of different aluminum source in synthesis makes this textural difference in these materials. SEM images (Fig. 5) showed that when aluminum sulphate was used as the aluminum source, cylindrical particles were formed (Fig. 5(a)). Also, agglomeration was observed in catalysts synthesized using aluminum sulphate as the aluminum source. On the other hand, when aluminum isopropoxide was used as the aluminum source, the length of the particles shortened and hexagonal-like structures were formed (Fig. 5(b)). For instance, for catalyst ALS-0.2B, the particle size is approximately 0.8 m

(Fig. 5(b)), on the other hand for ALS-0.2A, particle size becomes approximately 1 m (Fig. 5(a)). As a conclusion, the nature of the aluminum source used in the synthesis of catalyst has a signicant effect on the particle size and morphology. The TEM images of the synthesized materials are given in Fig. 6. These images showed well-ordered hexagonal arrays of uniform mesopores with channels. The pore structure of these materials was in honeycomb form that is common for mesoporous materials. The lighter regions in Fig. 6(b) indicated the presence of mesopores in the synthesized materials. The pore channels arranged regularly in cylindrical form both in aluminum sulphate and aluminum isopropoxide samples. These particles are combined with each other in an ordered manner and the particle size became 0.6 m approximately (Fig. 6(a) and (b)) when aluminum sulphate was used as the aluminum source. This result is in agreement with the result obtained from SEM analysis. On the other hand, these particles were shortened when aluminum isopropoxide was used as the aluminum source (Fig. 6(c) and (d)). When aluminum sulphate was used as the aluminum source, the average thickness of the wall was 4.8 nm and the pore diameter was around 6.0 nm, on the other hand in the case of using aluminum isopropoxide as the aluminum

Z. Obal et al. / Chemical Engineering Journal 176177 (2011) 202210

207

Fig. 7.

27

Al MAS NMR spectra of (a) ALS-0.03A, (b) ALS-0.2A, (c) ALS-0.03B, (d) ALS0.2B materials.

source the average wall thickness was 4.7 nm and the pore diameter was around 6.2 nm. All these results are in agreement with results obtained from N2 physisorption analysis (Table 2). The aluminum distribution in the SBA-15 material is very important in terms of catalytically active centers in the SBA-15 catalyst. Cations related with the oxygen atoms bonded to the aluminum atom at a nearest or next nearest neighbor affect Bronsted acidity [31]. The aluminum atoms may be located either in the channel surface or inside the channel walls [31]. Therefore, 27 Al NMR analysis was performed. 27 Al MAS NMR spectra of the synthesized materials using different aluminum sources are given in Fig. 7. The spectra of these materials showed two peaks at around 0 and 50 ppm. The peak at 50 ppm is assigned to tetrahedral framework aluminum formed in the mesoporous walls of the material. The peak at 0 ppm belongs to octahedral aluminum corresponding to extra-framework aluminum species. The NMR results showed that aluminum was incorporated into the SBA-15 structure, which are consistent with the EDS results. For the materials synthesized using aluminum sulphate, the peak intensity at 0 ppm increased with an increase in aluminum loading (Fig. 7(a) and (b)). In other words, most of the aluminum was not incorporated into framework of the catalyst structure. Probably, it was adsorbed on the catalyst surface. On the other hand, for the materials synthesized using aluminum isopropoxide, the peak at 50 ppm was more intense than the peak at

0 ppm when the aluminum content was low (Fig. 7(c) and (d)). For the materials synthesized using aluminum isopropoxide, with an increase in Al loading, while the peak intensity at 50 ppm decreased, the peak intensity at 0 ppm increased. This indicates adsorption of aluminum species on the materials surface. These results showed that the nature of the aluminum source used in the synthesis of the catalyst has an important effect on the location of aluminum in the SBA-15 framework. DRIFTS results of fresh catalysts are given in Fig. 8. The broad band between 4000 and 3000 cm1 is referred to the hydrogen bonded silanol groups and adsorbed water and the band observed at 3740 cm1 corresponds to free silanol groups [32,33]. The intensity of the band at 3740 cm1 decreased with an increase in aluminum loading. Probably, this is due to the bonding of aluminum to those free silanol groups. Additionally, the band around 1620 cm1 is generally associated to the bending of HOH from adsorbed water [3234]. Finally, the band observed at 1056 cm1 together with a shoulder at 1204 cm1 is referred to the asymmetric SiOSi stretching vibrations. There is also band at 790 cm1 , due to SiOSi symmetric stretching vibration [35]. It is remarkable that the incorporation of aluminum into catalyst structure causes a decrease in the intensity of the component assigned to the SiOSi symmetric stretching mode at 790 cm1 [35]. The band at 950 cm1 is assigned to the surface silanol groups. Incorporation of aluminum

208

Z. Obal et al. / Chemical Engineering Journal 176177 (2011) 202210

Fig. 8. DRIFTS spectra of Al-SBA-type catalysts synthesized using different aluminum sources; (a) Al2 (SO4 )3 18H2 O, and (b) C9 H21 AlO3 .

decreased the intensity of this peak, which indicates interaction of silanol groups with the aluminum species [34,35]. The peak around 1380 cm1 is assigned to SiOSi vibration [36]. The difference between DRIFTS spectra of pyridine adsorbed samples and fresh catalysts (Fig. 9) gave information about the Lewis and Brnsted acid sites in these materials. In other words,

the DRIFTS analysis is a good method to evaluate qualitatively relative strengths of Brnsted and Lewis acid sites [18]. Before FTS analysis of pyridine adsorbed materials, samples were the DRI dried under vacuum at 110 C for 9 h. DRIFTS bands observed at 1447 cm1 and 1598 cm1 corresponds to the Lewis acid sites in the material [37,38]. On the other hand, the bands observed at 1540 cm1 and 1640 cm1 are due to the Brnsted acid sites in the material. The band observed at 1489 cm1 is due to the contributions of both Lewis and Brnsted acid sites within the structure. Pure SBA-15 (S2 ) had only Lewis acid sites. But with aluminum loading, the presence of Brnsted acid sites in materials structure was also observed. In the catalyst synthesized using aluminum sulphate, IR bands corresponding to Lewis acid sites (1447 cm1 and 1598 cm1 ) were almost completely disappeared with aluminum loading. However, the band corresponding to Brnsted acid sites was strong in these materials. The relative intensity of the band corresponding to Brnsted acid sites (1540 cm1 ) signicantly increased with an increase in the aluminum loading. In the catalyst synthesized using aluminum isopropoxide, the relative intensity of the band corresponding to Lewis acid sites (1447 cm1 ) was higher than the band corresponding to Brnsted acid sites (1540 cm1 ). For these materials which were prepared using aluminum sulphate as the aluminum source, it was clear that Brnsted acid sites were much stronger than Lewis acid sites (Fig. 9(a)). On the other hand, Brnsted acid sites in the materials synthesized using aluminum isopropoxide were less strong than Lewis acid sites (Fig. 9(b)). Acid strengths of the sites and acid capacities of the catalysts affect catalysts performance in polypropylene degradation reaction. Hammett acidity function (Ho) of the synthesized catalysts was measured using an amine titration method, which is a good method to evaluate acid strength [39]. Titration results showed that the Ho values of all aluminum loaded SBA-15 catalysts were between +3.3 and +0.8. However the Ho value of pure SBA-15 was between +4.8 and +3.3, which is weaker with respect to aluminum incorporated SBA-15. Incorporation of aluminum into the synthesized SBA-15 materials increased the acid strength of the material. 3.2. Determination of kinetic parameters from TGA data Fig. 10 illustrates TGA plots resulted from both catalytic and non-catalytic thermal degradation of polypropylene. Pure polypropylene shows a steep weight loss in the temperature range of 350480 C. This steep weight loss was due to the chain degradation. Pure SBA-15 only possessed Lewis acid sites as shown in DRIFTS results of pyridine adsorbed materials. Incorporation of aluminum into the mesoporous SBA-15 material created the Brnsted acid sites and made the material more acidic. Therefore, degradation temperature decreased to a lower value in the presence of aluminum containing SBA-15 catalysts (Fig. 10). For instance, in the case of using aluminum sulphate as the aluminum source, the relative intensity ratio of Brnsted acid sites band to Lewis acid sites band (Fig. 9(a)) increased with an increase in aluminum loading so the polymer degradation temperature sifted to lower temperatures (Fig. 10(a)). On the other hand, when aluminum isopropoxide was used as the aluminum source the relative intensity ratio of Brnsted acid sites band to Lewis acid sites band increased with an increase of Al/Si ratio from 0.03 to 0.1 and then decreased with an increase of Al/Si ratio from 0.1 to 0.2. The same trend was observed in TGA analysis. ALS-0.1B catalyst lowered the degradation temperature much more than ALS-0.2 catalyst (Fig. 10(b)). This may be due to presence of non-accessible aluminum with increasing aluminum concentration in the SBA-15 material [31,40]. As a conclusion, decrease of Al/Si ratio is expected to cause a decrease in acidity of such catalysts and decrease of acidity is expected to have a negative effect on the activity of the catalyst for the degradation reactions. Degree of incorporation of Al into the catalyst structure and consequently

Fig. 9. DRIFTS spectra of the pyridine adsorbed Al-SBA-type catalysts synthesized using different aluminum sources; (a) Al2 (SO4 )3 18H2 O, and (b) C9 H21 AlO3 .

Z. Obal et al. / Chemical Engineering Journal 176177 (2011) 202210

209

Table 3 Activation energy value for the polypropylene degradation reaction in the presence of Al-SBA-type catalysts. Sample ID No catalyst SBA-15(S2) ALS-0.03A ALS-0.1A ALS-0.2A ALS-0.03B ALS-0.1B ALS-0.2B Activation energy (EA ) Values (kJ/mol) 172.0 139.7 69.9 56.1 50.7 89.2 86.3 81.9

Fig. 10. TGA plots describing the degradation of polypropylene over Al-SBA-type catalysts synthesized using different aluminum sources; (a)Al2 (SO4 )3 18H2 O, and (b)C9 H21 AlO3 .

the Al/Si ratio in the catalyst matrix are expected to have signicant effects on the activity of such catalysts. Using the TGA data, kinetic parameters for the polypropylene degradation reaction were evaluated in the absence and presence of aluminum containing SBA-type catalysts following the procedure reported in our previous study [5]. A standard power law model was used to describe the kinetics of polypropylene degradation: d = AexpEA /RT (1 )n dt (1)

where is the fraction of polymer decomposed at time t, n is the overall reaction order and A and EA are the pre-exponential factor and activation energy of the reaction, respectively. Here is dened as. = 1 ft 1 f (2)

in the absence of catalyst was found as 172 kJ/mol. When pure SBA15 (S2 ) was used as the catalyst in degradation reaction, it caused a decrease of activation energy to 140 kJ/mol. On the other hand, when aluminum loaded catalysts were used in degradation reaction, the activation energy of the reaction signicantly decreased because of the presence of the Brnsted acid sites in aluminum loaded SBA-15 catalysts as shown in the DRIFTS results. In other words, the acid strength of the SBA-15 materials was increased with aluminum loading and the activation energy of the reaction decreased more compared to the pure SBA-15 material. For instance, in the case using aluminum isopropoxide as the aluminum source in the synthesis of the catalysts, the activation energy values were obtained in the range of 8289 kJ/mol depending on the amount of aluminum loading and in the case using aluminum sulphate as the aluminum source in the synthesis of the catalysts, the activation energy values of degradation were found to be decreased down to about 5180 kJ/mol. These results were consistent with the DRIFTS results because the relative intensity ratio of Brnsted acid sites to Lewis acid sites of the catalysts synthesized using aluminum sulphate as the aluminum source were higher than that of the catalysts synthesized using aluminum isopropoxide. As a result, the presence of Brnsted acid sites has a positive effect on the activity of catalyst for the degradation reactions [40]. Gas chromatographic analysis indicated that in the catalytic pyrolysis reaction, the gas product was mainly composed of ethylene. In addition to ethylene, some amounts of propylene and butane gasses were also observed. In the liquid product, C7 hydrocarbons were the major components but some amounts of hydrocarbons in the range of C5C18 were also obtained. 4. Conclusions In this study, well-ordered mesoporous SBA-type catalysts from different aluminum sources with different Al/Si ratios were successfully synthesized and their performance was tested in the polypropylene degradation reaction using thermal analyzer. In the XRD patterns of the catalysts synthesized using aluminum isopropoxide as the aluminum source, deformation in the main peak of SBA-15 was observed which may be due to the structural irregularity. On the other hand, the well-ordered structure of SBA-15 did not deteriorate too much when the level of aluminum was kept low in the case of using aluminum sulphate as the aluminum source. All the synthesized catalysts showed typical Type IV isotherms with a H1 hysteresis loop and they have high surface areas. TEM images of SBA-type materials showed well-ordered hexagonal arrays of uniform mesopores with channels. The materials pore structure was in honeycomb form that is common for mesoporous materials. Aluminum incorporated effectively in all these catalysts. 27 Al MAS NMR spectra of the catalysts exhibited a mixture of tetrahedral and octahedral aluminum species. The Brnsted acid sites were much stronger than Lewis acids in the aluminum loaded SBA-type catalysts synthesized using aluminum sulphate. The TGA results showed a marked reduction in the degradation temperature in the

where ft and f are the instantaneous and the nal values of the weight fractions, respectively and by inserting a linear heating rate equation, q = dT/dt into Eq. (1) and integrating it yields,for the higher order reaction(n = / 1) ln 1 (1 )(1n) AR EA = ln qEA RT (1 n)T 2 for rst order reaction (n = 1) ln(1 ) AR EA ln = ln qEA RT T2 (4) (3)

By using the values, a plot of left hand side of Eqs. (3) and (4) versus 1/T should give straight lines for a value of n. The Arrhenius parameters, A and EA , can be estimated from the intercept and the slope of this line. The activation energy values for polypropylene degradation reaction evaluated from Eq. (4) are given in Table 3. Additionally, the overall reaction order was found to be one for all the synthesized catalysts. The activation energy of polypropylene degradation

210

Z. Obal et al. / Chemical Engineering Journal 176177 (2011) 202210 [15] A. Vinu, V. Murugesan, W. Bhlmann, M. Hartmann, An optimized procedure for the synthesis of AlSBA-15 with large pore diameter and high aluminum content, J. Phys. Chem. B. 108 (2004) 1149611505. [16] A. Katiyar, S. Yadav, P.G. Smirniotis, N.G. Pinto, Synthesis of ordered large pore SBA-15 spherical particles for adsorption of biomolecules, J. Chromatogr. A 1122 (2006) 1320. [17] D. Zhao, J. Feng, Q. Huo, N. Melosh, G.H. Fredrickson, B.F. Chmelka, G.D. Stucky, Triblock copolymer syntheses of mesoporous silica with periodic 50 to 300 angstrom pores, Science 279 (1998) 548552. [18] D. Zhao, Q. Huo, J. Feng, B.F. Chmelka, G.D. Stucky, Nonionic triblock and star diblock copolymer and oligomeric surfactant syntheses of highly ordered, hydrothermally stable, mesoporous silica structures, J. Am. Chem. Soc. 120 (1998) 60246036. [19] G. Oye, J. Sjblom, M. Stcker, Synthesis, characterization and potential applications of new materials in the mesoporous range, Adv. Colloid Interface Sci. 8990 (2001) 439466. [20] A. Taguchi, F. Schth, Ordered mesoporous materials in catalysis, Microporous Mesoporous Mater. 77 (2005) 145. ener, T. Do gu, G. Do gu, Effects of synthesis conditions on the structure of [21] C. S Pd incorporated MCM-41 type mesoporous nanocomposite catalytic materials with high Pd/Si ratios, Microporous Mesoporous Mater. 94 (2006) 8998. [22] S. Biz, M.G. White, Syntheses of aluminosilicate mesostructures with high aluminum content, J. Phys. Chem. B. 103 (1999) 84328442. [23] Y. Cesteros, G.L. Haller, Several factors affecting Al-MCM-41 synthesis, Microporous Mesoporous Mater. 43 (2001) 171179. [24] G.A. Eimer, L.B. Pierella, G.A. Monti, O.A. Anunziata, Preparation and characterization of aluminum-containing MCM-41, Catal. Commun. 4 (2003) 118123. [25] P.F. Fulvio, S. Pikus, M. Jaroniec, Short-time synthesis of SBA-15 using various silica sources, J. Colloid Interface Sci. 287 (2005) 717720. [26] S.A. Mirji, S.B. Halligudi, D.P. Sawant, N.E. Jacob, K.R. Patil, A.B. Gaikward, S.D. Pradhan, Adsorption of octadecyltrichlorosilane on mesoporous SBA-15, Appl. Surf. Sci. 252 (2006) 40974103. [27] C.S ener, Synthesis and characterization of Pd-MCM-type mesoporous nanocomposite materials, MSc Thesis, Middle East Technical University, Ankara, Turkey, 2006. [28] A. Zukal, H. Siklova, J. Cejka, Grafting of alumina on SBA-15: effect of surface roughness, Langmuir 24 (2008) 98379842. [29] A. Vinu, V. Murugesan, W. Bhlmann, M. Hartmann, An optimized procedure for the synthesis of AlSBA-15 with large pore diameter and high aluminum content, J. Phys. Chem. B 108 (2004) 1149611505. [30] Y. Li, Q. Yang, J. Yang, C. Li, Synthesis of mesoporous aluminosilicates with low Al/Si ratios using a single precursor under acidic conditions, J. Porous Mater. 13 (2006) 187193. [31] J. Dedecek, N. Zilkova, J. Cejka, Experimental study of the effect of Si/Al composition on the aluminum distribution in (Al)MCM-41, Microporous Mesoporous Mater. 4445 (2001) 259266. [32] S. Miao, Z. Liu, H. Ma, B. Han, J. Du, Z. Sun, Z. Miao, Synthesis and characterization of mesoporous aluminosilicate molecular sieve from K-Fledspar, Microporous Mesoporous Mater. 83 (2005) 277282. [33] R.S. Araujo, D.C.S. Azevedo, E. Rodriguez-Castellon, A. Jimenez-Lopez, C.L. Cavalcante Jr., Al and Ti-containing mesoporous molecular sieves: synthesis, characterization and redox activity in the anthracene oxidation, J. Mol. Catal. A: Chem. 281 (2008) 154163. [34] O.A. Anunziata, M.L. Martinez, M. Gomez Costa, Characterization and acidic properties of Al-SBA-3 mesoporous material, Mater. Lett. 64 (2010) 545548. [35] M. Gomez-Cazalilla, J.M. Mrida-Robles, A. Gurbani, E. Rodriguez-Castellon, A. Jimenez-Lopez, Characterization and acidic properties of Al-SBA-15 materials prepared by post-synthesis alumination of a low-cost ordered mesoporous silica, J. Solid State Chem. 180 (2007) 11301140. [36] L. Zhang, Y. Zhao, H. Dai, H. He, C.T. Au, A comparative investigation on the properties of Cr-SBA-15 and CrOx /SBA-15, Catal. Today 131 (2008) 4254. [37] S. Lim, G.L. Haller, Preparation of highly ordered vanadium-substituted MCM41: stability and acidic properties, J. Phys. Chem. B 106 (2002) 84378448. ay, The Surface acidity and characterization of Fe-montmorillonite [38] M. Akc probed by in-situ FT-IR spectroscopy of adsorbed pyridine, Appl. Catal. A: Gen. 294 (2005) 156160. [39] K. Tanabe, M. Misono, Y. Ono, H. Hattori, New solid acids and bases, their catalytic properties, Stud. Surf. Sci. Catal. 51 (1989) 525. [40] R.M. Martin-Aranda, J. Cejka, Recent advances in catalysis over mesoporous molecular sieves, Top. Catal. 53 (2010) 141153.

presence of aluminum containing SBA-type catalysts. These catalysts caused a signicant decrease in the activation energy of the reaction from 172 kJ/mol to a value in the range of 5189 kJ/mol depending on the amount of aluminum loading. Among the aluminum impregnated catalysts, the materials synthesized using aluminum sulphate were more effective in polypropylene degradation reaction.

Acknowledgements The authors thank Middle East Technical University Research Fund (BAP-2005-07-02-00-69 and BAP-2007-03-04-08) and The Scientic and Technological Research Council of Turkey (TUBITAK107M303) for the nancial support and Central Laboratory of Middle East Technical University for the analysis. The authors also thank Dr. Lucian Barbu-Tudoran from Babes-Bolyai University (Romania) for Transmission Electron Microscopy analysis and Nalan zbay from Gazi University for the measurement of acid strengths. References
[1] F. Pinto, P. Costa, I. Gulyurtlu, I. Cabrita, Pyrolysis of plastic wastes 2. Effect of catalyst on product yield, J. Anal. Appl. Pyrolysis 51 (1999) 5771. , S.N. Koc , G.S. Pozan, A. Kas gz, Thermal-catalytic degradation kinet[2] A. Durmus ics of polypropylene over BEA, ZSM-5 and MOR zeolites, Appl. Catal. B: Environ. 61 (2005) 316322. [3] Y.H. Lin, H.Y. Yen, Fluidised bed pyrolysis of polypropylene over cracking catalysts for producing hydrocarbons, Polym. Degrad. Stab. 89 (2005) 101 108. [4] K. Gobin, G. Manos, Polymer degradation to fuels over microporous catalysts as a novel tertiary plastic recycling method, Polym. Degrad. Stab. 83 (2004) 267279. [5] Z. Obal, N.A. Sezgi, T. Do gu, Performance of acidic MCM-like aluminosilicate catalysts in pyrolysis of polypropylene, Chem. Eng. Commun. 196 (2009) 116130. [6] W. Kaminsky, I.-J.N. Zorriqueta, Catalytical and thermal pyrolysis of polyolens, J. Anal. Appl. Pyrolysis 79 (2007) 368374. [7] R.A. Garcia, D.P. Serrano, D. Otero, Catalytic cracking of HDPE over hybrid zeolitic-mesoporous materials, J. Anal. Appl. Pyrolysis 74 (2005) 379 386. [8] J. Aguado, D.P. Serrano, G. San Miguel, J.M. Escola, J.M. Rodrigues, Catalytic activity of zeolitic and mesostructured catalysts in the cracking of pure and waste polyolens, J. Anal. Appl. Pyrolysis 78 (2007) 153161. [9] R. Van Grieken, D.P. Serrano, J. Aguado, R. Garcia, C. Rojo, Thermal and catalytic cracking of polyethylene under mild conditions, J. Anal. Appl. Pyrolysis 5859 (2001) 127142. [10] J. Aguado, J.L. Sotelo, D.P. Serrano, J.A. Calles, J.M. Escola, Catalytic conversion of polyolens into liquid fuels over MCM-41: comparison with ZSM-5 and amorphous SiO2 Al2 O3 , Energy Fuels 11 (1997) 12251231. [11] Y. Uemichi, Y. Makino, T. Kanazuka, Degradation of polypropylene to aromatic hydrocarbons over Pt- and Fe-containing activated carbon catalysts, J. Anal. Appl. Pyrolysis 16 (1989) 229238. [12] A. Marcilla, A. Gomez-Siurana, D. Berenguer, Study of the inuence of the characteristics of different acid solids in the catalytic pyrolysis of different polymers, Appl. Catal. A: Gen. 301 (2006) 222231. [13] J.S. Beck, J.C. Vartuli, W.J. Roth, M.E. Leonowicz, C.T. Kresge, K.D. Schmitt, C.T-W. Chu, D.H. Olson, E.W. Sheppard, S.B. McCullen, J.B. Higgins, J.L. Schlenker, A new family of mesoporous molecular sieves prepared with liquid crystal templates, J. Am. Chem. Soc. 114 (1992) 1083410843. bilmez, T. Do gu, S. Balc, Vanadium incorporated high surface area MCM[14] Y. Gc 41 catalysts, Catal. Today 100 (2004) 473477.

Potrebbero piacerti anche