You are on page 1of 14

Microporous and Mesoporous Materials 201 (2015) 176189

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

Synthesis and application of ZSM-5/SAPO-34 and SAPO-34/ZSM-5


composite systems for propylene yield enhancement in propane
dehydrogenation process
Marjan Razavian, Shohreh Fatemi
School of Chemical Engineering, College of Engineering, University of Tehran, P.O. Box 11365 4563, Tehran, Iran

a r t i c l e

i n f o

Article history:
Received 6 July 2014
Received in revised form 26 August 2014
Accepted 7 September 2014
Available online 16 September 2014
Keywords:
SAPO-34/ZSM-5
ZSM-5/SAPO34
Composite zeolites
Propane dehydrogenation
Propylene

a b s t r a c t
Two types of zeolitic composite systems with binary hierarchical structures comprising ZSM-5 and SAPO34 molecular sieves were synthesized employing different procedures. Obtained products were served as
catalytic carriers for propane dehydrogenation reaction so as to promote the physicochemical properties
of ZSM-5 support, enhance the propylene yield and reduce the formation of light compounds. ZSM-5/
SAPO-34 was fabricated in a series hydrothermal procedure employing pre-heated ZSM-5 suspension
followed by secondary growth of SAPO-34 layer whereas SAPO-34/ZSM-5 was synthesized using tetrapropylammonium bromide exchanged-SAPO-34 crystals and pre-reacted ZSM-5 slurry in a hydrothermal
one-step process. The products were characterized by XRD, FESEM, EDS, FTIR, NH3-TPD, and N2 adsorptiondesorption techniques to investigate the textural and structural properties of composite architectures. The catalytic performance of the bimetallic Pt-Sn-based composites were evaluated in propane
dehydrogenation reaction and compared with those of physical mixture and single ZSM-5-derived
catalysts. Either of composites represented improved catalytic performance due to the synergetic effect
between ZSM-5 and SAPO-34, which promoted the catalytic properties of the samples. Catalytic reactivity of the composite catalysts was strongly dependent on the synthesis method and employed zeolite/
zeotype ratio. Best result was acquired for PtSn-based SAPO-34/ZSM-5 (Si/Al = 60) brand-new efcient
composite with improved stability, boosted conversion and signicant selectivity towards light olens,
propylene in particular.
2014 Elsevier Inc. All rights reserved.

1. Introduction
Zeolitic molecular sieves, with uniform three-dimensional
nano-channels containing different portion and strength of acid
sites are inevitable members of catalytic and separation processes
[13]. It is well understood that both internal channels and
external crystal surfaces with different acid strengths and shape
sieving behaviors contribute in chemical reactions. Although external surface active sites are low in concentration, their role in
catalytic reactions is not negligible bringing in mind the diffusional
restrictions for reactants to reach active sites in the micropores [4].
Therefore, recent endeavors have been focused on the modication
of external surface area of zeolites to enhance their reactivity and
selectivity properties. The functionalization of the external surface
of zeolites by different techniques such as acid leaching, steaming
and element surface modication has been applied to induce
Corresponding author. Tel.: +98 21 61112229; fax: +98 21 669667784.
E-mail address: shfatemi@ut.ac.ir (S. Fatemi).
http://dx.doi.org/10.1016/j.micromeso.2014.09.022
1387-1811/ 2014 Elsevier Inc. All rights reserved.

mesopores and obtain a reasonable acid distribution [57].


However, partial destruction and species deposition may lead to
a reduction in the effective diameter of pore openings and consequently fall the catalytic activity.
Another strategy for modifying the surface structure in order to
manipulate the catalytic properties in a desired way is the growth
of a continuous shell over core crystals. In this case, access to a core
with specic properties is controlled by a shell which gives the
composite high stability and functionality [8,9]. Bouizi et al. synthesized a series of microcomposites of different zeolite structure
types including SOD-LTA (the rst zeolite type represents the core
and the second the shell material), BEA-LTA, FAU-MFI, MFI-BEA,
BEA/MFI and MFI/MFI (silicalite-1/ZSM-5 and the reverse)
[10,11]. The incompatibility between core material and shell
precursor mixture in preparing several types of composites was
circumvented successfully by the preliminary seeding of core crystals. However, the formation of core/shell structure in some cases
like SOD/LTA remained a problem due to the compositional and
crystallization elds incompatibilities between two structures.

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

Therefore, the foremost importance of chemical compatibility and


structural similarity of shell and core material, as well as the overlapping of crystallization conditions in producing an integrated
shell around single crystals is an indispensable matter of fact
[12,13]. Consequently, the successful synthesis of only few zeolitic
compounds like MFI/MEL [14] composed of identical building units
with different spatial arrangements, MFI/MFI [15,16] with similar
building units and crystallization conditions, BEA/MFI [11] with
entirely different structures and MOR/MFI [13,17] with a small
overlapping between framework compositions and crystallization
elds have been reported thus far. Above described samples exemplify the partial or complete overgrowth of a zeolitic material by
another one normally taking place in a two-step synthesis methodology. There are two other important categories of composite
materials, namely intergrowth as the result of intimate co-crystallization of zeolite materials with hybrid crystals and epitaxial
growth phenomenon with specic selective orientation on a
diversity of crystal faces which are usually obtained in a one-step
synthesis technique. ERI/OFF [18], MFI/MEL [19], MAZ/MOR (ERC1) [20], and BEA/MOR [21] are some frequent examples of intergrowth materials with similar planar building units. Besides, such
a heteroepitaxial growth was reported by Wakihara et al. for a
continuous chabazite lm on a sodalite substrate with a patterned
surface texture [22].
In spite of above described references, reports on the synthesis
and application of aluminosilicate and silicoaluminophosphate
binary structure composites are limited. Fan et al. [14] synthesized
ZSM-5/SAPO-11 in a form of core/shell structure by means of
in situ crystallization of SAPO-11 on ZSM-5 with modied dual
porosity and acidity which showed good desulfurization activity
and high liquid yield in hydro-upgrading FCC gasoline. Very
recently, the preparation of a ZSM-5/SAPO-5 core/shell structure
zeolite as FCC catalyst was reported applying a phosphorous precoating method [23]. Binary structure ZSM-5/SAPO-34 zeolitic
composite has been fabricated and developed for MTO [24] and
ethanol to propylene reactions [25]. In rst case, series composite
which was obtained through a consecutive hydrothermal crystallization procedure showed higher selectivity towards light olens
while in second one, physical mixture of individual ZSM-5 and
SAPO-34 showed the highest propylene yield. Performed synthesis
technique applying HZSM-5 powder and SAPO-34 precursor gel in
a one-step hydrothermal crystallization was not successful in
producing a composite with improved performance.
As far as we are concerned, no specic bi-phase composite
structure has been recommended for propane dehydrogenation
(PDH) reaction for the selective production of light olens in a
controlled trend; highlighting propylene as the major product.
In recent years, the growth in propylene demand exceeded the
growth in ethylene, opening up other technology opportunities
focused primarily on propylene. Generally, the worldwide
demand for propylene is well supplied from steam crackers and
FCC units providing propylene as a byproduct [26]. However,
the annual growing demand of propylene cannot meet the significant change of market, bringing in mind the instability of crude
oil price as well. PDH involves catalytic removal of hydrogen
from propane to yield propylene which has been proposed as a
great on-purpose alternative with high potential for covering
the major propylene deciency [27,28]. But it suffers from equilibrium limitation and thermodynamical constraints. It is crucial
to formulate a catalyst exhibiting high catalytic properties and
stability that can decelerate side reactions under severe conditions. PtSn-based zeolite Socony Mobil#5 (ZSM-5) catalyst is a
hydrothermally stable catalyst with good reactivity but low propylene yield since the present strong acid sites in its structure
transform propylene to heavy bulky products [2931]. It seems
that contribution of a system with high quality sieving network

177

like SAPO-34 can alleviate the aforementioned problem and


increase propylene yield [32,33].
In this contribution, synthesis of ZSM-5/SAPO-34 binary nanostructured composite by means of a sequential crystallization
method is reported. Additionally, an efcient convenient pathway
is proposed to preserve SAPO-34 crystals in the strong basic precursor media of ZSM-5 to produce a novel bi-phase SAPO-34/
ZSM-5 zeolite composite. Despite SAPO-34 and ZSM-5 zeolites
have neither similar framework compositions nor chemical
compatibilities, pre-treatment step accomplished to boost up
SAPO-34 robustness in the synthesis media. Combination of useful
physicochemical properties of these two types of zeolites with
structural codes of CHA and MFI into a hierarchical composite
can improve surface properties and lead to a modied distribution
of active centers. In this way a great organization of catalytic,
sieving and adsorptive functions of both anking structures is
combined that can increase the application area of these materials.
The inuence of properties on the catalytic performance of PtSnbased composite catalysts are compared with those of single ZSM5 and physical mixture of individual ZSM-5 and SAPO-34 in the
PDH chemical reaction to convert propane to highly valuable propylene as one of the most important petrochemical building
blocks.
2. Experimental
2.1. Materials
Tetraethyl orthosilicate (TEOS, 98%), aluminum isopropoxide
(AlP, 98%), polyethylene glycol (PEG, MW = 4000), tetraethylammonium hydroxide (TEAOH), wt. 20%), tetrapropylammonium bromide (TPABr, 99%), diethylamine (DEA, wt. 99%), ortho-phosphoric
acid (H3PO4, wt. 85%), hydrochloric acid fuming (HCl, wt. 37%),
nitric acid (HNO3, wt. 65%), Platinum(IV) chloride (PtCl4, 57.5%
Pt), Tin(II) chloride (SnCl2), 2-propanol (C3H8O) and sodium
hydroxide (NaOH pellets, 98%) were purchased from Merck
company. Colloidal silica (SiO2, wt. 30%), tetraethylammonium
hydroxide (TEAOH, wt. 35%) were obtained from Aldrich. Sodium
aluminate (NaAlO2, 54% Al2O3 and 41% Na2O) was bought from Riedel-de-Han company. Commercial PtSnK/cAl2O3 catalyst
(Pt = 0.6 wt.%, Sn = 0.8 wt.%, K = 0.8 wt.%, surface area = 200 m2/g,
total pore volume = 0.65 cm3/g) in form of spheres was supplied
by Procatalyse company.
2.2. Synthesis procedures
2.2.1. Preparation of ZSM-5/SAPO-34 composite zeolite
In this section two series of composites referred as ZS1 and ZS2
were greatly synthesized. In the rst case, ZSM-5/SAPO-34 binary
composite synthesis was carried out following a sequential crystallization. Firstly, ZSM-5 core crystals were synthesized according to
the seeding method applying silicalite-1 seeds as SDA (structure
directing agent). A template-free aluminosilicate precursor gel
was prepared by the hydrolysis of sodium aluminate and colloidal
silica as Si and Al sources with the molar composition of 60 SiO2:1
Al2O3:4.2 Na2O:1200 H2O. Sodium hydroxide was employed as the
alkali agent to attain the suitable PH of synthesis media (PH 12)
and thereafter a certain amount of silicalite-1 (equal to 5% of total
silica amount) was added to the reaction mixture. The homogenous gel was moved to a 150 ml stainless steel autoclave and
heated at 170 C for 24 h to crystallize. In the second step, SAPO34 precursor gel with molar ratio 1 Al2O3:0.5 SiO2:0.8 P2O5:0.1
HCl:1.8 DEA:0.2 TEAOH:0.05 PEG:60 H2O was obtained using TEOS,
AlP and H3PO4 as Si, Al and P sources, respectively. TEAOH and DEA
mixture was employed as structure directing agent and PEG

178

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

polymer acted as co-template. Appropriate quantity of ZSM-5


slurry and SAPO-34 gel were mixed together; thereafter, poured
into the autoclave and subjected to hydrothermal treatment at
130 C and 200 C for 5 and 12 h, respectively. The considered crystallization condition was similar with that of individual SAPO-34.
Finally, as-synthesized product crystals were recovered after
several cycles of centrifuging and washing with deionized water,
followed by drying and calcining at 550 C for 5 h. Second sample
was fabricated based on the idea of covering ZSM-5 core crystals
with a ne thin layer of SAPO-34 small crystals. The whole synthesis procedure was repeated except that the TEAOH aqueous
solution (wt.35%) and DEA with molar ratio of 1.8:0.2 were used
as SDA, SAPO-34 precursor gel was aged 24 h under vigorous stirring before ZSM-5 pre-heated slurry was added and crystallization
was performed under dynamic condition. The synthesis procedure
is schematically depicted in Fig. 1.
2.2.2. Preparation of SAPO-34/ZSM-5 composite zeolite
Two series of SAPO-34/ZSM-5 novel composites noted as SZ1
and SZ2 with regard to different core crystals content and applied
method of crystallization were prepared by one-step crystallization of ZSM-5 pre-reacted synthesis gel and laboratory made
pre-treated SAPO-34 powder. The synthesis procedure scheme is
demonstrated in Fig. 2. In a typical synthesis, SAPO-34 powder
was rstly synthesized from a reaction mixture with the following
molar composition 1 Al2O3:0.5 SiO2:0.8 P2O5:0.1 HCl:1.8 DEA:0.2
TEAOH:0.05 PEG:60 H2O. Similar organic amines-water system
using TEOS, AlP and H3PO4 as Si, Al and P sources was employed.
The precursor gel was hydrothermally treated at 130 C and
200 C for 5 and 12 h, respectively. The nal solid product was
recovered by centrifugation and multiple washing. Crystals were
dried at 100 C and calcined at 550 C C for 5 h to remove organic
template backbone from material structure. Finally the calcined
powder was ion-exchanged with 1 M tetrapropylammonium bromide propanolic solution twice at room temperature. The product
is noted as TPA-SAPO-34.
As the second step, ZSM-5 mixture gel was prepared by the
same procedure reported in Section 2.2.1. The initial gel was stirred
at room temperature and then transferred into an autoclave and
heated at 170 C for short 10 h as pre-crystallization step. Then, a
specied amount of TPA-SAPO-34 was added to the pre-reacted
ZSM-5 slurry under continuous stirring. After sufcient stirring,
the resulting mixture was transferred into the autoclave and

Fig. 2. Scheme of SAPO-34/ZSM-5 composite overall synthesis procedure.

treated at 170 C for 5 h under static condition. The solid as-synthesized product was recovered by centrifugation, washed with
distilled water and re-suspended in deionized water repeatedly.
Finally the sample was dried and calcined at 550 C for 5 h.
Another sample with higher content of SAPO-34 core crystals
was fabricated applying dynamic condition to enhance the
uniformity of combined composite crystals. Full details of the
batch composition and synthesis condition of each sample are
specied in Table 1.
For comparison, a physical mixture with equal mass percent of
SAPO-34 and ZSM-5 (Si/Al2 = 60, molar ratio) was prepared by fully
blending of both materials carefully.
2.2.3. Preparation of the catalysts
The bimetallic PtSn-based catalysts were prepared by sequential wet impregnation at 65 C with very dilute acidic solution of
SnCl2 and aqueous solution of PtCl4 with 0.03 and 0.0133 M
concentration, respectively. 0.9 M nitric acid was employed as a
catalyst to facilitate the solution of SnCl2 in water. Each impregnation step was followed by drying and calcination at 500 C for 6 h.
The nominal metal loading was 0.5% of Pt and 1% of Sn.
2.3. Characterization

Fig. 1. Scheme of ZSM-5/SAPO-34 composite overall synthesis procedure.

The X-ray powder diffraction (XRD) patterns of samples were


recorded at ambient temperature using a Philips Xpert system in
the 2h range of 550 using Cu-Ka1 radiation (k = 1.54056 ). A
Zeiss Sigma VP eld-emission electron microscope was used to
capture FE-SEM images. The elemental compositions were
characterized by Oxford Instruments energy dispersive X-ray
spectrometer (EDS). Temperature-programmed desorption (TPD)
measurements were carried out on conventional apparatus using
Micromeritics chemisorb 2750 analyzer. About 0.15 g of sample
was placed in a quartz reactor and saturated with ammonia at
room temperature. TPD was carried out from 50 to 650 C with a
heating rate of 10 C/min and nitrogen as the carrier gas. N2

179

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189


Table 1
Synthesis conditions for different composite samples.
Sample

ZS1
ZS2

SZ1
SZ2

Shell/core content

Mixing time (h)

Aging time (h)

Volume ratio

ZSM-5 suspension and SAPO-34 precursor gel

SAPO-34 precursor gel

1
6

1
2

0
24

Weight ratio

TPA-SAPO-34 powder and pre-reacted ZSM-5 slurry

ZSM-5 precursor gel

5.5
1.5

1
2

1
1

Crystallization technique

TEAOH/DEA molar ratio

Static condition
Dynamic condition (8.5 rpm)

1/9
9

Static condition
Dynamic condition (8.5 rpm)

1/9
1/9

adsorptiondesorption isotherms were determined in a Micromeritics Tristar-3000 porosimeter. Textural properties including total
surface area, total pore volume, micropore volume and average
pore width were measured using BET, t-plot and BJH (Barrett
JoynerHalenda) methods. Before the surface area measurements,
samples were degassed at 300 C for 3 h. FT-IR spectrum was measured using KBr-diluted pellet on an IR spectrophotometer
(PerkinElmer).
2.4. Catalytic evaluation
Propane dehydrogenation was carried out in a xed-bed tubular
quartz micro reactor with a propane and hydrogen mixture as feed.
0.5 g catalyst was placed into the center of reactor and was heated
up to 500 C and maintained at this temperature under nitrogen
ow for 2 h. Reaction conditions including temperature, pressure,
weight hourly space velocity (WHSV), and H2/C3H8 molar ratio
were xed on 600 C, 0.88 atm, 4 h1 and 0.8. Finally products
were analyzed with an online YL-6100 GC gas chromatograph
equipped with a HID detector. Propane conversion, product
selectivities, and product yield were dened as follows:

P
Propane Conversion

C iout  C propaneout
P ni
 100
i 3 C iout

ni
3

C iout
 100
C iout  C propaneout

conversion %  selectivity %
100

Selectivity P

Yield %

ni
i 3

ni
i 3

where i included all the components containing carbon atoms in the


exit gas stream, ni was the number of carbon atoms of component i
and Ci was its molar composition.
3. Results and discussion
3.1. Composite samples characterization
XRD spectrograms of prepared composite molecular sieves
along with reference components i.e. ZSM-5 and SAPO-34 are
represented in Fig. 3 to detect the crystalline structure of the samples. As it can be observed, all typical characteristic peaks of both
MFI (2h 7.9, 8.8, 23.1, 23.9, and 24.4) and CHA (2h 9.6,
13, 20.8 and 30.1) structures are present in all binary composite
except ZS2. In SZ series, this fact shows that SAPO-34 crystals are
preserved to a certain extent in the synthesis media. However,
the CHA reections intensities depend on the quantity of employed
SAPO-34 crystals. As it can be observed, the CHA major peaks are
very smoothed in SZ1 XRD pattern because of the low amount of
applied SAPO-34 in the synthesis procedure whereas they tended
to increase in SZ2 sample. CHA major peaks are buried in the massive MFI signals in ZS2 pattern. This sample was supposed to contain a very thin uniform layer of SAPO-34 around high siliceous

Fig. 3. XRD patterns of synthesized composites as well as pristine SAPO-34 and


ZSM-5 as references: (a) SAPO-34, (b) SZ1, (c) SZ2, (d) ZS1, (e) ZS2 and (f) ZSM-5.

ZSM-5 core crystals. It seems that this phenomenon has made


CHA characteristics reections to get imperceptible. Sometimes
lms are too thin to provide diffraction patterns with accurate
relative intensities using a conventional diffractometer with
BraggBrentano parafocusing geometry [34,35]. Additionally,
XRD peaks become broadened or disappear normally when the
order of domain area is decreased in tiny shell thickness [36]. Other
alternatives like glancing-incidence diffraction sounds better to
reduce substrate scatter and increase lm diffraction peaks intensities [3537]. More thorough inspection using spectrograms
shows a slight shift towards high 2h relative to the corresponding
peaks of individual ZSM-5 and SAPO-34. Therefore, the chemical
interaction between ZSM-5 and SAPO-34 can be inferred in combinational zeolite structures. Also no other unidentied phase in XRD
patterns is observed. Overall, the diffraction peaks of binary composites are weaker compared to the pristine ZSM-5 and SAPO-34
based on the shielding effect of materials surrounding core seeds.
Employing uncoated SAPO-34 crystals in the synthesis process
resulted in the formation of a fully amorphous material. XRD pattern (Fig. 4a) suggests complete dissolution of SAPO-34 crystals
within ZSM-5 basic reaction mixture (PH12). Even lighter basic
synthesis media of silicalite-1 (PH10) could not afford proper
conditions to stabilize the SAPO-34 crystals and has led to the partial dissolution of SAPO-34 (Fig. 4b). Low stability of SAPO-34 crystals due to the chemical incompatibility between SAPO-34 and
ZSM-5 phases has brought seed detachment, phase transformation,
and destruction. Framework density is believed to be closely
related to the stability of zeolitic materials [38]. Structure with
low framework density contains channels outlined by large rings
with low degree of branching of the framework and high energy
factor. Materials with high energy factors normally show low
stability and can be easily transformed under hydrothermal conditions. Consequently, CHA-type structure with low framework
density (15.1 T/1000 3) has been unstable in basic condition while
MFI-type structure with framework density of 18.4 T/1000 3 is
much denser and showed high stability in mild basic SAPO-34

180

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

Fig. 4. XRD patterns of synthesized (a) SAPO-34/ZSM-5 and (b) SAPO-34/silicalite-1


composite samples employing uncoated SAPO-34 crystals.

synthesis media (PH 9) [39]. Moreover, dissolution of SAPO-34


crystals probably produced enriched phosphorous, aluminum and
silicon based species which alter initial stoichiometry of shell

composition. So, chemical equilibrium at shell solution-core crystal


interface has changed so that adequate crystallization condition for
secondary growth was not achieved. Therefore it can be concluded
that core pre-treatment is a vital step so as to make the secondary
growth overtakes core dissolution and shell forms rapidly through
a non-equilibrium state. It seems that pre-treatment step is an
efcient route that can be applied in the synthesis of different composite structures utilized with SAPO-34 core crystals to circumvent
structural or chemical incompatibilities.
FESEM micrographs of SAPO-34 material and resultant SAPO34/ZSM-5 binary composite (SZ1) are shown in Fig. 5. Close inspection reveals the difference between composite and core phase
morphologies. It seems that cubic SAPO-34 crystals are covered
by plate like poly-crystalline ZSM-5 particles. However, presence
of huge number of isolated ZSM-5 particles in media due to the
large amount of employed ZSM-5 seeds made the diagnosis of
SAPO-34 crystals coverage hard. Distinct irregular particles in
Fig. 6 imply that formation of SAPO-34 layer around ZSM-5 particles was not achieved and two separate phases were formed. Based
on the occurrence of typical characteristic peaks of CHA structure
in the ZS1 XRD pattern, it seems that both SAPO-34 and ZSM-5
phases formed and grew without a strong interaction between
them. This pattern might be caused since SAPO-34 crystal nuclei
were too big to be able to attach and pile up on ZSM-5 crystals surfaces. Besides the lack of aging time for SAPO-34 precursor gel and
static crystallization condition were conducive to the formation of
big discrete SAPO-34 particles. The morphology might be nevertheless inuenced by partial extraction of ZSM-5 elements that
changed the nal SAPO-34 reaction mixture composition. High
molar ratio of TEAOH template, a long period aging time and
dynamic crystallization condition were then applied to fabricate

Fig. 5. FESEM images of (a) SAPO-34 molecular sieve and (b and c) SZ1 composite.

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

181

Fig. 6. FESEM images of (a) ZSM-5 molecular sieve and (b) ZS1 composite.

Fig. 7. FESEM images of (a) pre-treated SAPO-34 material and (be) SZ2 composite.

a thin integrated SAPO-34 layer around ZSM-5 seeds with low


crystalline ordering and enhance the uniformity avoiding SAPO34 gel accumulation and separate second phase arrangement.
FESEM images of SZ2 composite sample containing high content
of SAPO-34 material as well as pre-treated SAPO-34 powder with
TPABr alcoholic solution are represented in Fig. 7 with different
magnications. First photograph suggests huge change in the
morphology of SAPO-34 cubic crystals towards semi-spherical
big particles with rough surfaces being covered with TPA+ species.
Although the pre-treatment step for coating SAPO-34 crystals
altered morphological characteristics of these crystals, their structural properties were largely preserved during the pre-treatment
process according to the EDS result (Table 2). Fig. 7(be) attests
the attachment and stacking of ZSM-5 crystals on the surface of
SAPO-34 pre-treated particles and thereby producing a combinational joined-up system so as to organize catalytic, sieving and/or

adsorptive functions of both anking SAPO-34 and ZSM-5 zeolitic


structures. But is has to be pointed out regarding Fig. 7(e) that
several SAPO-34 particles could be selectively coated by ZSM-5
crystals wherein only specic faces were covered. FESEM
micrographs of ZS2 sample equipped with ZSM-5 core material
are displayed in Fig. 8. The morphology of the synthesized composite with pseudo-spherical coarse aggregates is close to the ZSM-5
plate like nanocrystals (100  500 nm) being packed to smoothed
pseudo-hexagonal prisms. A close view of single particle is shown
in Fig. 8(b).
Furthermore, SZ2 and ZS2 binary composites were selected as
optimum samples to undergo more characterizing analyses. EDS
technique was conducted to investigate the elemental composition
of resultant materials to validate further the presence of both
parent zeolite types. Since EDS analysis covers a surface much larger than the outer layer and thus is certainly affected by the core

182

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

Table 2
Elemental compositions of composites, pre-treated SAPO-34, and pure ZSM-5
molecular sieves obtained by EDS analysis.
Sample

TPA-SAPO-34
SZ2
ZS2
ZSM-5

Composition (wt.%)
Si

Al

3.9
33.1
37
54.8

15.9
19.7
6.4
2.6

13.8
16.4
2.5


Si/Al (molar ratio)

P/Al (molar ratio)

0.24
1.6
5.5
20.24

0.76
0.72
0.34


material, it was exploited to study the inuence of core material on


the surface layer composition [10,17]. The results are listed in
Table 2. Evidently, composites showed different Si/Al and P/Al
ratios from individual ZSM-5 and SAPO-34 denoting a strong interaction between two zeolitic structures. In case of ZS2 sample partial migration of non-framework aluminum to the zeolite surface is
probable owing to the insertion of ZSM-5 seeds in mild alkaline
media of SAPO-34 precursor gel. Abello et al. reported a porosity
development procedure for ZSM-5 upon selective aluminum
extraction using bulky tetraalkylammonium cations [40]. Similar
procedure was studied by Yingping et al. [41] for the modication
of ZSM-5 catalytic performance in MTG reaction. Therefore, partial
dealumination assumption can be the reason explaining lower Si/
Al and P/Al ratios. Elemental mapping of either of composite
zeolites are depicted in Figs. 9 and 10 in which homogenous
dispersion of main Si, Al, and P elements can be visualized. Summarizing these data, it can be said that required ZSM-5 and
SAPO-34 layers on SZ2 and ZS2 surface composites were formed
satisfactorily.
Infrared spectra of pristine and selected composite samples i.e.
SAPO-34, ZSM-5, SZ2 and ZS2 zeolites in the vibration frequency
region of 4004000 cm-1 are shown in Fig. 11. Possible assignment
of various vibration frequencies are gathered in Table 3, based on
the relevant reports [4245]. In terms of the variety of the mentioned IR bands, we can gain some useful information on the
framework structure of combinational molecular sieves. Both composite IR patterns comprise major characteristic bands which are
attributed to the SAPO-34 and ZSM-5 structures. Detected bands
in the range of 4001300 cm1 and 34003800 are because of
framework vibrations and bridging hydroxyl acidic sites stretching,
respectively.
According to the data provided in Table 3, it seems that SZ2
bi-phase composite which was equipped with SAPO-34 crystals
closely resembles SAPO-34 properties while ZS2 binary composite
with ZSM-5 acting as the core materials is akin to ZSM-5 structure.
At this juncture it is important to mention that the contribution of
CHA structure to the nal ZS2 structure is highlighted by the

appearance of two intense absorption bands at 1099 and


3437 cm1 (SAPO-34 characteristic bands) assigned to the OPO
groups and Bronsted acid centers and diminishing 3654 cm1 distinct band (ZSM-5 characteristic band) which is associated with OH
groups connected with extra lattice Al atoms that possess lower
acid strength compared to structural OH groups. This is denotative
of the atomic ordering of SAPO-34, even though its ordering was
not enough to be detected by X-ray diffraction method. It is worth
mentioning that observed IR bands shifts in composite samples are
due to the combination of the SAPO-34 and ZSM-5 frameworks by
chemical bonds which further conrm the idea of strong interaction between these two contributing phases. This result is consistent with that of XRD characterization.
Basic physicochemical and textural properties of obtained
products are summarized in Table 4. BET surface area and pore volume of either of composite zeolitic molecular sieves are within the
represented surface area and pore volume ranges of parent ZSM-5
and SAPO-34. ZS2 sample possesses the highest external surface
area which is tentatively attributed to the modication of its
surface based on the presence of thin SAPO-34 layer with small
crystals. Also its higher mesopore volume suggests higher relative
mesoporosity of the hierarchical ZSM-5/SAPO-34 composite that is
related to wider pore mouths located in the vicinity of the crystallite surface which explains the higher value of external surface
area. Mean pore width value of SZ2 sample is close to the relevant
SAPO-34 material value whereas ZS2 sample includes mesopores
with mean pore width close to ZSM-5 one.
Demonstrated nitrogen adsorptiondesorption isotherms in
Fig. 12 conrm mentioned results. Isotherms are a composite of
types I and IV, corresponding to microporous and mesoporous
materials according to IUPAC classication [46]. The steep uptake
in the low-pressure region (P/P0 < 0.1) is the typical property of
microporous compounds. Hysteresis appearing in the physisorption isotherms at the relative pressure P/P0 ranging from ca. 0.45
to 1.0 is usually associated with capillary condensation taking
place in mesopore structures. Moreover, hysteresis loop shape in
all isotherms is close to H4 type which is indicative of the presence
of narrow slit-like pores implanted in a matrix of much smaller
tiny pores. Smaller hysteresis loop in ZSM-5 isotherm veries that
ZSM-5 has the lowest amount of mesopores which might be
caused by the formation of extra-framework aluminum oxide species and interspaces between plate-like single crystals. This result
is in well agreement with those of FESEM and EDS analyses.
The presence of mesopores would effectively interconnect the
micropores to offer additional diffusional paths which are of great
benet for catalysis as a whole. Therefore it is worth investigating
pore size distribution by means of BJH calculation method. Fig. 13
shows that all samples had a mesopore size distribution

Fig. 8. FESEM images of (a and b) ZS2 composite.

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

183

Fig. 9. Elemental mapping (O, Na, Al, Si, and P) of the developed SZ2 composite.

centralizing around 5 nm. These curves indicate that these mesopores in composites were inherited from the parent materials i.e.
SAPO-34 and ZSM-5. However there are several humps within
2060 nm in SZ2 and ZS2 composites patterns suggesting the
formation of intercrystalline pinholes through close packing of
surface layer material crystallites. Also, some of mesopores in
SZ2 binary structure is believed to be induced from recrystallization of SAPO34 during the preparation of composite. It can be
deduced that more mesopores are formed combining two zeolites,
due to the increscent interspaces among the different crystals.
Considering the facts mentioned above, these cavities might be
served as transport paths for bulky compounds such as intermediate carbocations to alleviate the limitations in catalytic reactions.
NH3-TPD is an efcient technique for the evaluation of acidic
properties of the surface including quantity and strength of acid
sites. Generally, the specic peak area of a TPD prole can be
determined by Gaussian curve-tting that signies the amount of
desorbed ammonia which is proportional to the total acidity
whereas the peak position corresponds to the strength of acid sites.
NH3-TPD proles over ZSM-5, physical mixture and SZ2 molecular
sieves are illustrated in Fig. 14. ZS2 sample was not characterized
since its catalytic performance was similar to SZ2 sample implying
the dominance of same acidic regime in both samples which will
be shown and discussed in next section. Quantitative information
on the position and strength of major desorption peaks of each
sample are provided in Table 5. Peaks in the temperature range
of 50500 C, express the concentration of weak (50150 C),

medium (150350 C) and strong (>350 C) acid sites. Moreover,


according to Zhang et al. [47] only Bronsted-bound ammonia is driven off at temperatures higher than 440 C which is directly related
to the density of Bronsted acid sites while desorption at lower temperatures may arise from either of Bronsted and Lewis acid sites.
As shown in Fig. 14, three desorption steps at 80, 290 and
480 C were observed for ZSM-5 leading to this assumption that
three types of acid sites exist. The rst peak is due to the existence
of surface hydroxyl groups as weak acid sites while third peak
arose possibly from structural acidity. It is generally accepted that
the weak acid sites in ZSM-5 zeolite result from the non-framework aluminum or SiOH structural defects. FTIR spectrum of
ZSM-5 revealed that there were extra-framework AlOH groups
with low acid strength which is consistent with the result herein.
However, the symmetric localized HT (high-temperature) desorption peak around 480 C is indicative of the dominance of strong
Bronsted acid sites. Two distinctive desorption peaks at around
97 and 480 C were detected for the physical mixture of SAPO-34
and ZSM-5, which are usually ascribed to ammonia desorption
from weak and strong acid sites. As shown in Table 5, composite
molecular sieve possessed the highest acidity, in particular the
weak acid sites, in comparison with physical mixture and ZSM-5
with the lowest acidity. There is a well-resolved peak at 106 C
and a minor peak at 360 C, clearly suggesting that most of the
acid sites were weak. A distinct increase of ammonia desorption
at low-temperature and almost no ammonia desorption at high
temperature after the incorporation of SAPO-34 to the structure,

184

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

Fig. 10. Elemental mapping (O, Na, Al, Si, and P) of the developed ZS2 composite.

without taking part in cracking side reaction and enhance the control over the selectivity of desired products. In case of physical
mixture and composite zeolite, the LT (low-temperature) desorption peaks shifted towards higher temperatures slightly meaning
that the strength of weak acid sites were increased which was
probably caused by the contribution of POH and Si and/or Al
bridges hydroxyl groups. These ndings demonstrate that the synergy between two structures with different features can change the
acid character of the material effectively.
3.2. Catalytic performance of composite catalysts

Fig. 11. FT-IR patterns of composite samples as well as pristine SAPO-34 and ZSM-5
as references: (a) SAPO-34, (b) SZ2, (c) ZS2 and (d) ZSM-5.

point out that the strong acid sites on the ZSM-5 zeolite were preferentially transformed and/or neutralized during the synthesis
procedure. It is a huge advantage to propane dehydrogenation
reaction which is assumed to proceed through Olahs monomolecular mechanism [48,49]. Based on this mechanism, zeolite protonates alkane at high temperature and creates a pentacoordinated
carbonium ion which can further rearrange yielding molecular H2
and a C+3 carbenium ion. Strong acid sites can promote the protolytic cracking of the original hydrocarbon chains of the initially
formed carbenium ion through recombination of carbenium ion
with an alkyl moiety lowering propylene selectivity. It is worth
noting that the overall weak acidity of the zeolitic support is a
substantial requirement to convert intermediates to propylene

Catalytic performance analyses of all prepared samples were


carried out in a micro-reactor for direct dehydrogenation of propane and results are compared with those of individual ZSM-5
and physical mixture-derived catalysts (Figs. 15 and 16). All binary
structures performed better than ZSM-5 single structure catalyst.
Except ZS1, rest of composite catalysts showed higher conversion
and propylene selectivity than physical mixture. We expected this
outcome in accordance with the given explanation in Section 3.1
analyzing FESEM observations. PtSn/SZ1 sample suggests the huge
effect of SAPO-34 contribution in determination of nal catalytic
properties of obtained support which could not be obtained from
thorough physical blending of SAPO-34 and ZSM-5. It can be seen
from Figs. 15 and 16 that increasing the quantity of incorporated
SAPO-34 efcaciously led to an increase in converted propane
towards desired propylene product with higher reactivity. And
besides, the stability of PtSn/SZ2 catalyst was increased comparing
with PtSn/SZ1 composite catalyst. Both PtSn/SZ2 and PtSn/ZS2
showed high performance with similar conversion and selectivity.



3654 (w)
3436 (s)
3437 (s)
3500 (w)
1634 (ms)
1631 (ms)
1635 (w)
1099 (s)
1099 (s)


1224 (w)
1224
(ms)


1096 (s)

s, ms and w express the intensity of IR band. s: strong, ms: mild-strong, w: weak.


a
T = Si and Al.
b
T = Si, Al, and P.
c
T = Si and Al.

803 (ms)
797 (ms)
794 (ms)
546 (ms)
546 (s)
545 (s)



627 (w)








3434 (s)
1642 (ms)
1099 (s)




875 (w)
720 (w)
638 (w)

490
(ms)
459 (s)
451 (s)
451 (s)
SAPO34
SZ2
ZS2
ZSM-5

Hydroxyl SiOHAl
groups
Adsorbed
water
TO bond
SiOT
linkagesc
TOT
groupsb

Asymmetric stretch

D5R
rings
SiO

Symmetric stretch

PO (AlO)
groups
TO4a

D6R
rings
TO bend

Type
Sample

Table 3
Assignment of the observed IR bands of the selected SZ2 and ZS2 composite and individual ZSM-5 and SAPO-34 refrence samples.

OPO
groups

Stretching vibration

Hydroxyl AlOH
groups

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

185

The differences between catalytic performances of the SZ2 and


ZS2 composites in comparison with physical mixture suggest that
the coreshell structure of composite materials effectively contributes to the enhancement of catalytic performance. There is no
intimate contact between two phases in physical mixture which
thereby downgrades the synergism and impacts on the subsequent
performance of material. This fact proves that catalytic promotion
cannot be achieved unless a strong interaction is developed
through a chemical reaction that strictly inuences the governing
acidity pattern, topology and pore structure. These parameters
are of great signicance for acid and shape-selective catalytic
properties.
Meanwhile, the presence of smaller and shorter channels and
channel intersections in CHA topology bring the steric constraints
which hinder the formation of coke deposits and heavy compounds
as coke precursors and enhance light olens selectivity. Koyama
et al. [50] proposed a plausible mechanism for propylene production based on the volume of zeolite pores or in other words the
diameter of the largest sphere that can be included in a zeolite
framework (Di). They concluded that the selective production of
propylene proceed over zeolite, whose pore volume accommodates
the volumes of the olens and/or their carbenium cations. The
maximum propylene selectivity was observed at a Di value
between 7.0 and 7.6 which corresponds to an estimated pore volume of 180230 3. As a result, 8-membered rings SAPO-34 with
crystallographic pore diameter of 3.8 and Di = 7.31 was recognized as a superior sieve for selective production of propylene. In
view of above discussed issues, providing control over the stereochemistry of PDH reaction by using a robust selective catalyst is
a crucial factor, as if the cracking products i.e. methane and ethane
become considerable, it would be very difcult to convert them
into olens.
Also, mild acidity of SZ2 composite material enhanced the synergetic effect between two structures by the neutralization and/or
transformation of some of strong acid sites due to the hydrothermal effect of subsequent synthesis procedure and thus improved
the catalytic activity of the composite derived catalyst. It is fully
comprehended that the dehydrogenation of light parafns is
directly associated with acidic properties of the catalyst [32,33].
Besides, it is known that a predominant decline in the quantity
of strong acid centers occur after metallic incorporation of Pt and
Sn species [29,32]. Pt atoms locate on Bronsted acid sites and consume some of them while Sn atoms migrate into the pore channels
and neutralize some of the acid sites within zeolite channels [51].
This phenomenon can be the possible explanation for the lower
catalytic reactivity of PtSn-based ZSM-5 catalyst bringing in mind
the low acidity of ZSM-5 mainly consisting of strong acid sites.
On the contrary, SZ2 sample possessed a high density of weak acid
sites which are supposed to be slightly inuenced by impregnation
step. Therefore, the match between acid and metal functions
resulted in the favorable formation of the light olens especially
propylene inhibiting cracking of C+3 carbenium ions intermediates.
Catalytic reactivity of either of PtSn supported SZ2 and ZS2
catalysts were almost the same which signies the dominance of
similar acidic pattern in ZS2 sample. But it is not sensible to isolate
a specic factor to analyze its effect on the reaction. The reactions
of hydrocarbons are vastly inuenced by the residence time of
reactants molecules inside the pores. Given that, the presence of
mesopores and a very thin SAPO-34 shell layer assisted the diffusion of reactants and subsequent intermediate products. Rapid
adsorption and desorption of reactant and product molecules from
the PtSn/ZS2 hybrid catalyst might be another reason for high conversion since this would release active sites quickly for further
adsorption and conversion of reacting species. On the other hand,
the high yield of propylene benet from thin SAPO-34 porous shell
with the highest external surface area as its short diffusion lengths

186

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

Table 4
Physicochemical properties of SAPO34, composites, and ZSM-5 zeolites.

a
b
c
d

Sample

SBET (m2/g)a

Smic (m2/g)b

Vt (cm3/g)

SAPO-34
SZ2
ZS2
ZSM-5

427.4
308.7
279.8
275.7

390
256.8
185.7
214.2

0.29
0.22
0.23
0.18

Vmic (cm3/g)b

Mean pore width (nm)d

0.18
0.12
0.09
0.1

9.3
10.3
7.6
8.4

SBET was obtained by analyzing nitrogen adsorption data at 77 K in a relative vapor pressure ranging from 0.05 to 0.3.
Micropore area and micropore volume were determined using t-plot method.
Total pore volume was estimated based on the volume adsorbed at P/P0  0.99.
Mean pore width was determined by using BJH method.

Fig. 12. The N2 adsorptiondesorption isotherms of ZSM5, SZ2, ZS2 and SAPO-34
samples. (The Y-axis was shifted a little for ZS2 sample to show the change in the
isotherms.)

impeded secondary reactions of terminal products while diffusing


outward the pores. Such interpretation is suitably in line with pore
continuum theory wherein a minimum energy barrier will be
confronted in the inward and outward diffusion [52].
Figs. 17 and 18 represent the average product distributions and
yields over PtSn-based catalysts with various carriers. Each of SZ2
and ZS2 samples displayed >83% of average propylene selectivity
and average propylene yield up to ca. 31.5%. It is interesting to note
that channel intersections of ca. 9 in intermediate pore ZSM-5

zeolite have an important role in shape-selective catalytic reactions. It has been reported that up to 75% of total acid sites could
be present in channel intersections and the rest along the channels
(d  5.5 ) [53]. Besides hydrocarbons are believed to diffuse
through channels and channel intersections on the basis of their
molecular sizes and activation energies (attraction or repulsion)
[54]. It can be concluded that SAPO-34 with channel intersecting
windows of 7 sounds to be a superior molecular sieve for propylene and ethylene products. To sum up, acidity amount and
strength, topological aspects and specic pore structure of catalyst
(Sext and 3-D channels) can be named as the key parameters playing a critical role in the production of light olens in a controlled
trend.
Lower propylene selectivity dropping rates and ethylene selectivity average values in Figs. 16 and 17 for PtSn/SZ2 and PtSn/ZS2
catalysts suggest that cracking (carboncarbon bond) dissociation
phenomenon has been slightly encumbered underlining positive
inuence of SAPO-34 contribution. All mentioned facts imply better distribution of acidic and consequently active centers which
collaborated with high sieving property of SAPO-34 so as to
impede major side reactions including cracking and hydrogenolysis to control the composition of formed light olens mixture and
enhance propylene yield. On the other hand, the synergetic
catalysis of ZSM-5 and SAPO-34 structures has occurred well in
SZ2 and ZS2 selected samples. In summary, preparation method
and composition have to be pointed out as factors with paramount

Fig. 13. The BJH pore size distribution of ZSM-5, SZ2, ZS2 and SAPO-34 samples.

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

187

Fig. 17. Average product distributions over different PtSn-based composite


catalysts.
Fig. 14. NH3-TPD plots of ZSM-5, physical mixture and SZ2 samples.

Table 5
NH3-TPD data and acid amount of ZSM-5, physical mixture and SZ2 composite
samples.
Sample

ZSM-5
Physical mixture
Composite

Tpeak (C)

Acid amount (mmol/g)

LT peak

HT peak


96.6
105.4

482.7
482.1


Weak acidity
1.53
1.85

Strong acidity
0.17
0.38

Fig. 18. Average yields of products over different PtSn-based composite catalysts.

Fig. 15. Propane conversion as a function of time on stream over different PtSnbased composite and ZSM-5 catalysts.

Fig. 16. Propylene selectivity as a function of time on stream over different PtSnbased composite and ZSM-5 catalysts.

importance in obtaining catalysts with desired reactivity and


shape-sieving property.
An optimum composite catalyst was prepared according to the
represented synthesis procedure for SZ2 sample and by using

ZSM-5 with Si/Al ratio of 60. This ratio is selected based on our
ndings for ZSM-5 zeolites with three Si/Al levels which are going
to be published elsewhere. Additionally, TPD and FTIR spectroscopy proles of different Si/Al ratios ZSM-5 zeolites have been
characterized by Zhu et al. [55]. Results showed that support acid
strength was associated with Si/Al ratio inversely. Nawaz et al.
[30] found out remarkable effect of Si/Al ratio of PtSn supported
on ZSM-5 zeolites on the catalytic performance of catalysts owing
to the suppression of secondary reactions. Propylene selectivity
increased as the Si/Al ratio and consequently acidity decreased.
Effect of Si/Al ratio parameter on surface acidity of HZSM-5/
SAPO-34 catalysts and its reactivity in ethanol to propylene reaction was investigated as well [25]. SZ2 novel catalyst was chosen
based on the simplicity of the short-time synthesis procedure
when using pre-coated SAPO34 powder and economical issues
regarding the usage of low amount of TEAOH (20%) template. As
it has been mentioned before, ZS2 sample with similar catalytic
performance to SZ2 support was prepared employing high amount
of expensive TEAOH (35%) template. In addition, in case of SAPO-34
particles being covered by ZSM-5 particles, they are believed to be
strengthened for steam exposure wherein the credit goes to the
hydrothermally stable carrier for PDH reaction in the presence of
steam as additive and these types of reactions. Fig. 19 pictures time
course of propane conversion on PtSn/SZ2 (Si/Al = 60) through PDH
reaction and is an overall reaction network to evaluate the efciency of synthesized composite catalyst. As it can be observed,
intensication of support by adapting ZSM-5 with Si/Al = 60 in
the synthesis procedure resulted in an improvement in the performance of catalyst and more control on stereo-chemistry of
reaction. Reliable stable selectivity towards light olens,

188

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189

Fig. 19. Catalytic performance of PtSn/SZ2 (Si/Al = 60) composite catalyst versus time on stream (h).

Table 6
Comparison of reaction performance of PtSn/SZ2 (Si/Al = 60) and commercial PtSnK/cAl2O3 catalysts (Reaction condition T = 600 C, Catalyst weight = 0.5 g, WHSV = 4 h1, and
H2/C3 = 0.8).
TOSa (h)

0.5
6
a
b
c

PtSnK/cAl2O3

PtSn/SZ2 (Si/Al = 60)


Conversion (%)

Spropyleneb (%)

Solens (%)

Dc (%)

Conversion (%)

Spropylene (%)

Solens (%)

D (%)

47.7
40.9

92.2
85.9

94.6
92.2

14.3

46.5
43.5

87.4
85.4

91.2
90.6

6.4

Time-on-stream.
Selectivity.
Deactivation parameter which was dened as D = [(X0  Xf)/X0]  100, where X0 is the initial propane conversion and Xf is the nal propane conversion.

prominent selectivity towards propylene and high reactivity of


designated composite catalyst conrm the success of catalyst formulation procedure. The performance of the optimum catalyst in
terms of conversion and selectivity is compared with commercial
tri-metallic PtSnK/cAl2O3 and the results are provided in Table 6.
Higher initial activity and propylene selectivity of our catalyst
proves the efciency of synthesized catalyst although a gradual
decline of the conversion was involved due to the deactivation of
the active sites. It is noteworthy to mention that the high stability
of commercial catalyst stemmed from the potassium species
impregnated on the surface which has a great impact on the acidity
and interactions between support and metals. It is well known that
the addition of suitable amount of potassium can hinder deactivation phenomenon in which active sites and pores will be covered
with coke deposits [56].
4. Conclusion
In summary, two series of ZSM-5/SAPO-34 and SAPO-34/ZSM-5
hybrid composites with binary hierarchical structures were
successfully synthesized with hydrothermal methods. A facile
new technique has been proposed for the synthesis of zeolitic
SAPO-34/ZSM-5 bi-phase composite utilized with TPA-exchanged
SAPO-34 crystals. It seems that pre-treatment of SAPO-34 crystals
in organic amine solution containing TPA+ species was a determining step to preserve SAPO-34 crystals from phase transformation

and dissolution when they were exposed to harsh basic condition


of ZSM-5 pre-crystallized gel mixture. In this case, ZSM-5 layer
was formed without any problem rising from cores dissolution
and considerable detachment phenomenon. Moreover, role of
composites as carrier were evaluated in direct dehydrogenation of
propane reaction and compared with those of physical mixturederived composite and individual ZSM-5. Results showed that
binary structure of the composites did not only combine reactivity
and dominant sieving property of ZSM-5 and SAPO-34, but also had
inuential impacts on the product features to adjust the reaction
towards higher activity and propylene selectivity. Physical mixture
consisted of two separate phases while they were intimately
bonded in the composites by the interfacial effects. Thus their
synergism in acidity, topology, and pore structure were enhanced
which could not be achieved simply by mixing the same single zeolites. The yield of propylene and the catalytic stability of the PtSnbased composite catalysts were closely relative to the composition
and applied synthesis strategy. PtSn-SAPO-34/ZSM-5 (Si/Al = 60)
novel catalyst represented a high potential as a promising PDH
catalyst to meet a change of market demand for propylene.

References
[1] R. Singh, P.K. Dutta, in: S.M. Auerbach, K.A. Carrado, P.K. Dutta (Eds.),
Handbook of Zeolite Science and Technology, Marcel Dekker Inc., New York,
Basel, 2003, pp. 2164.

M. Razavian, S. Fatemi / Microporous and Mesoporous Materials 201 (2015) 176189


[2] A. Corma, Chem. Rev. 97 (1997) 23732419.
[3] M.W. Ackley, S.U. Rege, H. Saxena, Microporous Mesoporous Mater. 61 (2003)
2542.
[4] H. Nur, S. Ikeda, B. Ohtani, J. Braz. Chem. Soc. 15 (2004) 719724.
[5] R. Chal, C. Gerardin, M. Bulut, S. van Donk, Chem. Catal. Chem. 3 (2011) 6781.
[6] D.P. Serrano, J. Aguado, J.M. Escola, Catalysis 23 (2011) 253283.
[7] K. Kubo, H. Iida, S. Namba, A. Igarashi, Microporous Mesoporous Mater. 188
(2014) 2329.
[8] J. Cejka, S. Mintova, Catal. Rev. Sci. Eng. 49 (2007) 457509.
[9] Z. Xie, Z. Liu, Y. Wang, Q. Yang, L. Xu, W. Ding, Int. J. Mol. Sci. 11 (2010) 2152
2187.
[10] Y. Bouizi, L. Rouleau, V.P. Valtchev, Chem. Mater. 18 (2006) 49594966.
[11] Y. Bouizi, I. Diaz, L. Rouleau, V.P. Valtchev, Adv. Funct. Mater. 15 (2005) 1955
1960.
[12] J. Zheng, X. Zhang, Y. Wang, Y. Bai, W. Sun, R. Li, J. Porous Mater. 16 (2009)
731736.
[13] D. Kong, J. Zheng, X. Yuan, Y. Wang, D. Fang, Microporous Mesoporous Mater.
119 (2009) 9196.
[14] Yu Fan, Duo Lei, Gang Shi, Xiaojun Bao, Catal. Today 114 (2006) 388396.
[15] A. Lombard, A. Simon-Masseron, L. Rouleau, A. Cabiac, J. Patarin, Microporous
Mesoporous Mater. 129 (2010) 220227.
[16] D.V. Vu, M. Miyamoto, N. Nishiyama, S. Ichikawa, Y. Egashira, K. Ueyama,
Microporous Mesoporous Mater. 115 (2008) 106112.
[17] Y. Bouizi, L. Rouleau, V.P. Valtchev, Microporous Mesoporous Mater. 91 (2006)
7077.
[18] S. Yang, N.P. Evmiridis, Microporous Mater. 6 (1996) (1996) 1926.
[19] M.S. Francesconia, Z.El. Lpezb, D. Uzcteguib, G. Gonzlezb, J.C. Hernndeza,
A. Uzcteguia, A. Loaizaa, F.E. Imberta, Catal. Today 107108 (2005) 809815.
[20] M.E. Leonowicz, D.E.W. Vaughan, Nature 329 (1987) 819820.
[21] X. Qi, D. Kong, X. Yuan, Z. Xu, Y. Wang, J. Zheng, Z. Xie, J. Mater. Sci. 43 (2008)
56265633.
[22] T. Wakihara, S. Yamakita, K. Iezumi, T. Okubo, J. Am. Chem. Soc. 125 (2003)
1238812389.
[23] Q. Zhang, H. Shan, C. Li, S. Xu, C. Yang, J. Porous Mater. 20 (2013) 171176.
[24] H.J. Chae, Y.H. Song, K.E. Jeong, C.U. Kim, S.Y. Jeong, J. Phys. Chem. Solids 71
(2010) 600603.
[25] C. Duan, X. Zhang, R. Zhou, Y. Hua, L. Zhang, J. Chen, Fuel Process. Technol. 108
(2013) 3140.
[26] A. Corma, F.V. Melo, L. Sauvanaud, F. Ortegaa, Catal. Today 107108 (2005)
699706.
[27] M.M. Bhasin, J.H. McCain, B.V. Vora, T. Imai, P.R. Pujad, Appl. Catal. A 221
(2001) 397419.
[28] C. Yu, Q. Ge, H. Xu, W. Li, Catal. Lett. 112 (2006) 197201.

189

[29] Y. Zhang, Y. Zhou, A. Qiu, Y. Wang, Y. Xu, P. Wu, Catal. Commun. 7 (2006) 860
866.
[30] Z. Nawaz, T. Xiaoping, F. Wei, Korean J. Chem. Eng. 26 (2009) 15281532.
[31] L.G.A. van de Water, J.C. van der Waal, J.C. Jansen, T. Maschmeyer, J. Catal. 223
(2004) 170178.
[32] Z. Nawaz, X. Tang, Q. Zhang, D. Wang, W. Fei, Catal. Commun. 10 (2009) 1925
1930.
[33] Z. Nawaz, F. Wei, J. Ind. Eng. Chem. 16 (2010) 774784.
[34] B. Howorth, B.S. Thesis, Oregon State University, 2013.
[35] K. Ozalas, B.F. Hajek, Clays Clay Miner. 44 (1996) 811817.
[36] Y. Lv, X. Qian, B. Tu, D. Zhao, Catal. Today 204 (2013) 27.
[37] B.K. Tanner, T.P.A. Hase, T.A. Lafford, M.S. Goorsky, Adv. X-Ray Anal. 47 (2004)
309314.
[38] M.D. Foster, O.D. Friedrichs, R.G. Bell, F.A. Almeida Paz, J. Klinowski, J. Am.
Chem. Soc. 126 (2004) 97699775.
[39] Crystallographic information les made by reference to http://www.izastructure.org/databases/.
[40] S. Abello, A. Bonilla, J. Perez-Ramirez, Appl. Catal. A 364 (2009) 191198.
[41] H. Yingping, L. Min, D. Chengyi, X. Shutao, W. Yingxu, L. Zhongmin, G. Xinwen,
Chin. J. Catal. 34 (2013) 11481158.
[42] P.A. Jacobs, H.K. Beyer, J. Valyon, Zeolites 1 (1981) 161168.
[43] S. Sistani, M.R. Ehsani, Iran. J. Chem. Chem. Eng. 29 (2010) 99104.
[44] J. Tan, Z. Liu, X. Bao, X. Liu, X. Han, C. He, R. Zhai, Microporous Mesoporous
Mater. 53 (2002) 97108.
[45] M. Salmasi, S. Fatemi, A.T. Najafabadi, J. Ind. Eng. Chem. 17 (2011) 755761.
[46] K.S.W. Sing, I.U.P.A.C. 54 (1982) 22012218.
[47] W. Zhang, E.C. Burckle, P.G. Smirniotis, Microporous Mesoporous Mater. 33
(1999) 173185.
[48] G.A. Olah, R.H. Schlosberg, J. Am. Chem. Soc. 90 (1968) 27262727.
[49] G.A. Olah, G. Klopman, R.H. Schlosberg, J. Am. Chem. Soc. 91 (1969) 3261
3268.
[50] T.R. Koyama, Y. Hayashi, H. Horie, S. Kawauchi, A. Matsumoto, Y. Iwase, Y.
Sakamoto, A. Miyaji, K. Motokura, T. Baba, Phys. Chem. Chem. Phys. 12 (2010)
25412554.
[51] L. Bai, Y. Zhou, Y. Zhang, H. Liu, M. Tang, Catal. Lett. 129 (2009) 449456.
[52] R. Le Van Mao, N. Al-Yassir, D.T.T. Nguyen, Microporous Mesoporous Mater. 85
(2005) 176182.
[53] P. Hemalatha, M. Ganesh, M. Palanichamy, V. Murugesan, P. Yong-Ki, C.W.
Choon, J.H. Tae, Chin. J. Catal. 34 (2013) 294304.
[54] J. Xiao, J. Wei, Chem. Eng. Sci. 47 (1992) 11431159.
[55] X. Zhu, S. Liu, Y. Song, L. Xu, Appl. Catal. A 288 (2005) 134142.
[56] S. Zhang, Y. Zhou, Y. Zhang, L. Huang, Catal. Lett. 135 (2010) 7682.

You might also like