Você está na página 1de 7

Microporous and Mesoporous Materials 164 (2012) 120126

Contents lists available at SciVerse ScienceDirect

Microporous and Mesoporous Materials


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

Improved hydrothermal stability and acidic properties of ordered mesoporous SBA-15 analogs assembled from nanosized ZSM-5 precursors
Xuan Hoan Vu, Norbert Steinfeldt, Udo Armbruster, Andreas Martin
Leibniz-Institut fr Katalyse e.V. an der Universitt Rostock, Albert-Einstein-Strae 29a, D-18059 Rostock, Germany

a r t i c l e

i n f o

a b s t r a c t
Ordered hexagonal mesoporous materials have been successfully synthesized in a two-step process from assembly of pre-formed nanoscale ZSM-5 precursors at elevated temperature by pH adjusting during synthesis. The resulting materials were characterized by small-angle X-ray scattering (SAXS), N2 adsorption and desorption (BET), transmission electron microscopy (TEM), temperature-programmed desorption of ammonia (NH3-TPD), FTIR spectroscopy of pyridine adsorptiondesorption (pyr-FTIR), atomic absorption spectroscopy (AAS) and inductively coupled plasma-atomic emission spectroscopy (ICPAES). It was found out that a highly ordered mesoporous material with both improved hydrothermal stability and acidic properties can be optimally prepared at pH 3.5 and 200 C. The ordered mesostructure of such material is well preserved even after steaming at 800 C for 4 h, with little loss in BET specic surface of 10.7% only, while its acidic properties are signicantly enhanced in terms of both acid strength and total acidity. These improved properties can be attributed to the synergistic advantages of the use of zeolite nanoseeds, high temperature synthesis and pH adjustment. 2012 Elsevier Inc. All rights reserved.

Article history: Available online 5 June 2012 This article is dedicated to Prof. Dr. Jens Weitkamp on the occasion of his 70th birthday. Keywords: Mesoporous material Acidity Preparation ZSM-5 nanoseeds Characterization

1. Introduction Ordered mesoporous materials, typically like MCM-41 [1] and SBA-15 [2], have attracted considerable attention due to their potential application in catalysis and separation. Compared to microporous zeolites, which are widely used as industrial catalysts, mesostructured materials with high surface areas and large, uniform pore sizes might overcome the diffusion limitation encountered in zeolite catalysis, being suitable for the conversion of bulky molecules [3,4]. Unfortunately, these mesoporous materials exhibit relatively weak acidity and poor hydrothermal stability due to the amorphous nature of the mesopore walls, which severely hinder their practical uses [5]. Much effort has been devoted to improve both the poor acidity and low hydrothermal stability of mesoporous molecular sieves over the past two decades [69]. It has been demonstrated that increasing the atomic order of the mesopore walls is an effective strategy to enhance both the acidity and hydrothermal stability [1013]. Kloetstra et al. [10] rst reported the Al-MCM-41 with zeolitic structures in the mesopore walls, generated by partial crystallization of the pre-assembled amorphous walls of Al-MCM-41 in the presence of microporous zeolite structure directors. However, later studies [1113] have shown that this approach only allows achieving low degrees of crystallinity, not higher than 42%. Further
Corresponding author. Fax: +49 381 1281 51246.
E-mail address: andreas.martin@catalysis.de (A. Martin). 1387-1811/$ - see front matter 2012 Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.micromeso.2012.05.036

crystallization induced extra-wall growth of zeolite crystals and nally destroyed the mesoporous framework. To avoid this phenomenon, an alternative route was developed by Pinnavaia et al. [1416] and others [17,18] based on the use of pre-formed zeolite nanoclusters that consist of primary or secondary building units as precursors to assemble the mesostructure. By this way, SBA-15 analogs MAS-7 [19], MAS-9 [20] and MSU-S/H [21] were successfully synthesized in strong acidic media using nanoseeds of zeolite beta, ZSM-5, FAU-Y with triblock copolymer pluronic P123 as structure directing agent. The resulting materials exhibit strong acidity and high hydrothermal stability, which was attributed by the authors to a better local order in the mesopore walls introduced by pre-formed zeolite precursors compared to conventional mesoporous aluminosilicates [1922]. However, it should be noted that the aluminum content in SBA-15 analogs prepared by such a route is relatively low due to the strong acidic media used during their syntheses. Furthermore, most syntheses were carried out at mild temperatures (typically 100150 C) to prevent the decomposition of organic templates in strong acidic media, which resulted in imperfectly condensed mesopore walls. Consequently, the acidity and hydrothermal stability of these mesoporous materials are still unsatisfactory for industrial applications [23]. Herein, in order to enhance the acidity and hydrothermal stability of the ordered mesoporous materials assembled from ZSM-5 nanoseeds, we aim at increasing i) the silica condensation degree and ii) incorporated aluminum content in the mesopore walls by pH adjustment before hydrothermal treatment at high tempera-

X.H. Vu et al. / Microporous and Mesoporous Materials 164 (2012) 120126

121

ture. The obtained materials exhibit signicant improvements in both acidity and hydrothermal stability while their ordered mesostructures are well preserved. 2. Experimental 2.1. Chemicals The chemicals used in the synthesis were tetraethyl orthosilicate (TEOS, 99%, Aldrich), tetrapropylammonium hydroxide (TPAOH, 20% in water, Aldrich), aluminum isopropoxide (AIP, 98%, Aldrich), triblock copolymer pluronic P123 (EO20PO70EO20, MW = 5800, Aldrich), hydrochloric acid (HCl, 37%, J. T. Baker) and ammonium hydroxide (NH4OH, 25%, Acros). All agents were used as received without further purication. 2.2. Synthesis There are two steps involved in the synthesis of SBA-15 analogs assembled from ZSM-5 nanoseeds. The pre-formed ZSM-5 precursor solution was rst prepared, and then added to the surfactant solution to form the mesoporous framework in strong acidic media. After the mesostructure has basically formed, the pH value of the system was adjusted to higher values, followed by hydrothermal treatment at higher temperature for further silica condensation and more incorporation of aluminum. In the typical synthesis, 6.0 g of TEOS, 10.0 g of TPAOH, 2.0 g of distilled H2O, and 0.19 g of AIP were mixed at room temperature and stirred overnight to complete the hydrolysis, followed by hydrothermal treatment at 90 C for 6 h in a reux system to obtain the ZSM-5 precursor solution with the nal molar composition of 60 SiO2:1 Al2O3:20 TPAOH:1200 H2O:240 EtOH. Simultaneously, the P123 solution was prepared by dissolving 2.0 g of P123 in 75 ml of 1.6 M HCl at room temperature for 4 h to get the clear solution. Then, the ZSM-5 precursor solution prepared as described above was added dropwise to the P123 solution, followed by aging at 40 C for 24 h. Before transferring the mixture into a Teon-lined autoclave for further condensation at 200 C for 24 h, the pH value was adjusted in steps from 1.5 to 5.5 with aqueous NH3 solution to receive different solid samples. The nal product was ltered off, washed with distilled water, and dried at 100 C for 12 h. The assynthesized materials were calcined in air at 550 C for 5 h with a heating rate of 2 K/min to remove the organic template. The calcined samples are denoted as SAZ_x, where SAZ stands for SBA-15 analogs assembled from ZSM-5 nanoseeds, x is the pH value used in the hydrothermal treatment of the second synthesis step. For comparison, Al-SBA-15 was synthesized by the same procedure as SAZ_3.5, but using the conventional silica based sources instead of ZSM-5 nanoseeds. SBA-15 and H-ZSM-5 were prepared according to the synthesis procedures reported by Zhao et al. [2] and Kulkarni et al. [24] respectively. 2.3. Characterization SAXS measurements were carried out using a Kratky-type instrument (SAXSess, Anton Paar, Austria) operated at 40 kV and 50 mA in slit collimation using a two-dimensional CCD detector (T = 40 C). The 2D scattering pattern was converted into a onedimensional scattering curve as a function of the magnitude of the scattering vector q = (4/k)sin(h/2) with SAXSQuant Software (Anton Paar). A Gbel mirror was used to convert the divergent polychromatic X-ray beam into a collimated line-shaped beam of Cu K radiation (k = 0.154 nm). Slit collimation of the primary beam was applied in order to increase the ux and to improve the signal quality. The sample cell consisted of a metal body with two windows for the X-ray beam. The powdered samples were sealed be-

tween two layers of Scotch tape. Scattering proles of the mesoporous materials were obtained by subtraction of the detector current background and the scattering pattern of the Scotch tape from the experimental scattering patterns. Correction of instrumental broadening effects (smearing) was carried out with SAXSQuant software using the slit length prole determined in a separate experiment. Nitrogen physisorption measurements were carried out at 196 C on an ASAP 2010 Micromeritics apparatus. Before measurements, the samples were degassed at 180 C in vacuum for 10 h. The BET specic surface area was calculated using adsorption data at a relative pressure (p/po) of 0.050.25, and the total pore volume was estimated from the amount adsorbed at a relative pressure of about 0.976. The pore size distributions were obtained from the desorption branch of the isotherm using the corrected form of the Kelvin equation by means of the Barrett-Joyner-Halenda method with a cylindrical pore model. Temperature-programmed desorption of ammonia (NH3-TPD) measurements were carried out in a quartz tube reactor in the range of 100550 C. The samples were rst activated at 550 C for 0.5 h under helium ow. After the reactor cooled down to 100 C, the samples were swept with ammonia for 0.5 h for adsorption, then the feed gas was switched to helium to remove the physically adsorbed ammonia molecules, until the baseline was at. After that, the temperature was increased to 550 C with a heating rate of 10 K/min, with the desorbed ammonia being continuously detected by a thermal conductivity detector (TCD, Gow-Mac Instruments Co.) and quantitatively analyzed by the external standard method using the desorption peak area integrated to the standard curve. Infrared spectra of adsorbed pyridine (pyr-FTIR) were recorded on a Tensor 27 FT-IR spectrometer (Bruker). The sample was evacuated at 400 C for 2 h and cooled to room temperature. Then the sample was exposed to pyridine containing helium gas at room temperature for 1 h, and then ushed with helium to remove physisorbed pyridine. After that, the temperature was increased gradually up to 400 C while the infrared spectra were recorded. The Al and Si contents of the calcined samples were determined by ICP-AES (715-ES, Varian) and AAS (Analyst 300, Perkin Elmer) respectively. For this purpose, the samples were treated with a mixture of HClHNO3HF in order to dissolve them completely. The TEM measurements were performed at 200 kV on a JEMARM200F (JEOL) which is aberration-corrected by a CESCOR (CEOS) for the scanning transmission microscopic (STEM) applications. The sample was deposed on a holey carbon supported Cu-grid (mesh 300) as received and transferred to the microscope. 2.4. Hydrothermal stability tests The hydrothermal stability of calcined samples was evaluated by steaming at 800 C with 30% water vapor in helium ow (30 mL/min) for 4 h. 0.2 g of samples were sieved with the particle size between 300700 lm, and then loaded into the tube reactor. The samples were heated to 800 C under helium ow with a rate of 20 K/min. As soon as the temperature reached 800 C, the helium ow containing 30 vol.% steam was introduced into the tube reactor. After 4 h treatment, the samples were cooled down and characterized by SAXS and nitrogen physisorption to re-evaluate the structural and textural properties. 3. Results and discussion 3.1. Particle size distribution of the pre-formed ZSM-5 precursor solution Several studies have shown that the particle size of zeolite nanoseeds has a great effect on the ordering of the formed meso-

122

X.H. Vu et al. / Microporous and Mesoporous Materials 164 (2012) 120126

structures [25,26]. To understand such an effect in the present synthesis, the particle size of the seeds in the precursor solution was determined by SAXS. Fig. 1 shows the scattering pattern with vector q of the ZSM-5 precursor solution after aging at 90 C for 6 h. The pattern reects scattering from two populations. The scattering intensity observed between 0.5 nm1 < q < 6 nm1 is attributed to the presence of small primary ZSM-5 units. At low q values, a steeply increasing intensity in scattering is observed, which reects the presence of larger ZSM-5 units. Guinier analysis [27] was applied to obtain information about the particle size of the different populations. For the primary ZSM-5 units, a Guinier radius (Rg) of 2.4 nm was determined. The Rg for the larger structures amounted to 14.4 nm. The parameter Rg has a model dependent relation to the size of the particles. Assuming spherical particle shape, its radius RG can be calculated using the following equation: Rg2 = (3/5) RG2. From the presented data, it is reasonable to conclude that the ZSM-5 precursor solution contains the primary units and their aggregates, which are presumed to promote the zeolite nucleation to form ZSM-5 crystals. These results are consistent with previous reports [2830]. 3.2. Structural and textural characterization It has been reported that the acidity of the synthesis solution is a decisive factor for not only the mesoporous regularity but also the successful incorporation of aluminum into the SBA-15 framework [3133]. In this work, therefore, such effects of hydrothermal treatment pH (htt-pH) in the preparation of SBA-15 analogs from ZSM-5 nanoseeds at high temperature were investigated systematically. The structural and textural properties of the resulting materials were characterized by SAXS, TEM, and nitrogen adsorption/desorption. The results are shown in Fig. 2, Fig. 3, Fig. 4 and Table 1. From the SAXS patterns (Fig. 2) it can be seen that the samples prepared at the lower pH values in the second synthesis step (SAZ_1.5, SAZ_2.5, SAZ_3.5) exhibit three well-resolved reections indexed as the 100, 110 and 200 reections of 2D hexagonal (p6mm) symmetry, indicating an ordered mesostructure. Upon increasing the htt-pH value to 4.5 (SAZ_4.5) or 5.5 (SAZ_5.5), the 110 and 200 reections become less pronounced, indicating the formation of a disordered mesostructure. These results demonstrate that the htt-pH has a signicant effect on the ordering of mesostructures, which is in accord with the report by Pan et al. [31] when

100

110 200

log intensity / a.u.

(a) (b) (c) (d) (e)

0.5

1.0

1.5

2.0

2.5

2 Theta/degree
Fig. 2. SAXS patterns with 2h of (a) SAZ_1.5; (b) SAZ_2.5; (c) SAZ_3.5; (d) SAZ_4.5; (e) SAZ_5.5.

A
(a) Volume adsorbed (cm /g)
3

(b)

(c) (d) (e)

0.0

0.2

0.4

0.6

0.8

1.0

Relative pressure (P/Po)

B
(a)

Rg = 14.4 nm

Pore volume (cm3/g)

(b)

(c)

log intensity / a.u.

Rg = 2.4 nm

(d)

(e)

12

16

20

Pore diameter (nm)


Fig. 3. N2 sorption isotherms (A) and the corresponding pore size distribution curves (B) of (a) SAZ_1.5; (b) SAZ_2.5; (c) SAZ_3.5; (d) SAZ_4.5; (e) SAZ_5.5.
0.1 1

q/nm

-1

Fig. 1. SAXS pattern with vector q of ZSM-5 precursor solution at 90 C for 6 h.

using the conventional silica precursors. On the other hand, it is worthy to note that the reection positions shift only very slightly over the whole range of the htt-pH value used, suggesting that the

X.H. Vu et al. / Microporous and Mesoporous Materials 164 (2012) 120126

123

Fig. 4. TEM images of SAZ_3.5 viewing along [100] (A) and [110] (B) directions.

Table 1 Structural and textural properties of SAZ_x (x = 1.55.5). Sample SAZ_1.5 SAZ_2.5 SAZ_3.5 SAZ_4.5 SAZ_5.5 D100 (nm) 11.13 11.14 11.18 11.15 11.13 a0 (nm) 12.85 12.86 12.91 12.87 12.85 Dp (nm) 9.0 9.1 9.1 9.1 W (nm) 3.85 3.76 3.80 3.77 SBET (m2/g) 424 412 372 297 184 Vt (cm3/g) 1.05 1.10 0.97 0.80 0.77

a0: unit cell parameter (a0 = 2 d100/31/2); Dp: pore diameter; W: pore wall thickness (W = a0 Dp); Vt: total pore volume.

unit cell of the SAZ_x samples seems to be not affected by the httpH. The data of N2 adsorption/desorption are also in accordance with those of SAXS. The samples SAZ_x (x = 1.53.5) show differing textural properties compared to the two other ones (x = 4.55.5) (Fig. 3, Table 1). As depicted in Fig. 3A, all SAZ_x samples exhibit type IV isotherms with H1-type hysteresis loop, conrming their mesoporous nature. However, the capillary condensation steps of SAZ_1.5, SAZ_2.5 and SAZ_3.5 are very steep and occur in the relative pressure (p/po) range of 0.70.9, which is characteristic for an ordered mesoporous material with a large and uniform pore size. The two other samples (SAZ_4.5, SAZ_5.5) display a less steep hysteresis loop being indicative for a disordered mesostructure. The dimension and uniformity of mesopores can be directly reected by the pore size distribution (Fig. 3B). With the increase of pH values used in the hydrothermal treatment, the pore size distribution of SAZ_x becomes broader. The pore size of the rst four samples (x = 1.54.5) is about 9.09.1 nm, larger than that of MAS-9 [20] prepared at a lower temperature (100 C), implying that the silica based walls of these samples have been further condensed under the hydrothermal treatment at high temperature, being similar to the previous reports [31,34,35]. For SAZ_5.5, the mesopore system has been partially destroyed; therefore, its pore size cannot be calculated. The disordered mesostructure of SAZ_4.5 and SAZ_5.5 is also evidenced by their relatively low surface areas and pore volumes compared to those of the three other samples. As summarized in Table 1, SAZ_1.5, SAZ_2.5 and SAZ_3.5 show BET specic surface areas ranging from 424 to 372 m2/g and pore volumes from 1.10 to 0.97 cm3/g, while these textural parameters reduce to 279 m2/g and 0.80 cm3/g for SAZ_4.5 and 184 m2/g and 0.77 ce:hsp sp="0.25"/>cm3/g for SAZ_5.5. Hence, in order to obtain ordered mesoporous materials, the htt-pH value should not be higher than 3.5. According to Pan et al. [31], at higher pH values the surfactantsilanol interaction becomes weak, leading to the decline of the surfactant content adsorbed within the pores. As a result, the mesostructure partially collapsed while the framework shrinkage occurred at elevated temperature (200 C).

The ordered mesostructure of SAZ_3.5 is further conrmed by TEM characterization. The TEM images (Fig. 4A, B) exhibit well-ordered hexagonal arrays of mesopores with one-dimensional channels taken in [100] and [110] directions. From the bright-dark contrast in the TEM image (Fig. 4A), the mesopore size and wall thickness can be estimated to be 8.6 nm and 3.7 nm, respectively, which are comparable to the data of SAXS and N2 adsorption/ desorption. It was reported [33] that many void defects were formed in the mesoporous materials assembled from zeolite nanoclusters due to their relatively strong rigidity and large volume. However, in this case, it seems that no obvious void defects have been found from TEM analysis of SAZ_3.5, suggesting that the nanoseeds of ZSM-5 precursor solutions can be accommodated within pore walls of the dimensions observed. These results indicate that most large nanoclusters, present in the synthesis mixture, the size of which was evaluated by SAXS (see Fig. 1), have been separated into smaller sized units under the applied synthesis condition, which are suitable for assembly of the mesostructure with a wall thickness of 3.7 nm. Therefore, by using the proper zeolite precursor solution, the formation of separated defects in the synthesis of SBA-15 analogs at high temperature can be effectively suppressed. 3.3. Acidity study The inuence of the htt-pH on the total acidity of SAZ_x materials was studied by NH3-TPD. The data are presented in Fig. 5 and Table 2. It can be seen that the total acidity (Table 2) enhances considerably with increasing of the pH value. The total acidity of SAZ_5.5 (0.35 mmol NH3/g) is two times higher than that of SAZ_1.5 (0.16 mmol NH3/g). These results can be attributed to larger amounts of grafted aluminum at higher pH values, as conrmed by elemental analysis, which gave rise to more acid sites in the products. Notably, in the htt-pH range of 3.55.5, almost all aluminum added to the initial solution of zeolite seeds has effectively been introduced into the nal products. SAZ_3.5, for example, has a Si/Al ratio in the product of 34, which is very close to the Si/Al ratio = 30 in the initial gel. In contrast, MAS-9 prepared in strong acidic media showed a relatively low aluminum content; for example, the Si/Al ratio in the initial gel of 60 resulted in a product with Si/Al = 256 [36]. It is generally accepted that the aluminum incorporation into the nal product is favored in less acidic environment. At lower pH values, the AlOSi bonds in the ZSM-5 precursors might get hydrolyzed, leading to the leaching of Al during the hydrothermal treatment. Consequently, the total acidity decreases in the samples prepared at lower pH values. On the other hand, it was shown that the increase of pH values over 3.5 brought about less ordering of mesostructures. Therefore, SAZ_3.5 was chosen for further studies.

124

X.H. Vu et al. / Microporous and Mesoporous Materials 164 (2012) 120126

(a) Intensity / a.u. Intensity / a.u.

(b)

(a) (b) (c) (d)

(c) (d) (e)

100 100 200 300 400 500 600

200

300

400

500

600

Temperature /C
Fig. 5. NH3-TPD proles of (a) SAZ_1.5; (b) SAZ_2.5; (c) SAZ_3.5; (d) SAZ_4.5; (e) SAZ_5.5.

Temperature / C
Fig. 6. NH3-TPD prole of (b) SAZ_3.5 compared to those of (a) HZSM-5, (c) Al-SBA15, and (d) SBA-15.

Table 2 Acidity of SAZ_x (x = 1.55.5), Al-SBA-15, H-ZSM-5 and SBA-15. Sample SAZ_1.5 SAZ_2.5 SAZ_3.5 SAZ_4.5 SAZ_5.5 Al-SBA-15a HZSM-5b SBA-15c Si/Al ratio in gel 30 30 30 30 30 30 40 1 Si/Al ratio in productsd 67 40 34 32 29 45 n.a. 1 Total acidity (mmol NH3/g) 0.16 0.22 0.26 0.27 0.35 0.18 1.04

a Synthesized by the same procedure as SAZ_3.5, but using the conventional silica based source. b Synthesized according to the patent of Kulkarni et al. [24]. c Prepared by the original recipe of Zhao et al. [2]. d Si, Al content in the products analyzed by AAS and ICP, respectively.

Fig. 7. FTIR spectra of pyridine desorption of SAZ_3.5 at various temperatures (L-Py Lewis sites, PyH+ - Bronsted sites).

From the point of view of acid strength, Fig. 5 shows that all samples exhibit two ammonia desorption peaks. The rst sharp peak around 200 C can be attributed to weak acid sites, while the broader peak in the range of 350450 C shows the presence of medium and strong acid sites. However, it seems that the acid strength of SAZ_x samples is dominated by weak and medium sites and still lower than that of HZSM-5 crystals, but higher than that of AlSBA-15 prepared from conventional silica based precursors under the same conditions in terms of both acid strength and total acidity. The pure SBA-15, as expected, exhibits signicantly less response, which is indicative of little or no acid sites present (Table 2, Fig. 6). In addition, the nature of the acid sites in the representative sample SAZ_3.5 was evaluated by pyridine adsorption/desorption. The pyr-FTIR spectra are shown in Fig. 7. After evacuation at various temperatures (100400 C), the spectra of pyridine desorption show two bands at 1543 cm1 and 1451 cm1 which are characteristic for Brnsted and Lewis acid sites, respectively. Furthermore, these two bands are still present at 400 C, indicating the presence of both strong Brnsted and Lewis acid sites in the solid sample. The nature and mechanism of formation of the Lewis sites are not fully clear, but these kinds of acidity could be generated from either extra framework aluminum species, such as AlO5 and AlO6, or defect sites such as AlO3 or both [37]. Notably, while the strong Lewis acid sites are often observed in zeolites and mesoporous aluminosilicates, the strong Brnsted acid sites are only typical for zeolites [13,36,38]. The presence of strong Brnsted acid sites in SAZ_3.5 can be

presumed to be originated from the zeolite primary and secondary structure units in the wall provided by the ZSM-5 precursors, which is consistent with the previous work [17]. Therefore, by changing the htt-pH, the acidity behavior of SAZ_3.5 shows signicant improvement in terms of acid strength and total acidity that can be attributed to the synergistic advantages of both the high incorporated aluminum content and the use of zeolite nanoseeds. 3.4. Hydrothermal stability The hydrothermal stability was evaluated by treatment of the typical sample SAZ_3.5 as mentioned above (T = 800 C, 30% steam in He ow (30 mL/min)); subsequently SAXS and N2 sorption analysis were carried out. SBA-15 was also tested under the same conditions for comparison. The results are shown in Fig. 8, Fig. 9, Table 3. It is clear that SAZ_3.5 has much better hydrothermal stability than SBA-15. After the steam treatment for 4 h, SAZ_3.5 still exhibits three well-resolved peaks assigned to the 100, 110 and 200 reections of highly ordered hexagonal symmetry though the reection positions shift to higher angles, suggesting the contraction of the aluminosilicate framework (Fig. 8A). In contrast to SAZ_3.5, the 110 and 200 reections of SBA-15 become less obvious, indicating partial destruction of the hexagonal mesostructure (Fig. 8B) and the transition to a disordered structure. These results

X.H. Vu et al. / Microporous and Mesoporous Materials 164 (2012) 120126


100

125

A
log intensity / a.u.

100

B
log intensity / a.u.

100 110 200

110 200

(a)

(a)

110 200 (b)

(b)

0.5

1.0

1.5

2.0

2.5

0.5

1.0

1.5

2.0

2.5

2 Theta/degree

2 Theta/degree

Fig. 8. SAXS patterns of SAZ_3.5 (A) and SBA-15 (B): (a) before and (b) after steam treatment at 800 C for 4 h.

demonstrate that SAZ_3.5 possesses remarkable hydrothermal stability after such treatment. This conclusion is further supported by N2 adsorption and desorption (Fig. 9, Table 3). Only little change can be observed in the N2 adsorption and desorption isotherms of SAZ_3.5 before and after the steam treatment, implying that its ordered mesostructure is well maintained. Moreover, the BET surface area reduces only by around 10% after the steam treatment. In comparison, SBA-15 submitted to the same treatment shows a signicant decrease in BET surface area of 65%. It was reported that many factors, such as zeolite-like connectivity [15], high aging temperature [31], the salt effect [39], disordered mesostructures [40] and the thick pore wall [41] are favorable for the hydrothermal stability of mesoporous materials. Considering these factors in the present synthesis procedure, it is conceivable that the high hydrothermal stability of SAZ_3.5 might originate from the advantages of the use of both zeolite nanoseeds and high temperature synthesis.

Table 3 Textural properties of SAZ_3.5 and SBA-15 before and after steaming at 800 C for 4 h. Sample SAZ_3.5 Calcined Steamed SBA-15a Calcined Steamed
a

Dp (nm) 9.1 9.1 6.4 4.6

Vt (cm3/g) 0.97 0.82 1.06 0.43

SBET (m2/g) 372 332 879 308

Area decrease (%)

10.7

64.9

Synthesized according to the original recipe of Zhao et al. [2].

4. Conclusions Ordered hexagonal mesoporous materials with greatly improved hydrothermal stability and total acidity have successfully been synthesized by assembly of pre-formed ZSM-5 precursors at high temperature (200 C) by pH adjusting method. The optimal hydrothermal treatment pH was found to be 3.5 under the studied conditions. It was shown that the material prepared at pH 3.5 (SAZ_3.5) possesses a highly ordered mesostructure with remarkable hydrothermal stability and increased acidity. The ordered mesostructure of SAZ_3.5 is well maintained after steam treatment

at 800 C for 4 h with only 11% reduction in BET specic surface area. Furthermore, its acidity is signicantly enhanced in terms of both acid strength and total acidity as evidenced by the results of pyr-FTIR and NH3-TPD measurements. Such synthesis techniques offer new possibilities for the preparation of hydrothermally stable mesoporous materials with increased acidity. In addition, such solids will be checked by next in acid catalyzed reactions.

Acknowledgments The authors would like to thank Dr. U. Bentrup for pyr-FTIR measurements, Dr. M.-M. Pohl for TEM images; Mr. R. Eckelt for N2 adsorption and desorption studies. Dr. Hoang is also acknowledged for his help to access NH3-TPD measurement. X.-H. Vu thanks the government of Vietnam and LIKAT for the nancial support.

800 700

A
Volume absorbed (cm3/g)

700 600 500 400 300 200 100 0 0.0

Volume adsorbed (cm3/g)

600 500 400 300 200 100 0 0.0

(a)

(a) (b)
0.2 0.4 0.6 0.8 1.0

(b)

0.2

0.4

0.6

0.8

1.0

Relative pressure (P/Po)

Relative pressure (P/Po)

Fig. 9. N2 sorption isotherms of SAZ_3.5 (A) and SBA-15 (B): (a) before and (b) after steam treatment at 800 C for 4 h.

126

X.H. Vu et al. / Microporous and Mesoporous Materials 164 (2012) 120126 [24] S. J. Kulkarni, S. Puvuuri, N. Nama, K. V. Raghawan, U.S. Pat. 6800,272 B2 (2004). [25] C.S. Carr, S. Kaskel, D.F. Shantz, Chem. Mater. 16 (2004) 31393146. [26] P. Li, L. Liu, G. Xiong, Phys. Chem. Chem. Phys. 13 (2011) 1124811253. [27] A. Guinier, G. Fournet, Small-Angle Scattering of X-rays, Wiley, New York, 1955. [28] T. Lubomira, V.P. Valtchev, Chem. Mater. 17 (2005) 24942513. [29] P.P.E.A. de Moor, T.P.M. Beelen, R.A. van Santen, J. Phys. Chem. B 103 (2001) 16391650. [30] C. Cheng, G. Juttu, S.F. Mitchell, D.F. Shantz, J. Phys. Chem. B 110 (2006) 22488 22495. [31] S. Wu, Y. Han, Y. Zou, J. Song, L. Zhao, Y. Di, S. Liu, F.-S. Xiao, Chem. Mater. 16 (2004) 486492. [32] D. Pan, P. Yuan, L. Zhao, N. Liu, L. Zhou, G. Wei, J. Zhang, Y. Ling, Y. Fan, B. Wei, H. Liu, C. Yu, X. Bao, Chem. Mater. 21 (2009) 54135425. [33] T. Jiang, H. Tao, J. Ren, X. Liu, Y. Wang, G. Lu, Microporous Mesoporous Mater. 142 (2011) 341346. [34] D. Li, D. Su, J. Song, X. Guan, Klein Hofmann, F.-S. Xiao, J. Mater. Chem. 15 (2005) 50635069. [35] N. Xiao, L. Wang, S. Liu, Y. Zou, C. Wang, Y. Ji, J. Song, F. Li, X. Meng, F.-S. Xiao, J. Mater. Chem. 19 (2009) 661665. [36] Y. Han, N. Li, L. Zhao, D. Li, X. Xu, S. Wu, Y. Di, C. Li, Y. Zou, Y. Yu, F.-S. Xiao, J. Phys. Chem. B 107 (2003) 75517556. [37] B. Dragoi, E. Dumitriu, C. Guimon, A. Auroux, Microporous Mesoporous Mater. 121 (2009) 717. [38] V.-T. Hoang, Q. Huang, A. Ungureanu, M. Eic, T.-O. Do, S. Kaliaguine, Langmuir 22 (2006) 47774786. [39] R. Ryoo, J.M. Kim, C.H. Ko, C.H. Shin, J. Phys. Chem. 100 (1996) 1771817721. [40] R. Ryoo, S. Jun, J. Phys. Chem. B 101 (1997) 317320. [41] F. Zhang, Y. Yan, H. Yang, Y. Meng, Y. Meng, C. Yu, B. Tu, D. Zhao, J. Phys. Chem. B 109 (2005) 87238732.

References
[1] 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, J. Am. Chem. Soc. 114 (1992) 1083410843. [2] D. Zhao, J. Feng, Q. Huo, N. Melosh, G.H. Fredrickson, B.F. Chmelka, G.D. Stucky, Science 279 (1998) 548552. [3] A. Corma, Chem. Rev. 97 (1997) 23732419. [4] Y. Wan, D. Zhao, Chem. Rev. 107 (2007) 28212858. [5] R. Chal, C. Grardin, M. Bulut, S. van Donk, ChemCatChem 3 (2011) 6781. [6] Y. Liu, T.J. Pinnavaia, J. Mater. Chem. 12 (2002) 31793190. ejka, S. Mintova, Catal. Rev. 49 (2007) 457509. [7] J. C [8] K. Egeblad, C.H. Christensen, M. Kustova, C.H. Christensen, Chem. Mater. 20 (2008) 946960. [9] F. Chen, X. Meng, F. Xiao, Catal. Survey Asia 15 (2011) 3748. [10] K.R. Kloetstra, H. van Bekkum, J.C. Jansen, Chem. Commun. (1997) 22812282. [11] A.A. Campos, L. Dimitrov, C.R. da Silva, M. Wallau, E.A. Urquieta-Gonzlez, Microporous Mesoporous Mater. 95 (2006) 92103. [12] M.J. Verhoef, P.J. Kooyman, J.C. van der Waal, M.S. Rigutto, J.A. Peters, H. van Bekkum, Chem. Mater. 13 (2001) 683687. [13] D.T. On, S. Kaliaguine, Angew. Chem., Int. Ed. 40 (2001) 32483251. [14] Y. Liu, W. Zhang, T.J. Pinnavaia, J. Am. Chem. Soc. 122 (2000) 87918792. [15] Y. Liu, W. Zhang, T.J. Pinnavaia, Angew. Chem., Int. Ed. 40 (2001) 12551258. [16] Y. Liu, T.J. Pinnavaia, J. Mater. Chem. 14 (2004) 10991103. [17] Z. Zhang, Y. Han, F. Xiao, S. Qiu, L. Zhu, R. Wang, Y. Yu, Z. Zhang, B. Zou, Y. Wang, H. Sun, D. Zhao, Y. Wei, J. Am. Chem. Soc. 123 (2001) 50145021. [18] C.W. Ingram, Y. Ghirmazion, T. Mehreteab, J. Porous Mater. 14 (2007) 717. [19] Y. Han, F.-S. Xiao, W. Wu, Y. Sun, X. Meng, D. Li, S. Lin, F. Deng, X. Ai, J. Phys. Chem. B 105 (2001) 79637966. [20] Y. Han, S. Wu, Y. Sun, D. Li, F.-S. Xiao, Chem. Mater. 14 (2002) 11441148. [21] Y. Liu, T.J. Pinnavaia, Chem. Mater. 14 (2002) 35. [22] F.-S. Xiao, Top. Catal. 35 (2005) 924. [23] X. Meng, F. Nawaz, F.-S. Xiao, Nano Today 4 (2009) 292301.

Você também pode gostar