Vous êtes sur la page 1sur 11

Synthesis, structure and acid characteristics of partially

crystalline silicalite-1 based materials


Yin Fong Yeong, Ahmad Zuhairi Abdullah, Abdul Latif Ahmad, Subhash Bhatia
*
School of Chemical Engineering, Engineering Campus, Universiti Sains Malaysia, Seri Ampangan, Nibong Tebal, 14300 Seberang Perai Selatan, Pulau Pinang, Malaysia
a r t i c l e i n f o
Article history:
Received 5 February 2009
Received in revised form 14 March 2009
Accepted 30 March 2009
Available online 5 April 2009
Keywords:
Synthesis
Characterization
Partially crystalline silicalite-1
Phenethyltrimethoxysilane (PE)
Acidity
a b s t r a c t
A series of partially crystalline silicalite-1 based materials were synthesized by varying the molar ratio of
organosilane source, phenethyltrimethoxysilane (PE) to tetraethylorthosilicate (TEOS) in the range of
0.050.50, using one step co-condensation hydrothermal synthesis method. The phenethyl group was
subsequently sulfonated to arenesulfonic acid group following strong acid treatment. The resulting mate-
rials were characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), transmission
electron microscopy (TEM), nitrogen adsorption and desorption and elemental analysis. The structure
of these materials was determined by Fourier transform infrared spectroscopy (FTIR),
29
Si and
13
C solid
state NMR. The % crystallinity of the partially crystalline silicalite-1 as determined from XRD was in
the range of 3373%. The average crystallite size decreased with the increase of PE concentration in
the synthesis mixture. The thermogravimetric analysis shows that the structures were thermally stable
up to 550 C after elimination of the structure directing agents (SDAs) by calcination at 420 C. The acid
capacities of these materials ranged from 2.52 to 6.63 mmol H
+
/g.
2009 Elsevier Inc. All rights reserved.
1. Introduction
Zeolites are crystalline microporous aluminosilicates materials
with well-dened micropore structure, good thermal and struc-
tural stability and resistance to relatively extreme chemical envi-
ronment [1,2]. The use of zeolites as acid catalysts for industrial
processes, particularly in petroleum rening and petrochemicals,
has been widely reported in the literature [3,4]. The most impor-
tant applications are found in the eld of cracking, hydrocracking,
isomerization, alkylation and reforming reactions [5,6]. In the uti-
lization of zeolitic catalysts, the reaction activity and product selec-
tivity depend strongly on the number, strength and nature of the
acid sites present, crystal/particle size and morphology, as well
as the shape and size of the micropores which can induce different
shape-selectivity effects on the product distribution [7]. These par-
ticular properties are obtained by varying Si/Al ratio, crystallite
size and morphology, or modications of extra-framework by cat-
ion exchange, pore blockage and elimination of external sites, iso-
morphous substitution and functionalization with organic group
[4,811].
Among all the zeolites, MFI-type (ZSM-5 and silicalite-1) has
been extensively studied for industrial processes because of its
medium pore-size dimension (0.54 0.56 nm). A signicant re-
search effort has been devoted to the synthesis of MFI zeolite with
smaller crystal size due to its advantages [12]. In the early 1980s,
Jacobs et al. [13] reported the synthesis of partial crystalline
ZSM-5 zeolite which contained small crystallites of less than
8 nm in size within an amorphous matrix, using shorter hydrother-
mal synthesis times. Nicolaides et al. [3,4,14] reported the synthe-
sis of partially crystalline ZSM-5 based materials (NAS materials)
using lower synthesis temperatures, ranging from 25 to 140 C.
ZSM-5 based materials with XRD crystallinity level as low as 2%
exhibited superior catalytic performance (higher selectivities and
yields) in the skeletal isomerization of linear butenes to iso-butene,
due to the decreased of zeolite pore lengths presented in these low
crystalline materials [4]. These materials with XRD crystallinities
lower than 30%, partially crystalline samples possessing 3070%
crystallinity and highly crystalline materials with >70% XRD crys-
tallinity, were tested for their catalytic performance in n-hexane
cracking activity. They reported that the number of strong
Bronsted acid sites and n-hexane cracking activity were found to
be disproportionately low for the samples with XRD relative
crystallinities <30%, and both become signicantly higher only at
crystallinity levels higher than 30% [14].
In reviewing the reports on the improved catalytic performance
by using smaller crystal size and partially crystalline ZSM-5, it has
drawn interest to synthesize partially crystalline silicalite-1 based
materials. These materials could produce smaller crystal size and
extra-framework (amorphous species) with active acid sites. Silica-
lite-1 is an aluminum-free analogue of ZSM-5 (Si/Al = 1) which is
catalytically inactive in its pure form. In dening the acidity of the
1387-1811/$ - see front matter 2009 Elsevier Inc. All rights reserved.
doi:10.1016/j.micromeso.2009.03.038
* Corresponding author. Tel.: +60 604 5996409; fax: +60 604 5941013.
E-mail address: chbhatia@eng.usm.my (Subhash Bhatia).
Microporous and Mesoporous Materials 123 (2009) 129139
Contents lists available at ScienceDirect
Microporous and Mesoporous Materials
j our nal homepage: www. el sevi er . com/ l ocat e/ mi cr omeso
partially crystalline silicalite-1 based materials, the amount and
type of extra-framework or amorphous species presence is
important.
In order to create acid sites in the partially crystalline silicalite-
1 based materials, the organic-functional group should be intro-
duced [15,16]. Jones et al. [1719] reported that the organic group
is difcult to introduce into the zeolite micropore due to the large
size of organic species, which resulted in a disruption of the crystal
structure and multiphase, and thus, crystalline/amorphous mix-
ture is produced. It is expected that the transformation of the or-
ganic group attached on the amorphous materials into
organosulfonic acid group will result in a useful acid materials
which could be explored as catalyst for industrial reaction [20].
The main advantage of these materials is the absence of blockage
of silicalite-1 by acid sites from the extra-framework species that
located on the outer surface of the crystals.
In the present study, phenethyltrimethoxysilane (PE) as an org-
anosilance source has been utilized in the synthesis of partially
crystalline silicalite-1 based materials. The effect of the organosi-
lanes concentration present in the initial synthesis mixtures on
the formation of partially crystalline silicalite-1 based materials
is systematically studied. The organic group is subsequently trans-
formed into arenesulfonic acid via sulfonation. The crystalline and
extra-framework/amorphous phases of the samples were charac-
terized for their % crystallinity, crystal morphology, composition,
structure, thermal stability and surface area characteristic by a
number of physical and chemical techniques. The acid capacities
of the samples were obtained by acidbase titration. The acidic
properties of these materials are correlated with the composition
of the crystalline and amorphous phases present in the samples.
2. Experimental
2.1. Samples preparation
The synthesis of partially crystalline silicalite-1 based materials
and parent silicalite-1 (siliceous ZSM-5) sample are carried out fol-
lowing the method reported by Lai et al. [21] and co-condensation
method [10,2224].
A series of samples were synthesized using tetraethylorthosili-
cate (TEOS) as inorganic silica source and phenethyltrimethoxysi-
lane (PE) as organosilica source. The synthesis solution was
prepared by adding tetrapropylammonium hydroxide (TPAOH,
1 M, Merck) in a polypropylene bottle containing deionized (DDI)
water. TEOS (>98%, Merck) was added drop wise and the solution
was stirred gently. An appropriate amount of PE (>97%, Fluka)
was then added slowly to complete the reaction mixture. The nal
molar composition of the synthesis solution is presented as:
51 x TEOS : TPAOH : 1000 DDIH
2
O : 5xPE
where x is molar composition and x = 0.05, 0.1, 0.15 0.2, 0.3 and 0.5
in the present study and for the synthesis of silicalite-1, x = 0. The
reaction mixture was stirred vigorously for 1 day at room tempera-
ture. After vigorous stirring, the synthesis solution was transferred
into Teon-lined reaction pressure vessel. The reaction vessel was
sealed and heated to 175 C for 1 day. After hydrothermal synthesis,
the vessel was taken out and quenched to room temperature. The
samples were collected and washed by repeated centrifugation
and decanting until the pH of the seed suspension became 8. The
samples were then dried at 100 C overnight and further calcined
at 420 C for 15 h with heating and cooling ramping rate of 0.5 C/
min.
In the present work, the partially crystalline silicalite-1 based
materials were synthesized by increasing the PE concentration in
the synthesis mixture while keeping the synthesis temperature
and duration constant at 175 C and 1 day, different from the
method reported by Nicolaides [4].
2.2. Post-synthesis modication
The phenethyl group present in partially crystalline silicalite-1
based materials was sulfonated to arenesulfonic acid group follow-
ing the method reported by Holmberg et al. [25]. The calcined sam-
ples were dispersed in 96 wt.% concentrated sulfuric acid and
treated at 80 C for 24 h under stirring. After 24 h, the suspension
was ltered, washed extensively with DDI water and dried for
overnight at 100 C. Finally, the prepared samples were stored in
a desiccator before characterization.
The sample synthesized with a ratio of 0.95 TEOS/0.05 PE is
coded as SILPE5 and SILPE5SO
3
H after sulfonation. Six samples
with different PE composition were prepared and presented in Ta-
ble 1. For comparison purpose, parent silicalite-1 (SIL-1) sample
was also prepared.
2.3. Characterization
2.3.1. X-ray diffraction (XRD)
The crystallinity of the prepared samples was obtained from
XRD analysis. The analysis was done using X-ray diffractometer
Siemen (D5000) with Cu Ka radiation (k = 1.5406 ) operated at
40 kV and 30 mA. Data were collected stepwise over 5 6 2h 6
40 angular region. The crystallinity of samples was determined
based on the ratio of the major peak intensities of the samples
(at 2h 7.7, 8.8 and 23), relative to those of highly crystalline
reference material [4]. The crystallinity is dened as:
% XRD crystallinity
sum of peak int ensities of the samples
sum of peak int ensities of the reference
100
1
A highly crystalline silicalite-1 sample was used as a reference
material.
2.3.2. Scanning electron microscope (SEM)
Surface morphology of the samples was studied using scanning
electron microscope (Zeiss Supra 35VP) equipped with W-Tung-
sten lament, operated at 3.00 kV.
2.3.3. Transmission electron microscope (TEM)
TEM micrographs were obtained on a CM 12 Philips electron
microscope equipped with an image analyzer (Model Soft Imaging
System, SIS 3.0), operating at 80 kV. The samples were prepared by
evaporating one drop of powdered sample-EtOH suspension (after
sonication) onto a carbon coated lm supported on a 3 mm
diameter, 400 mesh copper grid.
Table 1
Parent and partially crystalline silicalite-1 based samples prepared in this study.
Sample Sample after sulfonation TEOS, 5(1 x) PE, 5x
Silicalite-1 (SIL-1,
as reference material)
1.00 0.00
SILPE5 SILPE5SO
3
H 0.95 0.05
SILPE10 SILPE10SO
3
H 0.90 0.10
SILPE15 SILPE15SO
3
H 0.85 0.15
SILPE20 SILPE20SO
3
H 0.80 0.20
SILPE30 SILPE30SO
3
H 0.70 0.30
SILPE50 SILPE50SO
3
H 0.50 0.50
130 Y.F. Yeong et al. / Microporous and Mesoporous Materials 123 (2009) 129139
2.3.4.
29
Si MAS NMR and
13
C CP-MAS solid state NMR
The
29
Si and
13
C CP NMR spectra were recorded at room tem-
perature under magic angle spinning (MAS) on a Bruker AV400
spectrometer. The spectra were recorded with a 4 mm probe at a
sample spinning rate of 7 kHz and quoted relative to tetramethyl-
silane (TMS). The single pulse
29
Si was acquired using pulses of
4 ls and a recycle delay of 240 s. The
13
C cross-polarization was
measured with a recycle delay of 5 s.
2.3.5. Fourier transform infrared spectroscopy (FT-IR)
The degree of chemical interaction of the organic and inorganic
phases and the nature of the acid sites were determined using FT-
IR technique. The IR spectra were recorded using a PerkinElmer
FT-IR (Model 2000) in the range of 4004000 cm
1
using KBr
method. In order to determine the natural acid sites, prior to KBr
method, the sample was rst exposed to excess pyridine for 1 h;
after degassing at 200 C, followed by desorption of physically ad-
sorbed of pyridine at 150 C under vacuum.
2.3.6. Elemental analysis
The chemical analysis of the samples was carried out using a
PerkinElmer CHNS/O analyzer where the sample was combusted
in an oxygen-rich environment at 975 C and analyzed for carbon,
hydrogen and sulfur content.
2.3.7. Acid capacity determination by titration technique
The acid capacity of the samples was determined by titration
technique [10,26]. The acid exchange capacity of the sample was
measured, using aqueous solutions of sodium chloride (NaCl,
2 M), tetramethylammoniumchloride (TMAC, 0.05 M) and tetrabu-
tylammonium chloride (TBAC, 0.05 M) as ion-exchange agents.
After treated 0.1 g sample at 200 C for one day, the sample was
added to 20 mL of the aqueous solution containing the correspond-
ing salt. The resulting suspension was equilibrated for 1 day and ti-
trated by dropwise addition of 0.01 M NaOH aqueous solution
using phenolphthalein as an indicator. The acid exchange capacity
is expressed as mmol H
+
/g of sample.
2.3.8. Thermal gravimetric analysis (TGA)
The thermogravimetric analysis was conducted under nitrogen
gas with a TGA/SDTA analyzer (Metler Teledo 851E) to study the
thermal stability of the synthesized samples. The synthesized
and calcined samples were subjected at the heating rate of 10 C/
min, from room temperature until 900 C.
2.3.9. Nitrogen adsorptiondesorption measurement
The pore characteristic (pore volume, pore-size and surface
area) of the samples were measured by nitrogen adsorption using
a Micromeritics ASAP 2000 instrument. Samples of 0.05 g were
outgassed overnight at 105 C under vacuum prior to the analysis.
3. Results and discussion
3.1. Crystallinity and crystal morphology by XRD, SEM and TEM
3.1.1. X-ray diffraction (XRD)
Fig. 1 shows the powder XRD patterns of the partially crystal-
line silicalite-1 based materials synthesized at various PE concen-
tration. Parent silicalite-1 is highly crystalline and its XRD
pattern shows the presence of the major peaks matching well with
the peaks reported for silicalite-1 [21]. Samples SILPE5SO
3
H to
SILPE20SO
3
H exhibited diffraction patterns similar to that of
parent silicalite-1 but at a relatively low intensity. This shows that
silicalite-1 structure was retained in the materials. The degree of
crystallinity was indicated by the peak intensity. It can be seen
from Fig. 1 and Table 2 that the intensity of the samples gradually
decreased with increasing PE concentration in the synthesis mix-
ture, from 73% for sample SILPE5SO
3
H to 33% for sample SIL
PE20SO
3
H. The XRD patterns for SILPE30SO
3
H and SILPE50
SO
3
H displayed a broad features (Fig. 1f and g) in the range
2h = 2030, which could be indicative of an amorphous phase.
This observation shows that crystalline materials cannot be formed
when the PE loading in the synthesis mixture reached 30% of the
total silica source. The XRD results were in agreement with the re-
sults reported by Jones et al. [1719]. The higher loading of orga-
nosilane resulted in the disruption of the crystal structure due to
the difculties of incorporation of longer organic fragment into
the crystalline zeolitic framework, and thus, resulting in the amor-
phous phases.
For all the partially crystalline samples (SILPE5SO
3
H to SIL
PE20SO
3
H) obtained, the peaks were less intense and broader
compared to parent silicalite-1, suggesting that the partially crys-
talline silicalite-1 based samples present in smaller crystalline do-
mains [27,28]. The average crystallite size of the samples is
presented in Table 2. It was calculated from the Scherrer equation
based on the half-width of diffraction lines assigned to (0 5 1). The
average crystallite size for parent silicalite-1 was 54 nm which was
comparable to the standard silicalite-1 reported in the literature
(60 nm) [29]. The crystallite size of the partially crystalline silica-
lite-1 based samples decreased, from 39 nm for sample SILPE5
SO
3
H to 29 nm for sample SILPE20SO
3
H, with the reduction in
crystallinity.
3.1.2. Scanning electron microscopy (SEM)
The SEM images for the parent silicalite-1 and partially crystal-
line silicalite-1 based samples are shown in Fig. 2. Silicalite-1
2, degree
(a)
(b)
(c)
(d)
(e)
(f)
(g)
I
n
t
e
n
s
i
t
y

(
a
.
u
.
)

5 10 15 20 25 30 35 40
(051)
(020) (101) (501)
(303)
(151)
Fig. 1. XRD pattern for (a) SIL-1, (b) SILPE5SO
3
H, (c) SILPE10SO
3
H, (d) SIL
PE15SO
3
H, (e) SILPE20SO
3
H, (f) SILPE30SO
3
H, and (g) SILPE50SO
3
H.
Table 2
X-ray results for the samples.
Sample % Crystallinity Average crystallite size
a
, D, nm
SIL-1 100 54.3
SILPE5SO
3
H 73 + Amorphous 39.1
SILPE10SO
3
H 60 + Amorphous 37.7
SILPE15SO
3
H 47 + Amorphous 30.5
SILPE20SO
3
H 33 + Amorphous 28.7
SILPE30SO
3
H Amorphous n.d
SILPE50SO
3
H Amorphous n.d
a
Crystallite size obtained from Scherrer equation, t = 0.9 k/bcosh measured at
d
051
.
Y.F. Yeong et al. / Microporous and Mesoporous Materials 123 (2009) 129139 131
exhibited a complete crystalline phase and cofn like crystal shape,
where the morphology was similar to the conventional MFI-type
zeolites (Fig. 2a) [30]. By loading PE into the synthesis mixture,
both of the partial crystalline and amorphous phases were ob-
served, but the cofn shape of silicalite-1 crystal was retained
(Fig. 2be). The amorphous phase increased with the increase of
PE loading, until a highly amorphous phase was produced with
30 and 50 mol% PE loading, respectively. Since the crystal sizes
were reported as the length of the prismatic crystals [31], the size
of the crystal reduced from 19 lm for silicalite-1 to 10, 6, 4.7, and
4.6 lm for SILPE5SO
3
H, SILPE10SO
3
H, SILPE15SO
3
H and
SILPE20SO
3
H, respectively. The reduction in the crystals size
and the presence of amorphous phases explain the drop in the
crystallinity of the samples; and vice verse, where the SEM results
were consistent with the XRD results.
3.1.3. Transmission electron microscope (TEM)
Fig. 3 shows the TEM micrograph of the parent silicalite-1 and
partially crystalline silicalite-1 based samples. The parent silica-
lite-1 was formed by spherical shape crystallites with dimension
Fig. 2. SEM micrographs for (a) SIL-1, (b) SILPE5SO
3
H, (c) SILPE10SO
3
H, (d) SILPE15SO
3
H, (e) SILPE20SO
3
H, (f) SILPE30SO
3
H, and (g) SILPE50SO
3
H.
132 Y.F. Yeong et al. / Microporous and Mesoporous Materials 123 (2009) 129139
of about 4060 nm (Fig. 3a). From TEM micrographs, a consider-
able reduction in the crystallite size was observed for the partially
crystalline silicalite-1 based samples when higher PE composition
was added into the synthesis mixture. It can be seen that the sam-
ples SILPE5SO
3
H and SILPE10SO
3
H also consist of small well-
formed crystallite of a nearly spherical shape, with an average size
ranged from 2040 nm. A similar image can be observed for the
SILPE15SO
3
H and SILPE20SO
3
H sample, except that the sizes
are slightly smaller, in the range of 1530 and 1030 nm, respec-
tively. They are also more separated from each other as compared
to the relatively more aggregated morphology of the SILPE5SO
3
H
and SILPE10SO
3
H samples. The crystallite size obtained from
TEM agreed well with the value determined from XRD data (indi-
cated in Table 2), and comparable with the crystallite size of NAS
materials reported by Triantafyllidis et al. [7]. The reduction in
the crystallite size was probably due to the formation of a large
number of nuclei in the rst steps that grow very slowly [32],
which reduced the relative crystallinity of the partially crystalline
silicalite-1 based samples. However, the presences of these small
particles are not small enough in order to generate interparticle
ordered mesoporosity [7,33]. The TEM micrograph for sample
SILPE30SO
3
H and SILPE50SO
3
H are shown in Fig. 3f and g,
Fig. 3. Transmission electron microscopy (TEM) images of (a) SIL-1, (b) SILPE5SO
3
H, (c) SILPE10SO
3
H, (d) SILPE15SO
3
H, (e) SILPE20SO
3
H, (f) SILPE30SO
3
H, and (g)
SILPE50SO
3
H.
Y.F. Yeong et al. / Microporous and Mesoporous Materials 123 (2009) 129139 133
respectively. The micrographs show that the irregular grains were
obtained, different from the spherical particles formed by other
partially crystalline silicalite-1 based samples with lower PE load-
ing (Fig. 3be). The difference in particles shape could be explained
by the presence of the amorphous phase for both of these samples
as indicated by XRD pattern (Fig. 1) and SEM images (Fig. 2).
3.2. Structural studies and incorporation of organic-functional group
as indicated by,
29
Si MAS,
13
C CP-MAS solid state NMR and FT-IR
3.2.1.
29
Si MAS and
13
C CP-MAS solid state NMR
Wang et al. [9,10] reported that
13
C and
29
Si solid state NMR
spectroscopies are useful for providing chemical information
regarding the condensation of organosiloxane.
29
Si MAS NMR spec-
trum for SILPE10SO
3
H sample is shown in Fig. 4a and the
description of the various structural units of silicon atoms for the
sample is presented in Table 3. Three peaks corresponding for Q
2
(92 ppm), Q
3
(103 ppm) and Q
4
(112 ppm) and three weaker
peaks assigned for T
1
(48 ppm), T
2
(56 ppm) and T
3
(65 ppm) were found in the SILPE10SO
3
H sample. The appear-
ance of the T
m
peaks indicates that some of the linkages between
the silicon atom containing the organic group and adjacent silicon
atoms were hydrolyzed to produce silanol groups during the acid
treatment process [25]. Fig. 4a shows that peak T
3
was predomi-
nant over T
2
implying that the organosiloxane (PE) are effectively
condensed as a part of the sample structure, probably grafted on
amorphous phases [10]. The relative area ratio of Q
4
/(Q
3
+ Q
2
) of
2.83 shows that very high condensation of TEOS under the synthe-
sis condition [9,10]. The
13
C CP-MAS NMR spectrum of the SIL
PE10SO
3
H sample is shown in Fig. 4b. Three peaks at 128, 141
and 150 ppm assigned to carbons on the aromatic ring, C
II
, C
I
and
C
III
, respectively. The corresponding carbon atoms on the methy-
lene group, C
IV
and C
V
are shown at peaks 13 and 29 ppm, respec-
tively [10,34]. These peaks are characteristic of the arenesulfonic
acid group grafted on the amorphous structure, which further con-
rms that phenethyl group is co-condensed in the sample struc-
ture and sulfonated to acid group [10].
3.2.2. Fourier transform infrared spectroscopy
FT-IR spectroscopy was used to identify the structure vibration
and surface hydroxyl groups of the samples [28]. The FT-IR spec-
trum of the parent silicalite-1 and pyridine adsorbed partially crys-
talline silicalite-1 based samples are shown in Fig. 5. The analyzed
IR absorption peaks of the samples and the corresponding refer-
ences are listed in Table 4. Parent silicalite-1 shows a strong peak
at 550 cm
1
, assigned to MFI-structured zeolite [29,36]. It can be
seen that with increasing of PE loading in the synthesis mixture,
the partially crystalline silicalite-1 based samples show the reduc-
tion in peak intensity at 550 cm
1
, or even disappears for samples
SILPE30SO
3
H and SILPE50SO
3
H. This observation was mainly
due to the reduction of crystallinity of the sample and the forma-
tion of amorphous phases for 30 and 50 mol% PE loading. The esti-
mated % relative crystallinity based on these IR peak intensities are
shown in Table 5. It was found that the value obtained was compa-
rable with the% relative crystallinity estimated from XRD analysis
(Table 2).
All the samples show the vibration band at 470, 800, and
1100 cm
1
, corresponding to the typical SiOSi bending, SiOSi
symmetric stretching and SiOSi asymmetric stretching with
the condensed silica network, respectively. An absorption peak at
around 960 cm
1
was attributed to stretching vibration of SiOH
group. In comparison to the spectrum of the parent silicalite-1,
the partially crystalline silicalite-1 based samples show additional
peaks at 700, 1250, 1490, 1542, 1638 cm
1
and a broad band cen-
tered at 3430 cm
1
. Two weak absorption peaks at 700 and
1490 cm
1
were assigned to the out-of-plane bending of the para
di-substituted aromatic ring as well as the bending vibration of
the sulfonic acid group and CH in aromatic rings, respectively
Fig. 4. (a)
29
Si MAS, and (b)
13
C CP-MAS solid state NMR spectra of sample SIL
PE10SO
3
H.
Table 3
Description of the various structural units in the partially crystalline silicalite-1 based
samples [10,34,35].
T species
T
1
T
2
T
3
R
H
H
O
O
Si Si O
H
Si
R
O
O
Si Si O
Si
Si
R
O
O
Si Si O
Q species
Q
2
Q
3
Q
4
H
Si
O H
O
O
Si Si O
Si
Si
O H
O
O
Si Si O

Si
Si
O Si
O
O
Si Si O
Arenesulfonic acid group
SO
3
H CH
2 CH
2
III
I
II II
II II
V IV
R = CH
2
, T
m
= RSi(OSi)
m
(OH)
3m
, Q
n
= Si(OSi)
n
(OH)
4n
.
134 Y.F. Yeong et al. / Microporous and Mesoporous Materials 123 (2009) 129139
[10,37]. These results suggested the condensation of PE group into
the sample structure, and the aromatic ring was successfully sulfo-
nated to arenesulphonic acid group, which was consistent with
13
C
CP-MAS solid state NMR spectra. These peaks intensities increase
with increasing the PE loading in the synthesis mixture conrming
that the higher amount of PE group was incorporated into the
structure. The absorbance at 1250 cm
1
assigned to the vibration
of SiC indicates that the SiC bond did not break after calcination
and strong acid treatment, thus, conrming the presence of organic
group. The peaks corresponding to the S'O stretching vibrations
of sulfonic acid are normally observed in the range of 1000
1200 cm
1
. However, these peaks cannot be resolved due to their
overlap with the absorbance of the SiOSi asymmetric stretch in
the 10001130 cm
1
range and that of the SiC stretch in the
12001250 cm
1
range [38]. The intense band at 3430 cm
1
was
assigned to OH group asymmetric stretching, indicates water
and hydroxyl functional group residual in the partially crystalline
silicalite-1 based samples.
Based on these results, the FT-IR spectra conrmed that both or-
ganic and inorganic structural units were present in the partially
crystalline silicalite-1 based samples. The existence of broad band
at 3430 cm
1
implied the hydrophilic characteristic [39] of the par-
tially crystalline silicalite-1 based samples compared to the hydro-
phobic characteristic of parent silicalite-1 sample. The nature of
the acid sites present in the samples was also conrmed by FT-IR
spectra. The typical peaks of 1638 and 1542 cm
1
indicated that
strong Bronsted acid sites are present in partially crystalline silica-
lite-1 based samples [40] and the silicalite-1 sample did not show
presence of acid sites.
3.3. Chemical properties by elemental analysis and acidbase titration
3.3.1. Elemental analysis
The carbon (C) and sulfur (S) content in the parent silicalite-1
and partially crystalline silicalite-1 based samples were further
Fig. 5. FT-IR spectrum for parent silicalite-1 and partially crystalline silicalite-1 based samples.
Table 4
Characteristic bands (cm
1
) in FT-IR spectrum obtained in the partially crystalline
silicalite-1 based samples.
Peak positions (cm
1
) Assignments Ref.
470 SiOSi bending vibration [37,39]
550 Double ring vibration in the
MFI-structured zeolite
[29,36]
700 Oop ring bend [37]
800 SiOSi symmetric stretching [37,39,41]
970 SiOH stretching [37]
1100 SiOSi asymmetric stretching [37,39]
1250 SiC [10,36,42,43]
1490 Ring [37]
1542 Strong Bronsted acid site [40]
1638 Strong Bronsted acid site [40]
3430 OH [37]
Table 5
Percentage relative FT-IR crystallinity based on the intensity peak at 550 cm
1
.
Sample % Relative FT-IR crystallinity
SIL-1 100
SILPE5SO
3
H 80
SILPE10SO
3
H 67
SILPE15SO
3
H 55
SILPE20SO
3
H 36
SILPE30SO
3
H Amorphous
SILPE50SO
3
H Amorphous
Table 6
C and S content in the synthesis mixture and obtained partially crystalline silicalite-1
based samples.
Samples Synthesis
mixture (wt.%)
a
Elemental
analysis (wt.%)
b
Elemental
analysis (wt.%)
c
C S C S C S
SIL-1 0 0 0 0 0 0
SILPE5SO
3
H 2.29 0 1.52 0 1.14 0.84
SILPE10SO
3
H 4.50 0 1.97 0 1.64 0.89
SILPE15SO
3
H 6.87 0 2.31 0 1.72 1.63
SILPE20SO
3
H 8.97 0 2.87 0 1.83 3.38
a
Relative to the weight of the C and S used in the synthesis mixture.
b
Relative to the weight of the C and S in the as-synthesized samples.
c
Relative to the weight of the C and S in the sulfonated samples.
Y.F. Yeong et al. / Microporous and Mesoporous Materials 123 (2009) 129139 135
characterized by elemental analysis. The results are summarized in
Table 6. Although the increase of PE loading in the synthesis mix-
ture increased the amount of C into the partially crystalline silica-
lite-1 based samples, this amount was relatively low as compared
to the carbon content loaded in the synthesis mixture. This was
mainly due to the difculties of incorporation of large organic
group into the sample structure. After the calcination and sulfona-
tion process, sulfur was present in the sample conrming the pres-
ence of the arenesulfonic acid group. However, there was loss of
small amount of carbon. This observation is in agreement with
the FT-IR spectra, where the SiC bond was present after calcina-
tion and strong acid treatment. Moreover, the FT-IR spectrum for
partially crystalline silicalite-1 based samples show reduction at
700 and 1490 cm
1
adsorption peaks (indicate aromatic ring, Table
4), compared to these peaks obtained for the as-synthesized sam-
ples (not shown).
The relationship between the samples crystallinity with the C
and S content is shown in Fig. 6. Based on this gure, the samples
crystallinity reduced gradually with the increase of C content in the
samples. Reasonably, the amount of S presence in the sample in-
creased with the increase of C content in the sample.
3.3.2. Acidity measurement by titration technique
The accessibility of sulfonic acid in the partially crystalline sili-
calite-1 based samples was determined by acidbase titration
using different ion-exchange agents and the results are presented
in Table 7. The amount of sulfonic acid sites increased from 2.52
to 6.63 mmol H
+
/g (Table 7) when NaCl was used as the ion-ex-
change agent. The acid capacity value is close to those obtained
SIL-PE5
-SO3H
Sample
C (synthesis mixture)
SIL-PE10
-SO3H
SIL-PE15
-SO3H
SIL-PE20
-SO3H
30
40
50
60
70
%

X
R
D

c
r
y
s
t
a
l
l
i
n
i
t
y

w
t
%

0
2
4
6
8
10
C (as-synthesized)
C (sulfonated) S (as-synthesized)
S (sulfonated) S (synthesis mixture)
80
Fig. 6. Relationship between the samples crystallinity with the C and S content.
Table 7
Acid capacity of the partially crystalline silicalite-1 based samples obtained by
titration technique.
Samples S content (mmol/g)
from EA
Acid capacity (mmol H
+
/g)
NaCl TMAC TBAC
SIL-1 0 0.00 0.00 0.00
SILPE5SO
3
H 2.65 2.52 1.18 0.95
SILPE10SO
3
H 2.80 2.72 1.48 1.44
SILPE15SO
3
H 4.10 4.03 1.92 1.57
SILPE20SO
3
H 7.50 6.63 2.07 1.92
Table 8
Acid capacity of other types of acid materials reported in the literature.
Samples Acid capacity (mmol H
+
/g) Ref.
NAS-150 0.53
a
[7]
SBA-15SO
3
H-10 0.82
b
[26]
SiMNPFSO
3
H 0.78
b
[44]
SiO
2
phSO
3
H-10 1.43
b
[10]
SBAphSO
3
H-10 1.12
b
[10]
SBAPrSO
3
H-10 1.04
b
[10]
a
Determined from TPD method.
b
NaCl as an ion-exchange agent.
Temperature (
o
C)
0 200 400 600 800
0 200 400 600 800
0 200 400 600 800
W
e
i
g
h
t

(
%
)
W
e
i
g
h
t

(
%
)
88
105
100
95
90
85
80
75
0 200 400 600 800
0.000
0.002
0.004
0.006
0.008
0.010
0.012
0.014
90
92
94
96
98
100
102
D
e
r
i
v
a
t
i
v
e

W
e
i
g
h
t

(
%
/
o
C
)
D
e
r
i
v
a
t
i
v
e

W
e
i
g
h
t

(
%
/
o
C
)
D
e
r
i
v
a
t
i
v
e

W
e
i
g
h
t

(
%
/
o
C
)
D
e
r
i
v
a
t
i
v
e

W
e
i
g
h
t

(
%
/
o
C
)
-0.007
-0.006
-0.005
-0.004
-0.003
-0.002
-0.001
0.000
Temperature (
o
C)
Temperature (
o
C)
Temperature (
o
C)
W
e
i
g
h
t

(
%
)
95
96
97
98
99
100
101
-0.004
-0.003
-0.002
-0.001
0.000
W
e
i
g
h
t

(
%
)
90
92
94
96
98
100
102
-0.014
-0.012
-0.010
-0.008
-0.006
-0.004
-0.002
0.000
PE
o
550
o
C
2.9 wt%
water
SDA
200
o
C
400
o
C
a b
d c
Fig. 7. TGA curves (a) as-synthesized SIL-1, (b) calcined SIL-1, (c) as-synthesized SIL-PE10, and (d) sulfonated SILPE10 or SILPE10SO
3
H.
136 Y.F. Yeong et al. / Microporous and Mesoporous Materials 123 (2009) 129139
from EA analysis, conrming the almost-complete sulfonation of
phenethyl group. When larger size cations (TMA
+
and TBA
+
) were
used as an ion-exchange agent, the acid capacities of the samples
decreased. These results suggest that most of the acid sites grafted
on the amorphous phases are conned within internal structure
environment, as well as the external surface of the sample.
It is noticeable that the acid capacities of the samples prepared
in the present study were found to be higher than the acid capacity
of other types of acid functionalized materials, which determined
by titration technique (mainly mesoporous type), as well as par-
tially crystalline ZSM-5 zeolite based materials as reported in the
literature (Table 8).
3.4. Thermal stability and surface area characterization
3.4.1. Thermal gravimetric analysis
The thermal stability of all of the samples was determined by
thermal gravimetric analysis under nitrogen atmosphere for the
as-synthesized and sulfonated samples. The TGA curves for SIL-1
and SILPE10SO
3
H samples are shown in Fig. 7. The TGA curves
for other partially crystalline silicalite-1 based samples displayed
the same curve trend as SILPE10SO
3
H sample (as-synthesized
and sulfonated), and thus, are not shown. However, the weight loss
analysis for all of the samples is presented in Table 9.
The rst and second weight loss occurred at 200 and 350
450 C (Fig. 7a and c) for all of the as-synthesized samples was
due to surface dehydration and decomposition of SDA molecules,
respectively. It can be observed that weight loss of partially crys-
talline silicalite-1 based samples due to the water desorption
was lower than that of parent silicalite-1. This indicates that as-
synthesized, partially crystalline silicalite-1 based samples were
more hydrophobic than the as-synthesized parent silicalite-1.
The as-synthesized partially crystalline silicalite-1 based samples
become more hydrophobic with increasing PE loading in the syn-
thesis mixture and their corresponding weight loss due to the
water desorption is reported in Table 9. An additional weight loss
was found in all as-synthesized partially crystalline silicalite-1
based samples, in temperature range of 500600 C. The weight
Table 9
Weight loss analyses obtained from TGA.
Samples % Weight loss of water
in as-synthesized samples
% Weight loss at 500600 C
in sulfonated samples
SIL-1 3.4 0
SILPE5 1.6 2.5
SILPE10 1.3 2.9
SILPE15 1.1 3.7
SILPE20 1.0 4.2
0
20
40
60
80
100
120
140
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
Relative Pressure, P/Po
0
100
200
300
400
500
600
Relative Pressure, P/Po
V
o
l
u
m
e
,

c
c
/
g
V
o
l
u
m
e
,

c
c
/
g
V
o
l
u
m
e
,

c
c
/
g
V
o
l
u
m
e
,

c
c
/
g
V
o
l
u
m
e
,

c
c
/
g
0
50
100
150
200
250
300
350
400
450
Relative Pressure, P/Po
0
50
100
150
200
250
300
350
400
Relative Pressure, P/Po
0
50
100
150
200
250
Relative Pressure, P/Po
N
2
desorption
N
2
adsorption
Silicalite-1
SIL-PE-5-SO
3
H
SIL-PE10-SO
3
H
SIL-PE20-SO
3
H
SIL-PE15-SO
3
H
Fig. 8. Nitrogen adsorptiondesorption isotherms for the samples.
Y.F. Yeong et al. / Microporous and Mesoporous Materials 123 (2009) 129139 137
loss was due to the removal of the incorporated PE group, since
there was no signicant weight loss in the parent silicalite-1 at
500600 C. The corresponding DTG prole displayed three peaks
centered at 200, 400, and 550 C, respectively for partially crystal-
line silicalite-1 based samples compared to two peaks centered at
200 and 400 C for parent silicalite-1.
TGA curves for calcined silicalite-1 and SILPE10SO
3
H samples
are shown in Fig. 7b and d, respectively. It can be seen that the
water and SDA were completely eliminated from the as-synthe-
sized samples through calcination and after strong acid treatment.
A weight loss was observed at 550 C for all the partially crystalline
silicalite-1 based samples, indicating that the arenesulfonic acid
group was thermally stable up to 550 C. Table 9 shows the total
weight loss of the acid group from the samples, estimated as 2.5,
2.9, 3.7, and 4.2 wt.% respectively, which was in agreement with
the C and S contents as determined in the elemental analysis (Sec-
tion 3.3.1).
3.4.2. Nitrogen physisorption
The nitrogen adsorptiondesorption isotherms for parent silica-
lite-1 and partially crystalline silicalite-1 based samples are shown
in Fig. 8. The isotherm for silicalite-1 has a long and horizontal pla-
teau showing typical Type I adsorption isotherm with an H4 hys-
teresis loop as dened by IUPAC [4,42]. Type I isotherm is well-
dened for micropore adsorbent, particularly for silicalite-1. Nev-
ertheless, H4 hysteresis loop is usually associated with thin slit-
like inter-crystalline pores where the pores are mainly in the
micropore range [9,45,46].
However, the isotherms for all of the partially crystalline silica-
lite-1 based samples show a different trend compared to parent sil-
icalite-1 isotherm. A sharp inection was observed in P/P
o
ranging
from 0.8 to 1 and overall nitrogen adsorption volume was in-
creased for SILPE5SO
3
H and reduced consistently when the PE
loading in the synthesis mixture increased. The isotherms exhib-
ited a notable shift of the hysteresis position toward high relative
pressures (P/P
o
= 0.81.0) suggesting that the sample itself under-
went a change in the structure until it produced H3 hysteresis loop.
This observation indicates that the shape of the pores had changed
from thin slit pore to slit-shaped pores, as shown by the presence
of H3 hysteresis loop [45]. The isotherms for partially crystalline
silicalite-1 based samples did not show any limiting adsorption
at high relative pressure, showing that the samples are typical of
open surface materials with large meso/micropores (2030 nm
diameter from BJH analysis), as well as high external surface areas
that allow the formation of multiple adsorbate layers as the P/P
o
increases [7].
The basic physicochemical and textural properties of the par-
tially crystalline silicalite-1 based samples are shown in Table 10.
The specic surface area (S
BET
) was obtained by analyzing nitrogen
adsorption data at 77 K in a relative vapor pressure ranging from
0.01 to 0.3. The total pore volume (V
tot
) was estimated based on
the volume adsorbed at a relative pressure of 0.99 and the micro-
pore volume (V
mic
) was determined by t-plot method. The BET sur-
face area of the parent silicalite-1 was 329 m
2
/g with the total pore
volume of 0.18 cm
3
/g and average pore diameter of 1.5 nm. These
data are comparable and consistent with the data reported in the
literature for silicalite-1 structure [47,48]. With the increase of
arenesulfonic acid in the partially crystalline silicalite-1 based
samples, the S
BET
decreased gradually from 326 to 244 m
2
/g, This
observation was consistent with the literature [9,49], where pres-
ence of organic group in the sample structure resulted in the drop
of surface area, which was due to the occupation of large organic
group in the pore channel.
The total pore volume and the average pore diameter for the
partially crystalline silicalite-1 based samples were higher than
the parent silicalite-1, as shown in Table 10. The presence of arene-
sulfonic acid group increased the total pore volume to 0.66 cm
3
/g
for SILPE5SO
3
H and reduced gradually to 0.26 cm
3
/g for SIL
PE20SO
3
H, the average pore diameter of SILPE5SO
3
H increased
to 7.9 nm for but the average pore diameter of SILPE20SO
3
H
dropped to 4.1 nm. The change in the pore diameter for the par-
tially crystalline silicalite-1 based samples was due to the incorpo-
ration of the PE group into the pore structure during co-
condensation resulted to the larger pore or pore openings [50], as
indicated also in the isotherms (Fig. 8). The further increase of PE
group gradually reduced the total pore volume and average pore
diameter in the partially crystalline silicalite-1 based samples,
which was due to the amorphous phases and partial crystalline
of the structure [46]. Table 10 shows a signicant reduction in
the micropore volume for the partially crystalline silicalite-1 based
samples with the increase of arenesulfonic acid group. The rela-
tively low micropore volume was due to the pore opening by the
acid group and the presence of the amorphous materials [48].
These changes were consistent with the reduction in the crystal
size, as shown in the SEM pictures [27].
4. Conclusion
Partially crystalline silicalite-1 based materials was successfully
synthesized by co-condensation of inorganic silica (TEOS) and or-
ganic silica sources (PE) under hydrothermal synthesis, followed
by sulfonation of aromatic rings which grafted on amorphous
phases, to arenesulfonic acid group. The partially crystalline silica-
lite-1 based materials were studied by a variety of chemical and
physical characterization techniques. XRD results show that the
crystallinity of the partially crystalline silicalite-1 based samples
reduced gradually, until highly amorphous phase was formed at
30 mol%, indicated that the crystalline materials can only be
formed from an organosilane source not more than 20 mol%. As
characterized by XRD, SEM and TEM, the surface morphology of sil-
icalite-1 was signicantly modied, leading to a smaller crystal
present in the amorphous phases. The combined interpretation of
the characterization data from
29
Si MAS,
13
C CP-MAS solid state
NMR, FT-IR and elemental analysis support the presence of the acid
group in the amorphous phases. Nitrogen adsorptiondesorption
analysis shows that, the presence of acid group in the amorphous
structure causing the reduction in the surface area, and micropore
volume. These partially crystalline silicalite-1 based materials
exhibited high acid capacity with a good thermal stability, up to
550 C. The present research shows a potential way to create tun-
able zeolite materials with new properties for better separations
and catalytic activities, as well as for other applications.
Acknowledgments
The nancial support provided by Ministry of Science, Technol-
ogy and Environment under e-Science Fund Grant (Account No:
Table 10
Surface area and porosity characteristics of parent silicalite-1 and partially crystalline
silicalite-1 based samples.
Sample S
BET
, m
2
g
1
Total pore
volume
a
,
V
tot
cm
3
g
1
Micropore
volume
b
,
V
mic
cm
3
g
1
Average pore
diameter,
d
p
nm
SIL-1 329 0.18 0.08 1.5
SILPE5SO
3
H 326 0.66 0.04 7.9
SILPE10SO
3
H 322 0.41 0.04 6.7
SILPE15SO
3
H 250 0.31 0.03 4.8
SILPE20SO
3
H 244 0.26 0.03 4.1
a
At P/P
o
0.99.
b
t-Plot method.
138 Y.F. Yeong et al. / Microporous and Mesoporous Materials 123 (2009) 129139
6013319), Universiti Sains Malaysia under Short Term Grant (Ac-
count No: 6035188), Ministry of Higher Education under FRGS (Ac-
count No: 6070021) and Research University Grant (Account No:
811043) are duly acknowledged.
References
[1] G. Bonilla, D. Vlachos, M. Tsapatsis, Micropor. Mesopor. Mater. 42 (2001) 191.
[2] J. Hedlund, J. Sterte, M. Anthonis, A. Bons, B. Carstensen, N. Corcoran, D. Cox, H.
Deckman, W. Gijnst, P. Moor, F. Lai, J. Mchenry, W. Mortier, J. Reinoso, J. Peters,
Micropor. Mesopor. Mater. 52 (2002) 179.
[3] C.P. Nicolaides, N.P. Sincadu, M.S. Scurrell, Catal. Today 71 (2002) 429.
[4] C.P. Nicolaides, Appl. Catal. A: Gen. 185 (1999) 211.
[5] N.Y. Chen, T.F. Degnan, Chem. Eng. Prog. 84 (1988) 32.
[6] J. Dwyer, A. Dyer, Chem. Ind. 2 (1984) 237.
[7] K.S. Triantafyllidis, L. Nalbandian, P.N. Trikalitis, A.K. Ladavos, T.
Mavromoustakos, C.P. Nicolaides, Micropor. Mesopor. Mater. 75 (2004) 89.
[8] A.S.M. Chong, X.S. Zhaoa, A.T. Kustedjo, S.Z. Qiao, Micropor. Mesopor. Mater. 72
(2004) 33.
[9] X. Wang, J.C.C. Chan, Y.H. Tseng, S. Cheng, Micropor. Mesopor. Mater. 95 (2006)
58.
[10] X. Wang, S. Cheng, J.C.C. Chan, J.C.H. Chao, Micropor. Mesopor. Mater. 96
(2006) 321.
[11] W. Zhang, X. Bao, X. Guo, X. Wang, Catal. Lett. 60 (1999) 89.
[12] R. Van Grieken, J.L. Sotelo, J.M. Menendez, J.A. Melero, Micropor. Mesopor.
Mater. 39 (2000) 135.
[13] P.A. Jacobs, E.G. Derouane, J. Weitkamp, J. Chem. Soc. Chem. Commun. (1981)
591.
[14] C.P. Nicolaides, H.H. Kung, N.P. Makgobaa, N.P. Sincadu, M.S. Scurrell, Appl.
Catal. A: Gen. 223 (2002) 29.
[15] Y. Shin, S.Z. Thomas, E.F. Glen, L.Q. Wang, L. Jun, Micropor. Mesopor. Mater. 37
(2000) 49.
[16] L.M. Yang, Y.J. Wang, G.S. Luo, Y.Y. Dai, Micropor. Mesopor. Mater. 84 (2005)
275.
[17] C.W. Jones, K. Tsuji, M.E. Davis, Micropor. Mesopor. Mater. 33 (1999) 223.
[18] C.W. Jones, M. Tsapatsis, O. Tatsuya, M.E. Davis, Micropor. Mesopor. Mater. 42
(2001) 21.
[19] K. Tsuji, C.W. Jones, M.E. Davis, Micropor. Mesopor. Mater. 29 (1999) 339.
[20] B.L. Su, M. Roussel, K. Vause, X.Y. Yang, F. Gilles, L. Shi, E. Leonova, M. Eden, X.
Zou, Micropor. Mesopor. Mater. 105 (2007) 49.
[21] Z. Lai, M. Tsapatsis, J.P. Nicolich, Adv. Funct. Mater. 14 (2004) 716.
[22] C.W. Jones, Science 300 (2003) 439440.
[23] J. Guo, A.J. Han, H. Yu, J.P. Dong, H. He, Y.C. Long, Micropor. Mesopor. Mater. 94
(2006) 166172.
[24] I.K. Mbaraka, D.R. Radu, S.Y. Lin, B.H. Shanks, J. Catal. 219 (2003) 329.
[25] B.A. Holmberg, S.J. Hwang, M.E. Davis, Y. Yan, Micropor. Mesopor. Mater. 80
(2005) 347.
[26] Y. Zheng, X. Su, X. Zhang, W. Wei, Y. Sun, Stud. Surf. Sci. Catal. 156 (2005)
205.
[27] D.P. Serrano, J. Aguado, J.M. Rodrguez, A. Peral, From zeolites to porous MOF
materials, in: The 40th Anniversary of International Zeolite Conference, 2007,
pp. 282288.
[28] S.C. Larsen, J. Phys. Chem. C 111 (2007) 18464.
[29] S. Tanaka, C. Yuan, Y. Miyake, Micropor. Mesopor. Mater. 113 (2008) 418.
[30] K. Yamamoto, T. Tatsumi, Chem. Mater. 20 (2008) 972.
[31] C.H. Cheng, T.H. Bae, B.A. McCool, R.R. Chance, S. Nair, C.W. Jones, J. Phys.
Chem. C 112 (2008) 3543.
[32] U. Diaz, A. Jose, V. Moya, A. Corma, Micropor. Mesopor. Mater. 93 (2006)
180.
[33] D.P. Serrano, J. Aguado, J.M. Escola, J.M. Rodriguez, A. Peral, Chem. Mater. 18
(2006) 2462.
[34] M. Rat, M.H. Zahedi-Niaki, S. Kaliaguine, Trong-On Do, porous MOF
materials, in: The 40th Anniversary of International Zeolite Conference,
2007, p. 1811.
[35] Y.H. Han, A. Taylor, M.D. Mantle, K.M. Knowles, J. Non-Cryst. Solids 353 (2007)
313.
[36] Y. Wang, Y. Tang, A. Dong, X. Wang, N. Ren, Z. Gao, J. Mater. Chem. 12 (2002)
1812.
[37] Z. Olejniczak, M. e czka, K. Cholewa-Kowalska, K. Wojtach, M. Rokita, W.
Mozgawa, J. Mol. Struct. 744 (2005) 465.
[38] X. Wang, S. Cheng, J.C.C. Chan, J. Phys. Chem. C 111 (2007) 2156.
[39] X. Zhang, Y. Wu, S. He, D. Yang, Surf. Coat. Technol. 201 (2007) 6051.
[40] Y.S. Ooi, S. Bhatia, Micropor. Mesopor. Mater. 102 (2007) 310.
[41] B. Samuneva, P. Djambaski, E. Kashchieva, G. Chernev, L. Kabaivanova, E.
Emanuilova, I.M.M. Salvado, M.H.V. Fernandes, A. Wu, J. Non-Cryst. Solids 354
(2008) 733.
[42] Y. Luo, P. Yang, J. Lin, Micropor. Mesopor. Mater. 111 (2008) 194.
[43] R. Xing, Y. Liu, Y. Wang, L. Chen, H. Wu, Y. Jiang, M. He, P. Wu, Micropor.
Mesopor. Mater. 105 (2007) 41.
[44] C.S. Gill, B.A. Price, C.W. Jones, J. Catal. 251 (2007) 145.
[45] F. Rouquerol, J. Rouquerol, K. Sing, Adsorption by Powders and Porous Solids
Principles, Methodology and Applications, Elsevier, Amsterdam, 1999.
[46] R.M.A. Roque-Malherbe, Adsorption and Diffusion in Nanoporous Materials,
CRC Press, Boca Raton, 2007.
[47] D.P. Serrano, G. Calleja, J.A. Botas, F.J. Gutierrez, Sep. Purif. Technol. 54 (2007)
1.
[48] Y.H. Chou, C.S. Cundy, A.A. Garforth, V.L. Zholobenko, Micropor. Mesopor.
Mater. 89 (2006) 78.
[49] S. Shylesh, S.P. Mirajkar, A.P. Singh, J. Mol. Catal. A: Chem. 212 (2004) 219.
[50] I.K. Mbaraka, B.H. Shanks, J. Catal. 244 (2006) 78.
Y.F. Yeong et al. / Microporous and Mesoporous Materials 123 (2009) 129139 139

Vous aimerez peut-être aussi