Você está na página 1de 5

Microporous and Mesoporous Materials 213 (2015) 192e196

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

Short communication

Microwave synthesis of AFI-type aluminophosphate molecular sieve


under solvent-free conditions
Xinhong Zhao a, *, Jiangbo Zhao a, Juanjuan Wen a, An Li a, Guixian Li a, Xiaolai Wang b
a

School of Petrochemical Engineering, Lanzhou University of Technology, Lanzhou 730050, China


State Key Laboratory for Oxo Synthesis and Selective Oxidation, Lanzhou Institute of Chemical Physics, Chinese Academy of Sciences, Lanzhou 730000,
China
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 13 February 2015
Received in revised form
18 March 2015
Accepted 29 March 2015
Available online 3 April 2015

AlPO4-5 aluminophosphate molecular sieve with AFI topology has been successfully synthesized by
microwave irradiation under solvent-free conditions. The key inuential factors controlling the crystallization of AFI structure were thoroughly investigated. The optimum synthetic conditions of this
material are as follows: the initial composition is 1.0Al2O3: 3.0P2O5: 2.0HF: 8.0TEABr(tetraethylammonium bromide): 125C, aluminum isopropoxide is the best aluminum source, and activated
carbon needs to be pretreated before used as the reaction medium and hard template. The resultant
AlPO4-5 molecular sieves were characterized by X-ray diffraction (XRD), N2 physisorption, scanning
electron microscopy (SEM) and transmission electron microscopy (TEM). The results reveal that the
obtained samples exhibit hierarchically porous characteristic.
2015 Elsevier Inc. All rights reserved.

Keywords:
AlPO4-5
Solvent-free
Hierarchical porous
Microwave

1. Introduction
AFI-type molecular sieve, composed of straight 12 MR onedimensional channels along the c crystallographic direction, is a
prominent one among the aluminophosphate molecular sieve
family. Recently, AFI molecular sieves have received much attention
due to their extensive use in catalysis, separation and nonlinear
optics [1]. Generally, aluminophosphate molecular sieves are synthesized by conventional hydrothermal or solvothermal method
with water or organic solvents as the reaction medium. The use of
these solvents inevitably results in many drawbacks such as waste
generation, low synthesis efciency and high autogenous pressure
[2]. Presently, considerable efforts have been devoted to solve these
problems.
Ionothermal synthesis, reported by Morris' group [3,4], can
eliminate safety concerns associated with high hydrothermal
pressure because it can be carried out under ambient pressure. But
this synthetic method involves the use of expensive ionic liquid,
and its recycling is not convenient [5]. Most recently, Xiao and coworkers developed a novel solvent-free synthesis, which exhibits
many advantages, such as the low waste production and high

zeolite yield. Additionally, similar to ionothermal synthesis, the


solvent-free synthesis also eliminates high pressure [2,6e9].
However, this route is energy intensive due to employing conventional heating.
As an efcient and selective heating method, microwave irradiation has been used to the synthesis of many new materials
[10e12]. In our previous work, FeAPO-16 [13], FeAPO-5 [14] and
SAPO-5 [15] molecular sieves were ionothermally synthesized by
microwave heating with cheap deep eutectic mixture as both solvent and template. This route is an energy-saving one compared
with conventional synthesis. Unfortunately, it is still not environmental benign owing to the heavy use of eutectic mixture. Inspired
by Xiao and co-workers results and considering the versatility of
hierarchically structured aluminophosphate molecular sieves, in
the present work we aim to develop a novel method of synthesizing
AFI molecular sieve with micro- and meso-porous structure, in
which activated carbon replaces the eutectic mixture solvent as the
reaction medium and hard template.
2. Experimental section
2.1. Synthesis of AlPO4-5

* Corresponding author. Tel.: 86 931 7823126.


E-mail address: Licpzhaoxh@lut.cn (X. Zhao).
http://dx.doi.org/10.1016/j.micromeso.2015.03.031
1387-1811/ 2015 Elsevier Inc. All rights reserved.

For the synthesis of AlPO4-5 molecular sieve, activated carbon


(Sinopharm Chemical Reagent Co. Ltd) was pretreated by nitric acid

X. Zhao et al. / Microporous and Mesoporous Materials 213 (2015) 192e196

(65 wt%, Yantai Shuangshuang Co. Ltd) before used as the reaction
medium. All other reagents were of reagent grade and used without
further purications: tetraethylammonium bromide (TEABr, Sinopharm Chemical Reagent Co. Ltd), aluminum hydroxyacetate
(Sinopharm Chemical Reagent Co. Ltd), aluminum isopropoxide
(Sinopharm Chemical Reagent Co. Ltd), pseudo-boehmite (Shandong Zibo Senchi Chemical Co., Ltd), phosphoric acid (85 wt% in
water, Tianjin De'en Chemical Reagent Co. Ltd), hydrouoric acid
(40 wt% in water, Sinopharm Chemical Reagent Co. Ltd), acetone
(Beijing Chemical Plant) and deionized water.
General synthesis procedure of AlPO4-5 molecular sieve was as
follows: the raw or pretreated activated carbon, TEABr, aluminum
source, phosphoric acid, hydrouoric acid and additional deionized
water if required were measured out in the molar ratio of 1.0Al2O3:
(2e6)P2O5: (1e4)HF: (4e8)TEABr: (0e20)H2O: (62.5e125)C and
ground in a mortar or ball mill (XQM-04L, speed of rotation
300 rpm, Nanjing Kexi Institute of Experimental Instruments Co.,
China) for 20 min. The resulting solid mixture was transferred to a
round-bottomed ask equipped with a condenser and placed in a
microwave reaction system (model XH-MC-1 from Beijing Xianghu
Science and Technology Development Co., Ltd, Beijing, China). Then
the mixture was heated to a designated temperature at about 20  C/
min and kept there for appropriate time (1e2 h). After crystallization, the synthesis mixture was cooled to room temperature. The
resultant product was washed thoroughly with acetone and
distilled water, dried in air, and calcined at 550  C for 5 h to remove
the template. Detailed synthesis conditions were listed in Table 1.
2.2. Characterization
X-ray powder diffraction (XRD) patterns of the synthesized
materials were conducted on a D/Max-2400 Rigaku diffractometer
with Cu Ka radiation operated at 40 kV and 150 mA. Nitrogen
adsorption/desorption studies were conducted on a Micromeritics
ASAP 2020 surface area and pore size analyzer at 196  C. Samples
were outgassed at 200  C for 4 h prior to measurements. Specic
surface areas of materials were calculated from the adsorption data
obtained at p/p0 between 0.07 and 0.20, using the BrunauereEmmetteTeller (BET) equation. The micropore volumes

Table 1
Initial compositions and crystallization conditions for the synthesis of AFI.
Sample

Al2O3:P2O5:HF:TEABr:H2Oa:C
(molar ratio)

Temperature
( C)

Time (h)

P-2
P-3
P-6
C-0
C-62.5
C-raw
TEABr-4
HF-1
HF-4
H2O-0
H2O-20
Al-IP
Al-PB
T180-2h
T200-1h
T200-2h
Mill-20e

1.0:2.0:2.0:8.0:10.0:125
1.0:3.0:2.0:8.0:10.0:125
1.0:6.0:2.0:8.0:10.0:125
1.0:3.0:2.0:8.0:10.0:0.0
1.0:3.0:2.0:8.0:10.0:62.5
1.0:3.0:2.0:8.0:10.0:125b
1.0:3.0:2.0:4.0:10.0: 125
1.0:3.0:1.0:8.0:10.0:125
1.0:3.0:4.0:8.0:10.0:125
1.0:3.0:2.0:8.0:0.0:125
1.0:3.0:2.0:8.0:20.0:125
1.0c:3.0:2.0:8.0:0.0:125
1.0d:3.0:2.0:8.0:0.0:125
1.0c:3.0:2.0:8.0:0.0:125
1.0c:3.0:2.0:8.0:0.0:125
1.0c:3.0:2.0:8.0:0.0:125
1.0c:3.0:2.0:8.0:0.0:125

180
180
180
180
180
180
180
180
180
180
180
180
180
180
200
200
180

1
1
1
1
1
1
1
1
1
1
1
1
1
2
1
2
2

193

were determined by the t-plot method. The mesopore volumes


were calculated from the difference between the total pore volume
and the micropore volume. The pore size distributions were
determined from desorption branches by BJH method. Thermogravimetric (TG) and differential thermal analysis (DTA) (Netzsch,
STA409) was performed in air at a heating rate of 10  C/min.
Scanning electron microscope images were obtained from a SEM
(JSM-6701F) instrument. Transmission electron micrographs (TEM)
were obtained on a JEOL JEM-2010 microscope. For TEM, the
sample was ground, dispersed in ethanol, and deposited on a holey
carbon lm supported on a copper grid.
3. Results and discussion
3.1. Optimization of synthesis conditions
For the synthesis of aluminophosphate molecular sieves, the
resulting structure types are commonly controlled by the P2O5/
Al2O3 ratio in the initial gel [16]. As shown in Fig. 1, when P2O5/
Al2O3 ratio is 2.0 (sample P-2), ve main diffraction peaks ascribed
to AFI phase can be clearly observed. Meanwhile, large amounts of
dense phases, i.e. tridymite, cristobalite and berlinite, are also
detected in this sample. As P2O5/Al2O3 ratio is increased to 3.0
(sample P-3), dense phases still exist, but the intensity of those
diffraction peaks attributed to tridymite and berlinite signicantly
declines. On the other hand, the intensity of ve diffraction peaks of
AFI phase increases in varying degrees. It is interesting to note that
imperfect AEL phase (sample P-6) can be produced when P2O5/
Al2O3 ratio arrives at 6.0. To the best of our knowledge, this may be
the rst case in which tetraethylammonium bromide is used as the
structure directing agent to synthesize AEL phase. Based on above
analysis, it can be found that the appropriate P2O5/Al2O3 ratio is 3.0
for the synthesis of AFI phase.
To synthesize AFI phase in a cost-effective manner, the effect of
the amount of activated carbon and the role of pretreatment were
studied in detail. As seen in Fig. 2, the product is pure berlinite
(sample C-0) when no activated carbon is added to the synthesis
mixture. As the C/Al2O3 is decreased from 125 to 62.5 (sample C62.5), the major product is dense cristobalite accompanied by minor AFI phase. It is worth mentioning that dense cristobalite phase
instead of AFI is produced when raw activated carbon is introduced
to the synthesis mixture. In Xiao and co-workers research, AFI
phase was successfully prepared in the presence of TEABr and di-n-

Additional water, not include the water contained in HF and H3PO4.


Raw activated carbon is used in the synthesis.
c
Aluminum isopropoxide.
d
Pseudo-boehmite.
e
20-min mechanical treatment by ball mill. Unless otherwise specied,
aluminum hydroxyacetate is used as the aluminum source.
b

Fig. 1. XRD patterns of three samples under the conditions of different P2O5/Al2O3
ratios.

194

X. Zhao et al. / Microporous and Mesoporous Materials 213 (2015) 192e196

Fig. 2. XRD patterns of four samples under the conditions of low C/Al2O3 and TEABr/
Al2O3 ratios.

propylamine [2]. If there was only one template in the synthesis,


AFI phase cannot be obtained. These results indicate that in our case
the interaction between pretreated activated carbon and TEABr
may play a key role in synthesizing AFI phase. To corroborate the
role of pretreated activated carbon, we performed the TG-DTA experiments of two mixtures prepared in mortar. One is the mixture
of pretreated activated carbon and TEABr (mixture A), and the other
is raw activated carbon and TEABr (mixture B). As shown in Fig. S1,
the initial exothermic effect accompanied by large weight loss,
which is associated with a Hoffmann elimination reaction of tetraethylammonium cations [17], can be clearly observed in the
210e275  C temperature range of DTA curves. The endothermic
peak centered at about 290  C with no obvious weight loss can be
ascribed to the melt of free TEABr. It should be noted that not only
Hoffmann elimination temperature but also the melting temperature of TEABr for mixture A is higher than those for mixture B. Such
results fully conrm that strong interactions occur between pretreated activated carbon and TEABr.
From an economic point of view, the high use of organic template is unfavorable in zeolite synthesis. Therefore, a synthetic
experiment was performed under a TEABr/Al2O3 ratio of 4.0. The
result shows that the product is dominated by dense cristobalite
phase (Fig. 2, sample TEABr-4). Obviously, the suitable ratios of C/
Al2O3 and TEABr/Al2O3 are 125 and 8.0, respectively.
Fluoride has recently been an extremely useful mineralizer for
aluminophosphate [18] and silicate synthesis [19], in which it may
help to solubilize the starting materials under the reaction conditions and in some cases play a structure-directing role. In the
present study, the inuence of hydrouoric acid on the crystallization was investigated by varying the initial HF/Al2O3 ratio. The
results demonstrate that among the three samples (Fig. 3, P-3, HF-1
and HF-4), P-3 has the highest crystallinity in terms of AFI phase.
High hydrouoric acid dosage is favorable for the generation of
dense berlinite. Thus, the optimum HF/Al2O3 ratio is 2.0 in this
study.
The most important factor affecting ionothermal synthesis was
considered to be water content in the original research of Morris'
group [3]. A very small amount of water was advantageous for
ionothermal synthesis [20]. However, the addition of larger
amounts of water to the synthesis mixture seemed to disrupt the
templating ability of the ionic liquid, leading to dense phase being
formed [21]. The synthetic method reported by us has something in
common with ionothermal synthesis, namely, water is not the

Fig. 3. XRD patterns of ve samples under the conditions of different H2O/Al2O3 and
HF/Al2O3 ratios.

dominant species in both methods. Thus, the added water in our


case may play a similar role as that in ionothermal synthesis. As
displayed in Fig. 3, the sample H2O-0 has the least dense phase
among the three samples (P3, H2O-20, H2O-0) synthesized under
the conditions of different H2O/Al2O3 ratios, again verifying the
effect of water. In fact, large amounts of water will compete with
activated carbon to interact with TEABr, thus disturbing the template ability of TEABr. In the subsequent study, additional water was
no longer introduced to the synthesis system.
Aluminum source also has a large impact on the synthesis of
aluminophosphate molecular sieves because different aluminum
sources have distinct reactivity with phosphoric acid. Comparing
the XRD proles of the three samples (Fig. 4, H2O-0, Al-AIP and AlPB), it is found that the diffraction peak intensity of AFI phase is the
highest in the sample Al-AIP. Therefore, aluminum isopropoxide is
chosen as the best aluminum source in the following optimization.
Crystallization temperature and time are two key factors in
controlling the purity of zeolite phase. As shown in Fig. 4, in comparison to the sample Al-AIP synthesized within 1 h microwave
heating, the diffraction peak intensity of dense phases in the
sample T180-2h is prominently decreased when the crystallization
time is extended to 2 h. This result suggests that longer crystallization time seems to favor the transformation of dense phases to

Fig. 4. XRD patterns of seven samples under the conditions of different aluminum
source, crystallization time, temperature and mechanochemical treatment.

X. Zhao et al. / Microporous and Mesoporous Materials 213 (2015) 192e196

other phase. Such phenomenon is quite unusual. In other words,


these so-called dense phases are not really stable. On the other
hand, higher crystallization temperature results in the AFI phase
amorphize because the curved baseline and minor AFI phase
conrm that the major phase is amorphous in the sample T200-1h.
Moreover, by comparing the XRD proles of the samples T200-1h
and T200-2h, it can be seen that amorphous phase is gradually
converted to AFI phase or dense phases with prolonged crystallization time. In Xiao and co-workers study, all AFI phases are
exclusively synthesized at 200  C [2]. But in the present work, the
optimum crystallization temperature and time are 180  C and 2 h.
The varying template types and their interaction with aluminum
and phosphor sources may account for the different crystallization
conditions.
Javier Perez-Ramirez and co-workers recently reviewed the
benets of (seed) milling as a tool for preparing and activating
precursor mixtures prior to their hydrothermal treatment for
zeolite synthesis [22]. A lot of advantages of this mechanochemical
process had been demonstrated in a series of studies. The main
ones are summarized as follows: faster crystallization kinetics [23],
reduced amount of the organic structure-directing agent, easy
attainment of a pure zeolite phase [24]. Motivated by these exciting
results, we performed further synthesis optimization through
replacing mortar with ball mill to pretreat the precursor mixtures.
As shown in Fig. 4, cristobalite disappears almost completely in the
sample Mill-20, at the same time the amount of berlinite is
signicantly decreased when compared with the sample T180-2h.
This result again veries the benecial role of high-energy milling
treatment, which can be attributed, in one hand, to the peculiar
mechanochemical effect, and in other hand, possibly to the
reduction of water. The ball milling process can lead to high
instantaneous local temperature, which may exceed 1000  C [25].
Thus, water in the initial synthesis mixtures will inevitably evaporate during this process, while low amount of water can restrict
the formation of dense phases, which had been discussed in previous section.
3.2. Structure and morphology analysis
After experiencing a series of optimization, the appropriate
conditions for the synthesis of aluminophosphate molecular sieve
with AFI topology are obtained, and showed as follows: the initial
composition is 1.0Al2O3: 3.0P2O5: 2.0HF: 8.0TEABr: 125C,
aluminum isopropoxide is the best aluminum source and activated
carbon needs to be pretreated before used as the reaction medium
and hard template. Among all synthesized samples listed in Table 1,
the samples T180-2h and Mill-20 reveal a relatively pure AFI phase
and high crystallinity. Thus, the two samples are selected and used
for further characterization.
Fig. 5 exhibits the nitrogen sorption isotherms and pore size
distributions (PSDs) of the samples T180-2h and Mill-20. Both
isotherms exhibit a hysteresis loop generated by capillary
condensation in the meso- and macropores, implying that these
materials have a mesopore characteristic [26]. This result is in
agreement with the ndings reported in the literature [26e32].
However, the hysteresises do not have a steep step, and their
shapes indicate that the samples do not contain ordered mesopores. The PSDs (Fig. 5, inset) of these samples determined from the
desorption branch of the isotherm, show two maxima, one at about
3.8 nm, which is probably attributed to the so-called tensile
strength effect [33], and a relatively broad one usually centered at
about 35 nm, representing the mesopores. The pore texture parameters summarized in Table 2 show that the BET surface area and
mesopore volume of the two samples are comparable. Xiao and coworkers reported that the hierarchical porosity of SAPO-5 can be

195

Fig. 5. N2 adsorption/desorption isotherms and pore size distributions (inset) of T1802h and Mill-20. The isotherms are shifted upwards with 40 cm3/g interval between
them.

formed in the conditions of solvent-free and absence of any


mesoscale organic templates [2], and the mesopore volume of
SAPO-5 reported by them (0.18 cm3/g) is comparable to the data
presented here (0.16e0.18 cm3/g). On the other hand, the micropore volume of the sample Mill-20 is a little higher than that in the
sample T180-2h, which can be explained that the former has less
dense phases.
Fig. 6 shows the scanning electron micrographs of calcined
T180-2h and Mill-20. It can be seen from the two images that
both samples are composed of the mixtures of nanoparticle aggregates and typical large hexagonal crystals. Minor irregular
crystals with high aspect ratio can also be observed (not shown),
which may belong to the dense phases. In addition, the large
hexagonal crystals are found to be covered by few amounts of
nanoparticles. The electron diffraction patterns shown in Fig. S2
indicate that these nanoparticles are crystalline in nature. For
the two samples, the heterogeneity regarding the morphology
may be due to the local over-heating effect related with microwave heating [34]. The restricted mass transfer occurred in solidstate reaction are also responsible for this result. Based on the
SEM images, the size of nanoparticles is estimated to be 60 nm in
both samples. According to the Scherrer equation, the average
particle sizes of T180-2h and Mill-20 are calculated to be 66 nm
and 62 nm, respectively, which are in agreement with the SEM
observation. As expected, the presence of mesoporosity in the two
samples can be ascribed to the intercrystalline void space between nanoparticles. It is the heterogeneity that results in the
wide pore size distribution, as has been revealed in previous N2
physisorption analysis.

Table 2
N2 physical adsorption-desorption data of representative samples.
Sample

SBET
(m2/g)a

SExt
(m2/g)b

Vmicro
(cm3/g)c

Vmeso
(cm3/g)d

PSDs
(nm)e

T180-2h
Mill-20

166
172

64
53

0.047
0.055

0.18
0.16

5e100
5e100

a
b
c
d
e

BET surface area.


External surface area.
Micropore volume, t-plot method.
Mesopore volume total pore volume e micropore volume.
Pore size distributions.

196

X. Zhao et al. / Microporous and Mesoporous Materials 213 (2015) 192e196

Fig. 6. SEM images of the typical samples T180-2h (left) and Mill-20 (right) (Inset is the image of large hexagonal crystals).

4. Conclusions
In conclusion, aluminophosphate molecular sieve with AFI topology has been synthesized by microwave heating under solventfree conditions. The appropriate initial composition for AFI structure is 1.0Al2O3: 3.0P2O5: 2.0HF: 8.0TEABr: 125C. Aluminum isopropoxide is the optimum aluminum source and activated carbon
needs to experience pretreatment. The characterization results
indicate that the resulting AFI molecular sieves have hierarchical
micro- and meso-porous structure.
The synthetic strategy reported here well integrates the features
of carbon template, microwave synthesis and solvent-free synthesis. It can be conveniently extended to the synthesis of other hierarchically structured aluminophosphate molecular sieves by
utilizing the interaction between activated carbon and organic
template. Such work is still under way.
Acknowledgements
This work was supported by the National Natural Science
Foundation of China (Grant No. 21306072, 51263012) and Development Program of Lanzhou University of Technology for excellent
teachers (Grant No. Q201113). We would like to thank Professor W.
Zhou for helpful discussion. We cordially thank the Reviewers for
providing us with valuable comments and suggestions.
Appendix A. Supplementary data
Supplementary data related to this article can be found at http://
dx.doi.org/10.1016/j.micromeso.2015.03.031.
References
[1] J. Caro, F. Marlow, K. Hoffmann, C. Striebel, J. Kornatowski, I. Girnus, M. Noack,
lsch, in: S.-K.I. Hakze Chon, U. Young Sun (Eds.), Studies in Surface SciP. Ko
ence and Catalysis, Elsevier, 1997, pp. 2171e2178.
[2] Y. Jin, X. Chen, Q. Sun, N. Sheng, Y. Liu, C. Bian, F. Chen, X. Meng, F.-S. Xiao,
Chem. Eur. J. 20 (2014) 17616e17623.
[3] E.R. Cooper, C.D. Andrews, P.S. Wheatley, P.B. Webb, P. Wormald, R.E. Morris,
Nature 430 (2004) 1012e1016.
[4] E.R. Parnham, R.E. Morris, Acc. Chem. Res. 40 (2007) 1005e1013.
[5] L. Han, Y. Wang, C. Li, S. Zhang, X. Lu, M. Cao, AICHE J. 54 (2008) 280e288.

[6] Q. Wu, X. Wang, G. Qi, Q. Guo, S. Pan, X. Meng, J. Xu, F. Deng, F. Fan, Z. Feng,
C. Li, S. Maurer, U. Mueller, F.-S. Xiao, J. Am. Chem. Soc. 136 (2014)
4019e4025.
[7] P. Zhang, L. Wang, L. Ren, L. Zhu, Q. Sun, J. Zhang, X. Meng, F.-S. Xiao, J. Mater.
Chem. 21 (2011) 12026e12033.
[8] L. Ren, Q. Wu, C. Yang, L. Zhu, C. Li, P. Zhang, H. Zhang, X. Meng, F.-S. Xiao,
J. Am. Chem. Soc. 134 (2012) 15173e15176.
[9] Y. Jin, Q. Sun, G. Qi, C. Yang, J. Xu, F. Chen, X. Meng, F. Deng, F.-S. Xiao, Angew.
Chem. Int. Ed. 52 (2013) 9172e9175.
[10] H.J. Wang, L. Wang, Y. Nemoto, N. Suzuki, Y. Yamauchi, J. Nanosci. Nanotechnol. 10 (2010) 6489e6494.
[11] H.S. Huang, K.H. Chang, N. Suzuki, Y. Yamauchi, C.C. Hu, K.C.W. Wu, Small 9
(2013) 2520e2526.
[12] G.A. Tompsett, W.C. Conner, K.S. Yngvesson, Chemphyschem 7 (2006)
296e319.
[13] X.H. Zhao, C.X. Kang, H. Wang, C.H. Luo, G.X. Li, X.L. Wang, J. Porous Mater. 18
(2011) 615e621.
[14] X.H. Zhao, H. Wang, B.F. Dong, Z.P. Sun, G.X. Li, X.L. Wang, Micropor. Mesopor.
Mater. 151 (2012) 56e63.
[15] X.H. Zhao, H. Wang, C.X. Kang, Z.P. Sun, G.X. Li, X.L. Wang, Micropor. Mesopor.
Mater. 151 (2012) 501e505.
[16] S. Oliver, A. Kuperman, G.A. Ozin, Angew. Chem. Int. Ed. 37 (1998) 47e62.
[17] F. Taborda, Z. Wang, T. Willhammar, C. Montes, X. Zou, Micropor. Mesopor.
Mater. 150 (2012) 38e46.
[18] R.E. Morris, A. Burton, L.M. Bull, S.I. Zones, Chem. Mater. 16 (2004) 2844e2851.
[19] S.I. Zones, R.J. Darton, R. Morris, S.J. Hwang, J. Phys. Chem. B 109 (2005)
652e661.
[20] H. Ma, Z. Tian, R. Xu, B. Wang, Y. Wei, L. Wang, Y. Xu, W. Zhang, L. Lin, J. Am.
Chem. Soc. 130 (2008) 8120.
[21] D.S. Wragg, A.M.Z. Slawin, R.E. Morris, Solid State Sci. 11 (2009) 411e416.
[22] G. Majano, L. Borchardt, S. Mitchell, V. Valtchev, J. Perez-Ramirez, Micropor.
Mesopor. Mater. 194 (2014) 106e114.
[23] V. Valtchev, S. Mintova, V. Dimov, A. Toneva, D. Radev, Zeolites 15 (1995)
193e197.
[24] L.M. Vtjurina, S.S. Khvoshchev, in: B.M.J.R.F. Rodriguez-Reinoso, K. Unger
(Eds.), Studies in Surface Science and Catalysis, Elsevier, 2002, pp. 671e675.
[25] C. Suryanarayana, Prog. Mater. Sci. 46 (2001) 1e184.
[26] K. Utchariyajit, S. Wongkasemjit, Micropor. Mesopor. Mater. 135 (2010)
116e123.
[27] D. Verboekend, M. Milina, J. Perez-Ramirez, Chem. Mater. 26 (2014)
4552e4562.
[28] N. Danilina, S.A. Castelanelli, E. Troussard, J.A. van Bokhoven, Catal. Today 168
(2011) 80e85.
[29] N. Danilina, F. Krumeich, J.A. van Bokhoven, J. Catal. 272 (2010) 37e43.
[30] K. Egeblad, M. Kustova, S.K. Klitgaard, K. Zhu, C.H. Christensen, Micropor.
Mesopor. Mater. 101 (2007) 214e223.
[31] J. Kim, S. Bhattacharjee, K.-E. Jeong, S.-Y. Jeong, M. Choi, R. Ryoo, W.-S. Ahn,
New. J. Chem. 34 (2010) 2971e2978.
[32] K. Murthy, S.J. Kulkarni, S.K. Masthan, Micropor. Mesopor. Mater. 43 (2001)
201e209.
rez-Ramrez, Micropor. Mesopor. Mater. 60 (2003)
[33] J.C. Groen, L.A.A. Peffer, J. Pe
1e17.
[34] Y.P. Xu, Z.J. Tian, S.J. Wang, Y. Hu, L. Wang, B.C. Wang, Y.C. Ma, L. Hou, J.Y. Yu,
L.W. Lin, Angew. Chem. Int. Ed. 45 (2006) 3965e3970.

Você também pode gostar