You are on page 1of 8

Microporous and Mesoporous Materials 151 (2012) 26–33

Contents lists available at SciVerse ScienceDirect

Microporous and Mesoporous Materials


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

Uniform mesoporous ZSM-5 single crystals catalyst with high resistance


to coke formation for methanol deoxygenation
Ali A. Rownaghi ⇑, Fateme Rezaei, Jonas Hedlund
Division of Chemical Technology, Luleå University of Technology, SE-971 87 Luleå, Sweden

a r t i c l e i n f o a b s t r a c t

Article history: In this study, an efficient procedure for controllable synthesis of uniform mesoporous ZSM-5 single crys-
Received 14 August 2011 tals with various crystal architectures and high mesoporosity was developed. Compared with conven-
Received in revised form 14 October 2011 tional ZSM-5 catalyst, mesoporous ZSM-5 single crystals synthesized by this method exhibited
Accepted 13 November 2011
significantly higher external surface area and larger mesopore volume than conventional one. The cata-
Available online 20 November 2011
lytic performance of mesoporous zeolites were evaluated in the conversion of methanol to hydrocarbons
using a fixed-bed reactor operating at 370 °C, atmospheric pressure and weight hourly space velocities
Keywords:
(WHSV) of 0.16–6.58 h 1. By controlling the ZSM-5 synthesis procedure, the activity and stability of
Mesoporous ZSM-5 single crystals
Methanol dehydration
the ZSM-5 catalyst in the conversion of methanol to gasoline-range hydrocarbons can be favorably tuned.
Deactivation We also found that mesoporosity plays a crucial role in catalyst stability. Good correlation was observed
Catalyst lifetime between catalyst lifetime and mesoporosity. While the catalyst activity is related to the acid site density,
Methanol-to-gasoline the catalyst stability (deactivation rate) correlates with the measured surface ratio of BET and external
surface area (SBET/Smeso). The novel ZSM-5 catalyst exhibited improved stability due to the faster removal
of products with shorter diffusion path-length and lower coke formation. The obtained results indicated
that the novel mesoporous ZSM-5 catalyst containing relatively large pores (mostly mesopores) enhances
reaction yield towards gasoline-range hydrocarbons. It is therefore concluded that the relatively slow
deactivation rate (coke formation) of novel mesoporous ZSM-5 single crystals catalyst makes this zeolite
the preferred catalyst for the conversion of methanol to gasoline-range hydrocarbons at mild conditions.
Ó 2011 Elsevier Inc. All rights reserved.

1. Introduction other chemicals [4–9]. In addition to Mobil’s methanol to gasoline


(MTG) technology [1], several methanol to hydrocarbon (MTH)
During the first global energy crisis in the early 1970s, alterna- processes are now being commercialized, including the Lurgi’s
tives to fuels derived from crude oil became necessary. Methanol methanol to propylene (MTP) [10], hydro/UOP’s methanol to ole-
to gasoline (MTG) process was developed by the Mobil company fins (MTO) [11], and Topsøe integrated gasoline synthesis (TIGAS)
in 1986 [1]. This process was not successfully commercialized [12]. These processes are all examples of methanol to hydrocar-
due to significant drop in crude oil prices in the 1990s. Recently, bons using zeolite materials as catalysts. Hence, a novel route to
with the problems associated with fossil resources use – namely, bring higher value on an energy content basis from renewable re-
greenhouse gas emission and looming petroleum crisis – biomass sources can be achieved by MTH processes.
to hydrocarbon process has received renewed interest as it will be- ZSM-5 zeolite with different SiO2/Al2O3 ratios has proven to be
come economically feasible. Biomass thermochemical-derived one of the promising catalysts for conversion of small and large
intermediates contain a large fraction of light mixed oxygenated oxygenates molecules to hydrocarbon products, which are similar
compounds (less than C6) that are chemically and thermally to gasoline-range molecules (C4–C10) of conventional petroleum
unstable and need to be deoxygenated to form compatible liquid [13–15]. The activity of a ZSM-5 catalyst for the methanol dehydra-
hydrocarbon molecules which could be a direct feed into current tion is determined by the amount, the nature, and the accessibility
petroleum refineries or that are direct replacements of diesel, of its acid sites. In addition, it is well known that the structural de-
gasoline, or chemical products [2,3]. Furthermore, conversion of fects in ZSM-5 zeolite can strongly influence the catalytic activity
methanol to hydrocarbons over solid acid catalysts (such as zeo- [16–18].
lites) has drawn increasing attention in the petrochemical industry One of the main drawbacks of the methanol to hydrocarbon
as one of the efficient methods for manufacturing olefins, fuels and process is related to the deactivation of the catalyst, due to coke
formation as a consequence of consecutive reactions between light
⇑ Corresponding author. Tel.: +46 920 49 12 97; fax: +46 920 49 11 99. olefins and Brønsted or Lewis acid sites. This problem is due to the
E-mail addresses: ali.rownaghi@ltu.se, ali.rownaghi@gmail.com (A.A. Rownaghi). limited diffusion of reactants/products within the zeolite materials

1387-1811/$ - see front matter Ó 2011 Elsevier Inc. All rights reserved.
doi:10.1016/j.micromeso.2011.11.020
A.A. Rownaghi et al. / Microporous and Mesoporous Materials 151 (2012) 26–33 27

[4,6]. To address this issue, mesoporous zeolites with faster diffu- the clear aluminate solution was added to tetraethylorthosilicate
sion rate can be synthesized. In fact, higher resistance to coke poi- in the polypropylene bottle. In order to achieve a total hydrolysis
soning and shorter diffusion path are the important advantages of of the silicon species, the final solution was placed on shaker for
mesoporous zeolites for their successful applications in catalytic 24 h at room temperature to hydrolyze TEOS and obtain a clear
reaction processes [19,20]. solution. Hydrothermal treatment in a Teflon-lined autoclave for
Several synthesis strategies for producing zeolite-based materi- 24–30 h at 140–170 °C yielded white crystals that were further
als with improved accessibility to the catalytically active sites lo- purified by repeated centrifugation at 15,000 rpm for 30 min fol-
cated in the microporous crystals have been pursued; they lowed by re-dispersion in water for four times. The purified prod-
include steaming, acid leaching treatments, carbon templating, uct was then freeze-dried. The calcined sample was obtained by
desilication, etc. [21–32]. Christensen et al. [21,22] reported the treatment of as-synthesized material in static air at 500 °C for
preparation of mesoporous ZSM-5 crystals by carbon-templating 16 h (the calcination heating and cooling rate were 5 °C/min).
and catalytic evaluation in several reactions including gas phase The SiO2/Al2O3 ratio was 50 on the basis of elemental analysis.
alkylation of benzene with ethylene and MTG. Another interesting After air drying, mesoporous ZSM-5 crystals as small as 5 nm in
study published by the Kustova et al. [23,24] was the use of mes- the form of mesoporous aggregates of approximately 1.95 lm in
oporous iron-containing zeolite as catalysts for selective catalytic length and an average pore diameter of 8 nm were obtained. This
reduction (SCR) of nitrogen oxide (NOx) with NH3. It was shown structure is very open giving rise to very large surface areas, mes-
that mesoporous Fe-ZSM-5 was more active than the conventional opore volume with a very narrow particle size distribution.
ZSM-5 at all temperatures studied. Recently, base-treatment
(desilication) of zeolite either alone or in combination with dealu- 2.1.2. Synthesis of conventional ZSM-5 crystals
mination, has been performed to obtain similar advantages. Perez- A conventional ZSM-5 (Con-ZSM-5) zeolite was synthesized
Ramırez et al. [25,26] synthesized a mesoporous ZSM-5 using the according to the procedure described elsewhere [35] and are omit-
post-synthetic steaming and alkali treatment and evaluated it in ted here for brevity.
the liquid phase benzene alkylation evidenced a higher activity
and selectivity to ethyl benzene. In a recent study published by 2.2. Catalyst characterization
Kim et al. [28], a synthesis method capable of synthesizing sheets
of crystalline MFI was reported using an elegantly designed bifunc- 2.2.1. X-ray diffraction (XRD)
tional surfactant as the structure-directing agent. Such MFI config- X-ray powder diffraction (XRD) was performed to verify the
uration significantly increased the catalytic lifetime for MTG crystallinity of the zeolites. The diffraction patterns were recorded
reaction by more than five times compared to a conventional zeo- on a Siemens diffractometer model D 5000 running in Bragg–
lite with similar Si/Al ratio. Even though some of these approaches Brentano geometry employing CuKa (k = 1.54439 Å) radiation to
are effective and result in improved catalytic performance in vari- generate diffraction patterns from powder crystalline samples at
ous reactions, troublesome pre-preparation is still unavoidable in ambient temperature. The spectra were scanned at a rate of 2.0°/
most cases. Furthermore, treating zeolite crystals by leaching agent min in the range 2h = 2–70°.
(desilylation and dealumination) can cause a change in the SiO2/
Al2O3 ratio and nature of framework of zeolite [33,34]. Therefore, 2.2.2. Surface area and porosity
a developed route is highly desired in terms of simplicity, repro- Nitrogen adsorption/desorption isotherms at 196 °C were re-
ducibility and large-scale applicability. corded using a Micromeritics ASAP 2010 instrument. Before the
Herein, we propose a facile and low cost route for synthesis of measurements, samples were heated to 350 °C in a vacuum for at
high Al-content mesoporous ZSM-5 single crystals. To investigate least 12 h. The Brunauer–Emmett–Teller (BET) equation was used
how the activity and, in particular, the deactivation relate to the to calculate the specific surface area (SBET) using adsorption data
mesoporosity (structural properties) of ZSM-5 catalyst, we present at p/p0 = 0.05–0.25. The mesopore size distribution was derived
measured catalyst lifetime (activities and deactivation rates) and from the adsorption branch by using the Barrett–Joyner–Halenda
hydrocarbon distribution in the conversion of methanol to hydro- (BJH) method. The micropore volume was determined from a
carbons of over a wide range of space time on novel mesoporous t-plot.
ZSM-5 single crystals and conventional ZSM-5.
2.2.3. Scanning electron microscopy (SEM)
The size, shape and size distribution of zeolite crystals were re-
2. Experimental section
corded using a FEI Magellan™ 400L instrument. Zeolite crystals
were deposited on a silicon wafer from dispersion and the sample
2.1. Catalyst preparation
was then calcined in air at 500 °C. SEM images were recorded with-
out any coating. The average crystal length was determined by
Both ZSM-5 zeolites were synthesized from clear colloidal syn-
measuring the length of 50 crystals for each sample.
thesis mixtures using different synthesis procedures. Tetraethylor-
thosilicate (TEOS, 98%, Aldrich) and aluminum isopropoxide (AIP,
2.2.4. Temperature-programmed desorption (TPD)
Aldrich) were used as Si and Al sources, respectively. Concentrated
The acid site density of synthesized zeolites was determined by
tetrapropylammonium hydroxide (TPAOH, 40 wt.% aqueous solu-
temperature-programmed desorption of ammonia (NH3-TPD)
tion, AppliChem) was also used as a template. The detailed synthe-
using ThermoFinnigan TPD/R/O 1100 Series instrument. The NH3-
sis descriptions of ZSM-5 samples are as follows.
TPD measurements were performed by pretreating the zeolites
(100 mg) by heating them to 773 K in NH3 flow (101 kPa,
2.1.1. Synthesis of mesoporous ZSM-5 single crystals 25 cm3 min 1) and holding catalysts under that flow at 773 K for
Meso-ZSM-5 zeolite was synthesized by using a mixture with 1 h before cooling them to ambient temperature.
the molar composition of 3TPAOH:0.5Na2O:0.5Al2O3:25SiO2:1500-
H2O:100EtOH:3PrOH according to following procedure. The 2.2.5. Inductively coupled plasma-atomic emission spectroscopy (ICP-
aluminum isopropoxide was added to an aqueous solution of AES)
tetrapropylammonium hydroxide, water and NaOH. The mixture The quality of synthesized zeolites were checked with induc-
was stirred at room temperature to obtain a clear solution. Then, tively coupled plasma-atomic emission spectroscopy (ICP-AES) to
28 A.A. Rownaghi et al. / Microporous and Mesoporous Materials 151 (2012) 26–33

determine the bulk chemical compositions of the crystals using a reaction temperature in flowing air. To avoid possible condensa-
Perkin–Elmer Emission spectroscopy Model Plasma 1000. tion of hydrocarbons, the temperature of the effluent line was
constantly maintained at 100 °C. An on-line Varian 3800 Gas
Chromatograph with a flame ionization detector (GC-FID)
2.3. Catalytic activity measurements equipped with a capillary column (CP-Sil PONA CB fused silica
WCOT) analyzed the reactor exit gas with hydrogen as carrier
The catalytic measurements were conducted at atmospheric gas. The first sampling of the product gas was done 1 h after the
pressure in a fixed-bed reactor connected to a Gas Chromatograph reactant feed started. During a typical run, subsequent samplings
(GC). The feed (N2) was saturated with methanol (99.9%) in a bub- and analyses were done with 1 h sampling periods. To simplify
ble saturator at 30 °C. The saturated temperature was maintained the discussion, we defined the aromatics as the sum of C6–C10 aro-
by a constant temperature bath. The inlet flow rate was adjusted matics, and the i-paraffins as the sum of C4–C10 i-paraffins.
using a mass flow controller (MFC). The reaction was carried out
in a tubular reactor with an internal diameter of 17 mm and a total
3. Results and discussion
length of 250 mm containing a standard mass of catalyst (i.e.,
0.50 g) at 370 °C with various ranges of weight hourly space veloc-
3.1. Catalyst characterization
ity (WHSV) of 0.16–6.58 g g 1 h–1. For each test, the powder ZSM-5
catalyst was sieved to 250–300 lm and diluted with inert quartz
Fig. 1 illustrates the X-ray powder diffraction (PXRD) patterns of
sand, then loaded to the reactor. The maximum temperature vari-
mesoporous and conventional ZSM-5 catalysts after calcination in
ation along the bed was ±1 °C. The catalyst and sand were mixed in
air for 16 h at 500 °C. The XRD patterns of both ZSM-5 samples
a 1:10 mass ratio. Soda glass beads and quartz glass were used as
show well-resolved diffraction peaks, which are well matched to
catalysts beds during catalytic test. Soda glass beads and quartz
MFI type structure, suggesting that their frameworks contain only
glass were placed on the top and bottom of the catalysts. A thermo-
MFI type zeolite [36]. Table 1 lists the chemical composition of fi-
couple was positioned in the center of the catalyst bed in order to
nal crystals in both mesoporous and conventional ZSM-5 samples.
monitor the temperature. The catalyst was then pretreated in situ
From Table 1, it can be seen that the silicon-to-aluminum ratio
at a heating rate of 5 °C/min under a flow of air (30 mL/min). When
(SiO2/Al2O3) in the ZSM-5 samples determined by ICP-AES was
the reactor temperature reached 500 °C, it was maintained at that
consistently 50 and did not change significantly at different syn-
temperature for 4 h. The temperature was then decreased to the
thesis temperatures and time conditions. This ratio is in agreement
with previously reported data on catalytic properties of solid acid
ZSM-5 zeolite [7,9,28,32].

The acidic properties (content and strength of acid sites) of both
♦ catalysts were evaluated by NH3-TPD technique and the corre-
♦ ZSM-5
sponding profiles are displayed in Fig. 2. As shown in this figure,
three desorption steps were observed, leading to the assumption
that at least three types of acid sites exist. The Con-ZSM-5 catalyst
Intensity/ a.u.

♦ contains three acid sites at 245 °C (weak acid strength), 265 °C



♦ (moderate acid strength), and at 420 °C (strong acid strength). In

the case of Meso-ZSM-5, two distinct desorption peaks were ob-
♦ ♦ ♦
served at 275 and 460 °C in NH3-TPD, which are usually ascribed
meso-ZSM-5
to NH3 desorption from moderate and strong acid sites. From Con-
ZSM-5 to Meso-ZSM-5, it is observed that the moderate peak shifts
to higher temperatures. This shift is accompanied by a gradual in-
crease in the intensity at intermediate temperatures (275 °C). The
Con-ZSM-5
relative values of total acidity of catalysts (i.e., the amount of
ammonia desorbing) correlates well with the amount of aluminum
20 40 60 in the framework for both catalysts, as estimated by ICP-AES, and a
2θ / Degree higher acidity is obtained for Meso-ZSM-5 catalyst. The tempera-
ture at which maximum amount of ammonia was desorbed by
Fig. 1. Powder XRD patterns of mesoporous ZSM-5 single crystals and conventional catalyst is also higher for Meso-ZSM-5 catalyst. Zhang et al.,
ZSM-5. 2h angles between 2° and 70° were collected.
reported that the NH3-TPD desorption at temperatures higher than

Table 1
Chemical and textural properties of ZSM-5 crystals.

Sample SiO2/Al2O3 Crystal lengthb Crystal widthb SBET Smeso SBET/Smeso Vtotal Vmicro Vmeso D
ratiosa (lm) (lm) (m2 g 1 c
) (m2 g 1 d
) (m2 g 1) (cm3 g 1 e
) (cm3 g 1 f
) (cm3 g 1 g
) (nm)h
Meso- 49.2 1.95 1.45 395 214 1.85 0.46 0.08 0.38 9.81
ZSM-5
Con-ZSM- 52.3 2.10 1.30 348 108 3.22 0.20 0.12 0.08 3.0
5
a
Determined by ICP-AES.
b
Estimated from SEM images.
c
Surface areas were obtained by the BET method using adsorption data in p/p0 ranging from 0.05 to 0.25.
d
Measured by t-plot method.
e
Total pore volumes were estimated from the adsorbed amount at p/p0 = 0.995.
f
Measured by t-plot method.
g
Vmeso = Vads,p/p0=99 Vmicro.
h
Average pore diameters were derived from the adsorption branches of the isotherms by the BJH method.
A.A. Rownaghi et al. / Microporous and Mesoporous Materials 151 (2012) 26–33 29

From each XHR-SEM image, approximately 50 ZSM-5 crystals were


Con-ZSM-5 measured and the average size was used as the crystal size (see Table
Meso-ZSM-5 1). As evident from Fig. 3a, Meso-ZSM-5 catalyst constitutes uniform
micro-rectangular crystals with about 1.95 lm  1.45 lm dimen-
Detector response/ a.u.

sions. It is also noted that for this sample, the individual aggregates
are composed of closely packed nanocrystals, whose sizes vary from
2 to 5 nm. Aggregation processes of colloid chemistry derived by
intercrystal forces, rather than the formation of chemical bonds, is
more likely to be responsible for this morphology. In contrast,
Con-ZSM-5 displayed the typical twinned shape of micron-size
ZSM-5 crystals with the dimensions of 2.10 lm  1.30 lm along
the length of the envisioned twinns (Fig. 3d). Surface pitting was fol-
lowed by the attack at the twin boundaries, revealing the curved
junctions of the 90° intergrowth described and discussed in detail
0 100 200 300 400 500 600 700 elsewhere [38,39]. The preparation-dependent morphology ob-
served in the above ZSM-5 syntheses is closely related to the XRD re-
Temperature/ C
sults mentioned earlier.
Fig. 2. NH3-TPD patterns of mesoporous ZSM-5 single crystals and conventional The different morphologies of obtained zeolites were caused to
ZSM-5 catalysts. a large extent by different synthesis conditions even when the
same template was used. Therefore, ZSM-5 zeolites with various
sizes and morphologies can be synthesized with the same template
440 °C can be directly related to the density of Brønsted acid sites by carefully controlling the synthesis conditions. In fact, the
[37]. In addition, desorption at lower temperatures may arise from growth orientation of crystals varies largely with synthesis condi-
both Brønsted and Lewis acid sites. From the results presented in tions giving rise to production of crystals with different morpholo-
Fig. 2, it can be observed that for Con-ZSM-5 catalyst the amount gies. The Meso-ZSM-5 single crystals (Fig. 3a) showed plough-land
of weak-medium acid sites is more than the strong acid sites. roughs with some small furrow like on crystal surface. According
Fig. 3 shows representative morphology of mesoporous ZSM-5 to SEM analysis and in accordance with X-ray diffraction patterns,
single crystals and conventional ZSM-5 crystals. The extreme high all obtained samples appear to be highly crystalline. It can be fur-
resolution-SEM (XHR-SEM) images, shown in Fig. 3 clearly indicate ther seen from high magnification images in Fig. 3b and c that the
that the crystals obtained by different techniques have significantly Meso-ZSM-5 crystals have hierarchical rectangular architecture,
different crystal morphology, crystal size, and size distribution. assembled by many stone pavement-like crystals of 2–5 nm in

Fig. 3. XHR-SEM images of (a–c) mesoporous ZSM-5 single crystals (three different magnifications), and (d) conventional ZSM-5 crystals.
30 A.A. Rownaghi et al. / Microporous and Mesoporous Materials 151 (2012) 26–33

300 micro- and mesopore volumes and pore size) are summarized in
Table 1. The BET surface area obtained from the adsorption iso-
therm includes the contributions from both the internal and the
250 external surfaces. The Meso-ZSM-5 zeolite with higher surface area
Volume Adsorbed (cm 3/g STP)

Meso-ZSM-5 and larger mesopore volume (Smeso = 214 m2 g 1 and


Vtotal = 0.46 cm3 g 1, see Table 1) displayed significantly higher
200 Con. ZSM-5 BET surface area of 395 m2 g 1. This value is very close to the value
reported by Abelló et al. for the same structure [41]. Kustova et al.
[42] synthesized a series of mesoporous samples consisting of a
150
ZSM-5 core and a silicalite shell and additional desilicated of mes-
oporous ZSM-5 samples resulted in mesopore volume of 0.39, 0.25
and 0.47 cm3 g 1, respectively. Kim et al. [28] reported the prepa-
ration of mesoporous ZSM-5 with mesopore volume of
100
0.32 cm3 g 1 via hydrothermal synthesis with the addition of orga-
nosilane surfactants. In another study, Christensen et al. [21,22,32]
compared the mesoporous ZSM-5 single crystals prepared by car-
50
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 bon-templating technique to a conventional ZSM-5 sample in
MTG reaction. Same authors reported the synthesis of mesoporous
Relative Pressure (P/P 0)
ZSM-5 single crystals with mesopore volume of 0.31 cm3 g 1 in
Fig. 4. N2 adsorption/desorption isotherms of mesoporous ZSM-5 single crystals
fluoride media [43]. Mesopore volume of 0.82 cm3 g 1 for meso-
and conventional ZSM-5 zeolites. porous ZSM-5 was achieved by Park et al. [27] using carbon-tem-
plating and microwave heating techniques. The degree of
mesoporosity in Meso-ZSM-5 zeolite strongly depends on the mes-
width. In addition, the stone pavement-like crystals are indeed opore volume as shown in Table 1. The N2-sorption data indicate
assembled layer-by-layer (see Fig. 3c). It is found that the obtained that the BET and mesopore surface areas of the zeolite samples de-
mesostructured crystals (ca. 2–5 nm in diameter) consist of arrays pend strongly on the synthesis mixture preparation method and
of parallel nanorods. The results obtained by both XHR-SEM and the hydrothermal treatment conditions.
ICP techniques are in agreement with that of XRD patterns. The In the case of conventional catalyst (Con-ZSM-5), the BET sur-
uniform mesoporous ZSM-5 single crystals are unique in the sense face area, mesopore surface area and total pore volume were esti-
that they contain both inter-connected micropores and mesopores mated to be 348 m2 g 1, 108 m2 g 1 and 0.20 cm3 g 1,
inside each individual single crystal. respectively. These values differ largely with those of novel
The N2 adsorption data of the mesoporous ZSM-5 catalyst pre- mesoporous zeolite. It is clear that the micropore volume (Vmicro)
sented in Table 1 indicate the presence of mesopores that is rea- of the Con-ZSM-5 catalyst is 1.5-fold higher than the micropore
sonably assigned to inter-crystalline voids between the stone volume of Meso-ZSM-5 catalyst (0.08 and 0.12 cm3 g 1, see Table
pavement-like crystals. Correspondingly, the average pore size 1). No significant dependence of micropore volume on the crystal
gives a diameter at about 10 nm. This indicates that ZSM-5 ob- size or relative crystallinity of the samples was detected. The
tained in this work has indeed a hierarchical structure (as also Meso-ZSM-5 catalyst displays as 2-fold higher mesopore surface
shown by XHR-SEM). It is believed that the preparation of synthe- area (214 m2 g 1) as Con-ZSM-5 catalyst (108 m2 g 1). Mesoporous
sis mixture and hydrothermal reaction conditions are key factors surface area of 75–244 m2 g 1 was achieved by Caicedo-Realpe and
influencing crystal morphology, surface area (internal and exter- Perez-Ramirez [33] using a two-step route comprising sodium alu-
nal), porosity and crystal size. Since inter-crystalline porosity is minate and acid treatments of ZSM-5 zeolite. In addition, it can be
developed, this simple method opens novel avenues for catalyst noted that, as mesopore surface area (Smeso) increases from 108 to
design by improved diffusion characteristics, and thus a better uti- 214 m2 g 1, mesopore volume (Vmeso) also enhances from 0.08 to
lization of the zeolite crystal for catalytic applications. Based on 0.38 cm3 g 1 indicating a linear relation between Smeso and Vmeso.
these results, the novel mesoporous ZSM-5 catalyst is potentially Average pore diameters of crystals were obtained by applying
a suitable candidate in various catalytic applications. the BJH model from nitrogen adsorption data listed in Table 1.
Fig. 4 shows different types of N2 adsorption/desorption iso- The Con-ZSM-5 zeolite has smaller (3 nm) pores than Meso-
therms corresponding to the prepared zeolites with SiO2/Al2O3 ra- ZSM-5 (with an average pore size of 10 nm). The later pore size,
tio of 50. Meso-ZSM-5 sample exhibited a type-IV isotherm [40], which is in the range anticipated for hydrocarbon cracking, is
almost similar to that of Con-ZSM-5 sample at low-pressure re- unprecedented among all previously reported mesoporous forms
gions, but with significant increase in adsorption amounts with of MFI zeolite structure.
pressure. Con-ZSM-5 sample, on the other hand, displayed a The characteristics of the final mesoporous zeolite are impor-
type-I isotherm (plateau at relative pressure and without hystere- tant in terms of nature of framework, Si/Al ratio, and distribution
sis), which is usually observed for microporous materials having of metal species in the crystal volume. The mesoporosity is intro-
relatively low external surfaces [40]. Characteristic features of duced into the zeolite crystals directly during the crystallization
type-IV isotherm are basically its hysteresis loop, which is associ- of the zeolite by having intercrystalline pores or voids. At the
ated with capillary condensation taking place in mesopores, and beginning of synthesis, amorphous materials in the synthesis mix-
the limiting uptake over a range of high p/p0 = 0.33. The Meso- ture begin to form zeolite nanocrystals. These nanosized crystals
ZSM-5 showed a type-H4 hysteresis loop, which is often associated tend to aggregate and self-assemble into uniform-shaped polycrys-
with porous materials with narrow slit like pores. This finding is in talline particles. Constant stirring applied during the synthesis pro-
accordance with SEM images obtained from the same sample. The cess results in the formation of uniform ZSM-5 crystals. The
isotherm corresponding to Meso-ZSM-5 crystal indicates that this intercrystalline spaces in nanocrystals are directly accessible from
zeolite is mostly mesoporous as large amounts of nitrogen adsorp- the outer surface of the zeolite. In this case, the condition that mes-
tion were observed at high p/p0. Meso-ZSM-5 possesses a total pore opores enhance the mass transport to/from the active sites in the
volume of 0.46 cm3 g 1 measured at p/p0 = 0.995. The textural zeolite materials has been satisfied [32]. The highly crystalline
properties of both zeolites (i.e., BET and mesoporous surface areas, mesoporous zeolite synthesized in this work has in fact the crystal-
A.A. Rownaghi et al. / Microporous and Mesoporous Materials 151 (2012) 26–33 31

during the initial reaction time. However, as reaction proceeds, it


100
can be clearly seen that the mesoporous ZSM-5 catalyst is more
Methanol Conversion (wt.%)

stable than the conventional ZSM-5 catalyst indicated by higher


80
methanol conversion rate. In the case of Con-ZSM-5 catalyst, the
methanol conversion started to decrease after an initial reaction
60 period of 10 h. Hereafter the catalytic lifetime is defined as the
time at which the catalytic conversion decreases by 50% (i.e., t1/2)
40 [44,45]. Thus, the t1/2 value obtained for conventional catalyst
(Con-ZSM-5) was 30 h (see Fig. 5). Unlike the conventional ZSM-
20 5 catalyst, mesoporous ZSM-5 catalyst is more stable (up to 95%)
over the same reaction time. It is worth noting that the comparison
0 of catalysts lifetime was performed under similar SiO2/Al2O3 ratios
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32
Time on stream/ h because the catalyst physicochemical properties and lifetime are
more likely to change by Al-content [44–46]. These results show
Fig. 5. Effect of mesoporosity on catalytic activity (deactivation curves); methanol that the catalytic lifetime is enhanced up to 2-fold for our meso-
conversion vs time-on-stream for both catalysts (370 °C, 1 atm, WHSV = 1.05 g g 1 h). porous ZSM-5 catalyst mainly due to the generation of mesoporos-
ity. Thus, we did succeed in producing a ZSM-5 catalyst with high
line form of MFI zeolite with both micropores and mesopores in its
activity and stability.
structure. Adjusting the composition of synthesis mixture and The catalyst activity is determined by the acidity, whereas the
hydrothermal synthesis time and temperature could easily modu-
catalyst stability (deactivation behavior) is determined by the acid
late the composition and mesoporosity of zeolite. For instance, un- site accessibility and diffusion rate, that is not correlated with the
der the conditions studied here, synthesis times ranging from 18 to
Brønsted acidity of the ZSM-5 crystals. The ratio of the BET and
72 h and synthesis temperatures ranging from 120 to 170 °C are external surface areas for two catalysts are presented in Table 1.
optimum conditions for obtaining highly mesoporous single crys-
As can be seen, a surprisingly clear correlation exists between
tals. These materials are an emerging and important class of cata- the deactivation rate (Fig. 5) and the SBET/Smeso ratio, indicating
lysts for a variety of reactions. Using this method to donate
that the higher mesopore surface area affects the catalyst stability
uniform mesopores to ZSM-5 single crystals should significantly and deactivation rate. This leads to the conclusion that the param-
extend its applications because such uniform single crystals with
eter determining catalyst deactivity is related to the acid site
additional mesoporosity lead to easier access to the active sites accessibility as well as the diffusion rate of reactants and products.
as well as resistance to cock formation and consequently catalyst In general, the interest in mesoporous ZSM-5 is motivated by the
deactivation. desire to improve the mass transport of reactants and products to
and from the active sites located inside the ZSM-5 micropores.
3.2. Catalytic performance in methanol dehydration reaction Schematic illustration of concentration profiles of a reagent
through mesoporous and conventional crystals are shown in
Fig. 5 shows an overview of the activities and deactivation Fig. 6. For a catalytic reaction involving reactants and products
behavior (catalyst stability) of two catalysts described and with relatively limited diffusion reaction, it is not possible to fully
discussed in this contribution. As evident from this figure, the cor- utilize the entire ZSM-5 crystals unless intracrystalline mesopores
responding conversion rates are very close to each other during the are introduced. As proposed by Bibby et al. for conventional ZSM-5
initial reaction period, implying similar extent of deactivation catalyst, the initial deposition of coke takes place inside the zeolite

Meso-ZSM-5 Single Crystal

Results
Methanol -higher activity and selectivity
-Light olefins
-Alkyl-aromatics

Reaction zone

Con-ZSM-5 Crystals

Results
-Lower activity and selectivity
-Wide-range hydrocarbons
-Coke
Methanol Unutilized part of
ZSM-5 crystals

Fig. 6. Combination of SEM-images and schematic, illustration of the diffusion pass rate of methanol through conventional and mesoporous ZSM-5 single crystals.
32 A.A. Rownaghi et al. / Microporous and Mesoporous Materials 151 (2012) 26–33

75
70
65
60

distribution, [wt%]
55
Dimethyl ether
50
Light paraffins (C1-C4)
45
40 Ligh olefins (C2-C3)
35
Other paraffins and olefins (C4+)
30
Alkyl aromatics (BTX)
25
Hydrocarbon

20
15
10
5
0
-5
0.1 1 10
WHSV, [g/g hr]

Fig. 7. Effect of space velocity on hydrocarbon distribution over mesoporous ZSM-5 single; selectivity vs space time (370 °C, 1 atm, methanol partial pressure: 0.21 atm).

channels [47,48] and results in an inactive or inaccessible zone in The results at the highest space velocity (WHSV = 6.58 g g 1 h)
the center of the crystals. Therefore, only the outer surface area of show that the product stream contains a relatively large amount of
the crystal is accessible to the reactant. In addition, the polyme- dimethyl ether (DME) formed by the reversible dehydration of
thylated benzenes formed inside the center of the crystal (unuti- methanol. However, the presence of a small fraction of DME at high
lized part in Con-ZSM-5 in Fig. 6) do not reach the gas phase, space velocity implies that the contact time was not sufficient to
and hence a fast deactivation occurs. On the contrary, for meso- convert all DME to light olefins. Based on the data, the DME
porous ZSM-5 single crystals, the polymethylated benzenes accu- forming reaction might be related as either parallel or sequential
mulation in the catalyst crystals is slower, because the transfer to the major reaction steps leading to hydrocarbons. Running the
to the gas phase is more efficient as a consequence of both a larger reaction at the highest velocity and lowest conversion (38% at
total surface area and the shorter distance to be covered by diffu- WHSV = 6.58 g g 1 h) gives rise to a 4% dimethyl ether,
sion. Given that the diffusion becomes faster with an increasing hydrocarbon distribution of 35% primary hydrocarbons (paraffins
external surface area, it is found that crystals with a high external and olefins), and 40% alkyl aromatics (benzene, toluene, and
surface area (Meso-ZSM-5) are expected to deactivate slow, which xylene). These results indicate that higher space velocity is required
agrees qualitatively with the observed trends in Fig. 5. for an efficient conversion of methanol to dimethyl ether over mes-
When the activity of two catalysts having similar Al-contents oporous ZSM-5 catalysts.
were correlated with the acid content and external surface areas, The light olefins formed during first steps then undergo conden-
a good correlation was observed between catalytic activity and sation and oligomerization over mesoporous zeolite catalyst, lead-
strong acidic sites on external surface area (see Fig. 5 and Table ing to formation of final alkyl aromatic products. The conversion of
1). The partial destruction of the crystalline structure improves methanol was completed between 0.16 and 1.05 WHSV, with the
the accessibility to the intra-crystalline active sites by large mole- higher formation of light olefins and substantial amounts of aromat-
cules. This favorably affects the catalytic reactivity of mesoporous ics. The highest selectivity for light olefins (17%) was reached at
ZSM-5 catalyst. Therefore, it is evident that introducing mesopo- 370 °C at longest contact time, although the selectivity for light par-
rosity in ZSM-5 structure can lead to a significant improvement affins was increased with enhanced space velocity. The existing
in the methanol dehydration reaction. hydrocarbons exhibit a relatively narrow range of molecular
As earlier findings reported by Bibby et al., indicate, coke forma- weights, terminating abruptly at about C6–C10, which correspond
tion occurs much more rapidly in a microporous ZSM-5 catalyst to a boiling point range of gasoline. Alkyl aromatics were the main
than in a mesoporous zeolite during the hydrocarbons reaction products at various contact times. Essentially no hydrocarbons
[47,48]. Therefore, coke in ZSM-5 catalyst is preferentially gener- above C10 were produced. These results imply that mesoporosity
ated inside the micropores prior to the deposition at the external not only plays a dominant role in stabilizing the lifetime of the cat-
surface, demonstrating that external coking does not play a domi- alyst but also in increasing the selectivity to alkyl aromatics. As
nant role when a significant amount of internal defects is present shown in Table 2, the selectivity to alkyl aromatics exceeds 50%
at the same time. This can be verified in observing a significant dif- while product distributions are very similar, in which light olefins
ference in the color of the two catalysts after 30 h reaction. and aromatics are main products at low space velocity. Furthermore,
Fig. 7 shows more complete collections of hydrocarbon distribu- it can be seen from Fig. 7 that the selectivity to aromatics and light
tion vs WHSV with five orders of magnitude. The corresponding data olefins decreased gradually with increasing WHSV. On the other
are also listed in Table 2. The hydrocarbon products are very similar, hand, the selectivity to aliphatic hydrocarbons increased step-by-
suggesting a common reaction pathway over various contact times. step. At the same space velocity, the conversion of methanol dramat-
The primary reaction is dehydration of methanol to dimethyl ether ically decreased from 100% (at low WHSV) to 38% (highest WHSV).
and water [5]. Methanol and dimethyl ether further react to give Through the choice of space velocity, the reaction can be tailored
light olefins (secondary reaction), which then are transformed to to maximize alkyl aromatics (BTX), light olefins (ethylene and pro-
higher olefins, aromatics and paraffins. The detailed mechanism of pylene), or liquefied petroleum gas (LPG). Moreover, due to shape-
this process has been discussed elsewhere [49]. selective constraints of ZSM-5 catalyst, this process yields a high
A.A. Rownaghi et al. / Microporous and Mesoporous Materials 151 (2012) 26–33 33

Table 2
Effect of space velocity on methanol dehydration and hydrocarbon distribution over mesoporous ZSM-5 single crystals (370 °C, methanol partial pressure: 0.21 atm).

WHSV [g of liquid methanol/(g of Methanol conversion Hydrocarbon distribution (%)


catalyst/h)] (%)
Paraffins Olefins Dimethyl Paraffins and Olefins Alkyl Aromatics Othersa
(C1–C4) (C2–C3) ether (C4+) (BTX)
0.16 100 14.6 16.6 – 3.2 65.5 0.1
0.43 100 19.0 12.1 – 7.7 61.0 0.2
1.05 100 22.8 8.0 – 12.0 57.0 0.2
2.63 40 26.9 3.0 – 19.2 50.5 0.4
6.58 38 33.6 1.3 4.0 21.2 39.4 0.5
a
Including CO, CO2 and coke.

quality hydrocarbon and minimizes the production of branched [4] F.J. Keil, Micropor. Mesopor. Mater. 29 (1999) 49.
[5] C.D. Chang, A.J. Silvestrl, J. Catal. 47 (1977) 249.
products.
[6] M. Stocker, Micropor. Mesopor. Mater. 29 (1999) 3.
[7] A.A. Rownaghi, F. Rezaei, J. Hedlund, Catal. Comm. 14 (2011) 37.
4. Conclusions [8] A.T. Aguayo, D. Mier, A.G. Gayubo, M. Gamero, J. Bilbao, Ind. Eng. Chem. Res. 79
(2010) 12371.
[9] A.A. Rownaghi, J. Hedlund, Ind. Eng. Chem. Res. 50 (2011) 11872.
The preparation of uniform mesoporous MFI single crystals [10] H. Koempel, W. Liebner, Stud. Surf. Sci. Catal. 167 (2007) 261.
using a novel technique was reported in this study. The role of [11] J.Q. Chen, A. Bozzano, B. Glover, T. Fuglerud, S. Kvisle, Catal. Today 106 (2005)
103.
mesoporosity of ZSM-5 catalyst was investigated in conversion of [12] J. Topp-Jørgensen, Surf. Sci. Catal. 36 (1988) 293.
methanol to hydrocarbons at rather mild conditions, i.e., 370 °C [13] P. Li, W. Zhang, X. Han, X. Bao, Catal. Lett. 34 (2010) 124.
temperature and ambient pressure. The mesoporous ZSM-5 zeolite [14] W.J.H. Dehertog, G.F. Froment, Appl. Catal. 71 (1991) 153.
[15] V.R. Choudhary, V.S. Nayak, Zeolites 5 (1985) 325.
discussed in the present study exhibits several attractive features,
[16] S. Donk, A.H. Janssen, J.H. Bitter, K.P. Jong, Catal. Rev. 45 (2003) 297.
such as large BET and mesoporous surface area, controlled pore [17] F. Rezaei, P. Webley, Sep. Purif. Technol. 70 (2010) 243.
size, large mesopore volume and low cost and facile synthesis pro- [18] K. Wandelt, Surf. Sci. 387 (1991) 251.
[19] J. Maier, Angew. Chem. Int. Ed. 32 (1993) 528.
cedure. It was found that the deactivation rate correlates well with
[20] Y. Tao, H. Kanoh, L. Abrahms, K. Kaneko, Chem. Rev. 106 (2006) 896.
the BET and external surface area ratio. A ZSM-5 catalyst with a [21] C.H. Christensen, K. Johannsen, E. Törnqvist, I. Schmidt, H. Topsøe, Today 128
high SBET/Smeso ratio is expected to show a slow deactivation rate (2007) 117.
in the conversion of methanol to hydrocarbons. Introducing meso- [22] C.H. Christensen, I. Schmidt, C.H. Christensen, Catal. Comm. 5 (2004) 543.
[23] A.L. Kustov, K. Egeblad, M. Kustova, T.W. Hansen, C.H. Christensen, Top. Catal.
porosity to ZSM-5 crystals increases external surface area resulting 45 (2007) 159.
in a faster diffusion of the hydrocarbon products in the ZSM-5 [24] A.L. Kustov, T.W. Hansen, M. Kustova, C.H. Christensen, Appl. Catal. B 76 (2007)
channels, and a correspondingly slower deactivation rate. The pre- 311.
[25] J. Perez-Ramırez, S. Abello, A. Bonilla, J.C. Groen, Adv. Funct. Mater. 19 (2009)
sented data clearly show that apart from the commonly recognized 164.
deactivation of ZSM-5 due to coking at the external surface, the [26] J.C. Groen, L.A.A. Peffer, J. Perez-Ramırez, Micropor. Mesopor. Mater. 60 (2003)
deactivation via coking at internal pores of ZSM-5 plays an impor- 1.
[27] J.B. Koo, N. Jiang, S. Saravanamurugan, M. Bejblova, Z. Musilova, J. Cejka, S.E.
tant role. Deactivation of catalyst occurred much slower in the case Park, J. Catal. 276 (2010) 327.
of mesoporous ZSM-5 than conventional ZSM-5, mainly due to fas- [28] J. Kim, M. Choi, R. Ryoo, J. Catal. 269 (2010) 219.
ter diffusion of products from the zeolite channels, hence reducing [29] M. Rivallan, G. Ricchiardi, S. Bordiga, A. Zecchina, J. Catal. 264 (2009) 104.
[30] B. Louis, F. Ocampo, H.S. Yun, J.P. Tessonnier, Chem. Eng. J. 161 (2010) 397.
the formation of coke. This resulted in simultaneous improvement
[31] J. Cejka, S. Mintova, Catal. Rev. Sci. Eng. 49 (2007) 457.
of catalytic activity and selectivity towards alkyl aromatics. Cou- [32] M.S. Holm, E. Taarning, K. Egeblad, C.H. Christensen, Catal. Today 168 (2011) 3.
pled with existing synthesis gas-based biomethanol technologies, [33] R. Caicedo-Realpe, J. Perez-Ramirez, Micropor. Mesopor. Mater. 128 (2010) 91.
[34] M. Ogura, Catal. Surv. Asia 12 (2008) 16.
these results are very promising in developing a highly efficient
[35] J. Hedlund, O. Öhrmana, L.V. Msimang, E.V. Steenb, W. Böhringer, S. Sibya, K.
alternative process for practical applications in renewable fuel Möller, Chem. Eng. Sci. 59 (2004) 2647.
products and chemicals. These results suggest that the use of mes- [36] Ch. Baerlocher, L.B. McCusker, D.H. Olson, Atlas of Zeolite Framework Types,
oporous ZSM-5 catalyst prepared in this study, will significantly re- Elsevier, 2007.
[37] W. Zhang, E.C. Burckle, P.G. Smirniotis, Micropor. Mesopor. Mater. 33 (1999)
duce the coke formation and enhance production of alkyl 173.
aromatics from methanol at mild operation conditions. [38] J.R. Agger, N. Hanif, C.S. Cundy, A.P. Wade, S. Dennison, P.A. Rawlinson, M.W.
Anderson, J. Am. Chem. Soc. 125 (2003) 830.
[39] D.G. Hay, H. Jaeger, K.G. Wilshier, Zeolites 10 (1990) 571.
Acknowledgments [40] K.S.W. Sing, D.H. Everett, R.A.W. Haul, L. Moscou, R.A. Pierotti, J. Rouquerol, T.
Siemieniewska, Pure Appl. Chem. 57 (1985) 603.
The Swedish Energy Agency, Knut and Alice Wallenberg foun- [41] S. Abello, A. Bonilla, J. Perez-Ramirez, Appl. Catal. A-Gen. 364 (2009) 191.
[42] M. Kustova, M.S. Holm, C.H. Christensen, Y.H. Pan, P. Beatoa, T.V.W. Janssensa,
dation is acknowledged for financially supporting the Magellan F. Joensen, J. Nerlov, Stud. Surf. Sci. Catal. 174 (2008) 117.
SEM instrument. [43] K. Egeblad, M. Kustova, S.K. Klitgaord, K. Zhu, C.H. Christensen, Micropor.
Mesopor. Mater. 101 (2007) 214.
[44] K.G. Ione, G.V. Echevskii, G.N. Nosyreva, J. Catal. 85 (1984) 287.
References
[45] S.M. Campbell, D.M. Bibby, J.M. Coddington, R.F. Howe, J. Catal. 161 (1996) 350.
[46] D.B. Luk’yanov, Zeolites 12 (1992) 287.
[1] S. Czernik, A.V. Bridgwater, Energ. Fuel. 18 (2004) 590. [47] D.M. Bibby, N.B. Milestone, J.E. Patterson, L.P. Aldridge, J. Catal. 97 (1986) 493.
[2] D. Mohan, C.U. Pittman, P.H. Steele, Energ. Fuel. 20 (2006) 848. [48] D.M. Bibby, C.G. Pope, J. Catal. 116 (1989) 407.
[3] J. Cobb, in: G. Connell (Ed.), New Zealand Synfuel: The Story of the World’s First [49] S.A. Tabak, S. Yurchak, Catal. Today 6 (1990) 307.
Natural Gas to Gasoline Plant, Cobb/Horwood Publications, Auckland, New
Zealand, 1995, p. 1.

You might also like