Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Microporous and Mesoporous Materials 291 (2020) 109699

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

Modification of H-[B]-ZSM-5 zeolite for methanol to propylene (MTP) T


conversion: Investigation of extrusion and steaming treatments on
physicochemical characteristics and catalytic performance
Maryam Sadat Beheshtia, Mahdi Behzada,∗, Javad Ahmadpourb, Hasan Arabic
a
Semnan University, P.O. Box:19111-35131, Semnan, Iran
b
Department of Chemical Engineering, Babol Noshirvani University of Technology, P.O. Box: 47148-71167, Babol, Iran
c
Group of Polymerization Catalyst, Iran Polymer and Petrochemical Institute, P.O. Box: 13115-14977, Tehran, Iran

A R T I C LE I N FO A B S T R A C T

Keywords: High-silica H-[B]-ZSM-5 zeolite was synthesized by a hydrothermal method and was then extruded by a con-
H-[B]-ZSM-5 ventional process. The influence of mesoporosity formation and acidity modification was investigated using
Zeolite extrusion and combined extrusion-steam treatment over the synthesized zeolite. The catalysts were character-
Extrusion ized by BET, FE-SEM, TGA, PXRD and NH3-TPD techniques. The catalytic performance of the powder H-[B]-
Steam treatment
ZSM-5 (BZP), extruded (BZE) and steam-treated extruded (BZEH) samples was evaluated in the methanol to
Methanol to propylene (MTP) reaction
propylene (MTP) conversion reaction. The reactions were performed in a fixed-bed reactor at 480 °C using a feed
containing a mixture of 50 wt% methanol in water with a methanol weight hourly space velocity (WHSV) of 0.9
h−1. The results showed that extrusion of H-[B]-ZSM-5 with alumina sol as a binder followed by steam-treatment
led to the formation of narrow and uniform mesoporosity without severely destructing the crystal structure. The
life time of the BZEH catalyst (750 h) was higher than that of BZE (520 h) and BZP (580 h) samples. This
considerable enhancement of the lifetime could be attributed to the synergetic effect of extruding and steaming
of H-[B]-ZSM-5 zeolite, leading to weakening of the acid sites strength and reduction of strong/weak acid sites
ratio as well as formation of mesoporous structure, which are helpful for attenuating coke deposition in the
micropore channels even at a high coking level. Besides, BZEH exhibited higher selectivity to propylene, bu-
tylenes and total light olefins, while its selectivity to C1–C4 alkanes and ethylene was relatively lower. The
insights attained in this work would help to design a promising catalyst for methanol to propylene process.

1. Introduction well as long catalyst lifetime [1,7,8]. However, the micro porosity and
high extent of strong acidity of the H-ZSM-5 zeolite result in relatively
Propylene, as one of the most important commodity petrochemicals, easy coke formation and deactivation of catalysts during this acid-cat-
is mainly produced as a by-product of both the petroleum refinery in alyzed reaction process [9,10]. In order to overcome this drawback,
FCC units and the ethylene production by the naphtha steam cracking numerous studies have been focused on proper tuning of the acidity
process. However, due to the increasing global demand of propylene (strength and density) [11,12], size and/or morphology-controllable
and rapid depletion of petroleum reserves, new processes with high synthesis [13–16] and fabrication of hierarchical pore structure of the
propylene yield are required. Methanol to propylene (MTP) process is H-ZSM-5 zeolite [17–20].
regarded as a promising alternative route for producing propylene from For H-ZSM-5 zeolite powder, variation of framework Si/Al ratio via
non-petroleum resources, since methanol can be industrially produced pre- or post-synthesis methods is a generally adopted procedure for
from natural gas and coal in large scale [1–4]. Lurgi's MTP process, adjusting the zeolite acidity [21–25]. Moreover, promoter incorpora-
based on an H-ZSM-5 catalyst in a fixed-bed reactor, has been com- tion is an efficient method to modify the acidity of the zeolite, thereby
mercialized since 2010 in China. And up to now, three MTP plants have improving the catalytic performance of the H-ZSM-5 catalyst in the
been constructed in China using this technology [4–6]. MTP reaction [26–31]. In recent years, many researchers have re-
Currently, high silica H-ZSM-5 zeolite is the preferred catalyst for peatedly reported that boron-modified H-ZSM-5 shows improved MTP
practical MTP process due to its high selectivity towards propylene as catalytic performance compared to the unmodified zeolite. The


Corresponding author.
E-mail address: mbehzad@semnan.ac.ir (M. Behzad).

https://doi.org/10.1016/j.micromeso.2019.109699
Received 27 April 2019; Received in revised form 15 August 2019; Accepted 2 September 2019
Available online 04 September 2019
1387-1811/ © 2019 Elsevier Inc. All rights reserved.
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

considerable enhancement of the catalyst lifetime could be attributed to catalyst upon steaming, as well as its catalytic performance during MTP
the reduction of strong/weak acid sites ratio, avoiding various hy- process.
drogen-transfer reactions producing aromatics [1,11,32,33]. Hence, the aim of the present work is to investigate the effect of the
However, for industrial application, the powder form of zeolite extrusion and steam treatments on the physicochemical properties and
needs to be extruded with a matrix, usually a metal oxide such as catalytic performance of H-[B]-ZSM-5 extruded catalyst in the MTP
alumina in order to achieve the necessary mechanical strength for in- reaction using a fixed-bed flow reactor under the same operating con-
dustrial applications. The binders may affect catalyst acidity, porosity, ditions. The physicochemical properties of these catalysts were char-
thermal properties, and catalyst coking behaviors. Hence, information acterized by BET, FE-SEM, TGA, XRD and NH3-TPD methods. Useful
on the influence of binder on the physicochemical properties and cat- insights concerning the synergetic effect of extruding and steaming of
alytic performance of zeolite is very important for developing an in- H-[B]-ZSM-5 zeolite were gained. It is proved that the hydrothermally-
dustrial catalyst [34–36]. Nevertheless, a few works have been done on treated extruded H-[B]-ZSM-5 is a promising catalyst for MTP reaction.
the effect of binder on activity, selectivity and stability of H-ZSM-5 This research is part of an extensive effort to prepare a suitable H-[B]-
zeolite during methanol-to-propylene reaction. Lee et al. [37] in- ZSM-5 catalyst for the methanol conversion to light olefins especially
vestigated the influence of single and binary binder system on the with high propylene selectivity in commercial scale, which is im-
acidity, and catalytic performance of H-ZSM-5 zeolite in MTP process. plemented by Petrochemical Research and Technology Company.
They observed that the propylene selectivity dramatically increased
from 23.8% to around 40% when the H-ZSM-5 zeolite bound with 10 wt 2. Experimental
% of aluminum phosphate single binder, or with the binary binder
system comprising aluminum phosphate-silica and aluminum phos- 2.1. Materials
phate-alumina. This is explained by the decrease in the number and
strength of strong acid sites due to the addition of binders. Freiding All the reagents, including silicic acid (SiO2.xH2O, ≥ 99 wt%), so-
et al. [38] also compared the catalytic performances of AlPO4/H-ZSM-5 dium aluminate (NaAlO2, Al2O3 wt.% = 55), boric acid (H3BO3, 99.8 wt
extrudates with conventional alumina- and silica-bound extrudates in %), tetrapropylammonium bromide (TPABr, C12H28BrN, ≥99 wt%),
the fixed-bed MTO reaction, and concluded that at temperatures above ammonium nitrate (NH4NO3, 99 wt%), sodium hydroxide (NaOH,
400 °C, AlPO4/H-ZSM-5 extrudates show superior activity, propylene 99.6 wt%) and sulfuric acid (H2SO4, 98 wt%) were purchased from
selectivity, and stability. Despite these superior results, the binder Merck Company and were used in the hydrothermal synthesis of the
might also partially block the zeolite micropore mouths, which lead to parent H-ZSM-5 sample. Aluminum oxide sol (Alumina sol) was used as
decreasing the accessibility of reactants to the active sites. So, there is a binder. The content of Al2O3 of alumina sol was 20–21%. Alumina sol
still a great challenge to solve the issues for creating a highly effective was purchased from AIOTEC Company. All the materials were extra
extrudates MTP catalyst. The hydrothermal post-treatment is a simple pure and were used as-received without any further treatments.
method to modify the textural properties and the structure of the ex-
truded catalyst. Moreover, in the industrial MTP process, methanol 2.2. Catalyst preparation
conversion on H-ZSM-5 extrudates progresses in a long run under hy-
drothermal conditions with a water co-feed [39–42]. Hence, the novelty The H-[B]-ZSM-5 sample was prepared by hydrothermal synthesis
of this research work is to study the influence of steaming on the tex- according to the following procedure. The calculated amounts of so-
tural property and catalytic performance of the extruded catalyst. It is dium aluminate, silicic acid, sodium hydroxide, tetra-propyl ammo-
of particular importance to mention that the extruded materials are nium bromide, boric acid and deionized water were mixed well to form
used in industrial reactors under the actual operating conditions in gel under fast stirring for 3 h. Concentrated sulfuric acid was used to
which a mixture of methanol and water is fed. adjust the gel pH. The molar ratio of the starting gel was
It is known that steaming leads to dealumination of the zeolite 20SiO2:0.05Al2O3:0.05B2O3:3TPABr:1.5Na2O:200H2O. Crystallization
structure, and is one of the feasible techniques to adjust the acidity of was carried out at 180 °C under autogenously pressure for 48 h with
H-ZSM5 zeolites for tailoring catalytic properties. Moreover, the deal- continuous stirring in a stainless-steel autoclave. The resultant solid was
umination of H-ZSM-5 zeolites by steam treatment is a suitable meth- recovered by filtration, washed several times with deionized water until
odology to shorten the effective diffusion pathways and enhance the the pH value of the decanted water reached 7, and then dried overnight
accessibility of acid sites via the creation of mesopores [21,22,43,44]. at 105 °C. The catalyst samples were calcined to remove the organic
So far, there are few literature studies that discuss the influence of template in a muffle furnace under air flow at 530 °C for 12 h at a
binder on the hydrothermal stability of extruded catalysts heating rate of 3 °C/min. Finally, the solid powder was ion-exchanged
[39–41,45,46]. Zhang et al. [40] have characterized and evaluated with 1 mol/L NH4NO3 solution at 90 °C for 10 h under continuous
hydrothermally treated H-ZSM‐5 powders and extrudates by varying agitation, followed by filtration and washing. This procedure was re-
temperatures and times in the MTP process, demonstrating that under peated four times to obtain NH4-[B]-ZSM-5 zeolite. After drying at
the optimized steaming conditions, the reduction of Brønsted acidity 105 °C for 12 h, the NH4-[B]-ZSM-5 zeolite was calcined at 530 °C for
content of dealuminated catalysts in powder or extruded form leads to 12 h (at a heating rate of 3 °C/min) to obtain the H-form of the zeolite.
increases in the propylene selectivity and catalytic lifetime. Chang et al. This sample was named as BZP. The resulting powders were formed by
[46] reported that the activity of high-silica H-ZSM-5 and B-ZSM-5 for tableting and then 16–25 mesh catalysts were made by crushing and
n-hexane cracking reaction increased by hydrothermal treatment in the sieving for catalytic evaluation in the reactor.
presence of alumina, due to transport and insertion of Al into the zeolite The extruded catalyst was prepared by compositing BZP with alu-
framework in tetrahedral coordination. It is further shown that Al mina binder by the following procedure: (1) physical mixing of BZP
substitution is facilitated by the presence of framework B because of the with hydroxyethyl cellulose and then milling by ball mill for 60 min, (2)
hydrolytic instability of B and hence its ready replacement. More re- pasting the above-mentioned mixture by H2O and alumina sol
cently Song et al. focused on the changes in properties of phosphorus- (70:20:10 wt%), (3) intermixing and extrusion, (4) drying at 105 °C for
modified H-ZSM-5 extruded catalyst during steam treatment. The re- 12 h and then calcination at 540 °C for 12 h under flowing of zero-air.
sultant catalyst exhibited an enhanced selectivity to propylene and a Finally, the extruded sample was sieved into 16–25 mesh size and the
prolonged lifetime which was due to its optimized acidic property in obtained extruded sample was named as BZE.
combination with its well-developed porosity [41]. However, to the Steam treatment of the BZE sample was performed in a fix-bed re-
best of our knowledge there is no information regarding the interaction actor. Prior to steam treatment, the sample was activated at 550 °C
between zeolite and alumina binder in the H-[B]-ZSM-5 extruded under N2 flow (30 mL/min) for 2 h. The sample was hydrothermally

2
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

treated with 100% water vapor (WHSV = 3 h−1) at 500 °C for 12 h and 2.4. Catalytic performance tests
then cooled to room temperature. This sample was named as BZEH.
For comparison and to confirm the presence of Al2O3 component as MTP reaction was performed in a fixed-bed continuous-flow reactor
a constituent of alumina sol in the structure of BZE and BZEH catalysts, apparatus with a stainless steel tube reactor (450 mm length, 11 mm
firstly, pure alumina sol was dried overnight at 105 °C and then the I.D) under atmospheric pressure at 480 °C. The weight hourly space
solid specimen was calcined in a muffle furnace under air-flow at 540 °C velocity (WHSV) for methanol was 0.9 h−1 with a methanol/H2O mass
for 12 h at a heating rate of 3 °C/min. These samples were named as ratio of 1:1. Some detailed information on the performance tests is
ALS-Dry and ALS-Cal., respectively. described in the ESI. Fig. S1 provides a schematic view of the experi-
mental set-up consisting of a liquid/gas feeding section, a fixed-bed
reactor, and an analysis section. The reaction performance results, in-
2.3. Catalyst characterization cluding methanol conversion and hydrocarbon products selectivity
were calculated. Due to very fast equilibrium reaction of methanol to
Thermogravimetric analysis (TGA) was performed to detect the DME, those two species can be combined as one lumped reactant spe-
thermal decomposition of templates of as-synthesized (non-calcined) cies (oxygenates) in the calculation of conversion and hydrocarbon
samples using a TGA instrument (Pyris Diamond TG/DTA, products selectivity. Hence, the conversion of methanol in the MTP
PerkinElmer). TGA was performed in presence of nitrogen as inert reaction was calculated through the following equation:
purge gas with a flow rate of 200 mL/min and heating rate of 10 °C/ i o o
NMeOH − (NMeOH + 2NDME )
min, heating from 30 to 700 °C. Also, the amount of coke deposited over Methanol conversion = i
× 100
the spent catalysts was determined by TGA. To this end, the spent NMeOH (1)
samples with the initial weight of 5–6 mg were placed in the platinum The product selectivity was defined as the mole ratio of each pro-
pan and heated up to 1000 °C at a rate of 5 °C/min under zero-air duct (on CH2 basis) referred to the moles of converted methanol:
flowing (200 mL/min).
The powder X-ray diffraction (PXRD) pattern was used to recognize xNCox Hy
Selectivity = i o o
× 100
the type of crystalline phase and crystallinity of samples. PXRD patterns NMeOH − (NMeOH + 2NDME ) (2)
of these samples were collected by a Bruker d8 advance diffractometer
using Cu Kα radiation (λ = 1.5406 Å) in the 2θ range of 5°–80° at the where superscript i refers to the components at the inlet of reactor and
rate of 2°/min. The crystalline phases were determined using Joint superscript o refers to the components at the reactor outlet; subscript x
Committee on Powder Diffraction Standard (JCPDS) files. refers to the number of carbon atoms.
The morphology and particle size of the catalyst samples were ex-
amined using a field emission scanning electron microscope (FE-SEM 3. Results and discussion
Model: MIRA3 TESCAN, USA) operating at 15 kV. The elemental com-
position and distribution of each element were also determined by 3.1. Catalyst characterization
Energy dispersive X-ray (EDX) and SEM mapping analysis. All samples
were subsequently sputter-coated with a thin gold film to reduce 3.1.1. Catalyst structure
charging effects. TGA was used to determine the weight loss during heating of
The Fourier transform infrared (FT-IR) spectra of the samples were sample. Also, in this work, the TGA experiment was performed to verify
collected on a Bruker Vertex80 spectrometer with the resolution of the temperature of calcination process of the synthesized sample. The
4 cm−1 for addressing surface functional groups. TGA curve of the H-[B]-ZSM-5 powder as the dry sample (before cal-
The textural properties of the fresh and deactivated catalysts were cination) is illustrated in Fig. S2. The TGA thermogram showed two
measured using N2 adsorption-desorption technique at 77 K on a NOVA weight-losses in the temperature ranges 25–300 and 300–530 °C. The
2000 Quanta Chrome USA instrument. Before measurement, all the initial weight loss (I) of about 1.32% is due to loss of moisture which
samples were evacuated at 300 °C under N2 flow for 3 h. The conven- was physically adsorbed over zeolite surface. The main weight loss (II)
tional Brunauer-Emmet-Teller (BET) equation was used to calculate the of about 13.53% which took place around 450 °C could be attributed to
specific surface area (Stotal) from the adsorption data obtained (p/ removing and the oxidative decomposition of the organic template. The
po = 0.05–0.25). The total pore volume (Vtotal) was based on the ni- results showed that the point where rapid weight reduction occurred,
trogen adsorbed volume at p/po = 0.99. The micropore area (Smicro) was the temperature where the combustion of tetra-propyl ammonium
and micropore volume (Vmicro) were calculated from the t-plot curve at (TPA+) ions took place, releasing a lot of toxic gases (such as NOx, CO
thickness range between 3.5 and 5.4 Å. The mesopore size distribution and CO2). TPABr as a template tends to be decomposed at a tempera-
was calculated using the Barrett-Joyner-Halenda (BJH) method on the ture range of 350 °C–470 °C. As shown in Fig. S2, above 530 °C, a very
adsorption branch of the isotherm. The mesopore volume (Vmeso) is the small weight loss was detected. Consequently, the TGA thermogram
difference of the calculated total volume and the corresponding mi- showed that the selected operation conditions (530 °C, 12 h) for calci-
cropore volume. nation of the dried H-[B]-ZSM-5 sample were suitable because as-dried
The temperature-programmed desorption of ammonia (NH3-TPD) precursor did not demonstrate any weight loss above 530 °C. The results
for the prepared catalysts (fresh and reduced after reaction/coke de- of this work are compatible with the literature data [2,11]. To create a
composition) was carried out with a conventional flow apparatus stable thermal structure for the BZE catalyst, the raw BZE sample was
(BELCAT-A, BEL Japan, Inc.) to determine the surface acidity of various calcined at 540 °C. The weight loss of the ALS-Dry sample was 18.4 wt%
catalysts. A 35 mg of sample was loaded in the U- tube. Before ad- during the calcination process. This weight loss can be attributed to the
sorption, the samples were initially degassed under He flow at 300 °C decomposition of the boehmite according to the following reaction:
for 2 h at a heating rate of 10 °C/min. After cooling to 60 °C, the sample
2γ-AlOOH γ-Al2O3 + H2O (3)
was saturated with ammonia for 1 h. After saturation, the sample was
purged with helium for 30 min to remove the physically adsorbed The XRD patterns of the different samples are depicted in Fig. 1. The
ammonia on the surface of the catalyst. The temperature of the sample XRD patterns of the ALS-Dry and ALS-Cal. samples are depicted in
was then raised at a heating rate of 5 °C/min from 35 to 850 °C. The Fig. 1(a). As clearly seen in Fig. 1(a), ALS-Dry and ALS-Cal. samples
amount of ammonia desorbed from the catalyst was measured by display the typical diffraction peaks of boehmite (γ-AlOOH) and γ-
comparing the TPD areas with that for a standard sample using a Al2O3, respectively. The pattern of ALS-Cal. sample included peaks in
thermal conductivity detector (TCD). 2θ = 37.2°, 46.0° and 66.9°, which the intensity of the characteristics

3
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

zeolite framework, FT-IR analysis was performed on the different


samples. Since no extra crystalline phase related to B2O3 observed in
the XRD patterns of the prepared H-[B]-ZSM-5 zeolites. These ob-
servations confirmed the two issues; firstly, the incorporation of boron
modifier in the ZSM-5 framework structure has a uniform dispersion.
Secondly, since the amount of boron in various H-[B]-ZSM-5 samples is
very low, XRD analysis cannot determine the presence or absence of the
boron-related crystalline phases in these materials. The results of FT-IR
characterization confirm the incorporation of boron into the zeolite
structure. As shown in Fig. S3, IR bands at 670 and ~920 cm−1 at-
tributed to the symmetric bending vibration of Si–O–B and the
stretching vibration of Si–O–B band (tetrahedrally coordinated boron),
respectively. The weak peak at 1385 cm−1 is assigned to trigonally
coordinated framework boron. As clearly seen in Fig. S3, because of the
relative low boron amount, the intensity of the characteristics peaks are
very relatively weak. The relative crystallinity using IR analysis was
calculated from [(I550/I450)/0.72] × 100%, which I550 and I450 are the
intensities of the infrared bands around 550 and 450 cm−1, respectively
[48]. The IR crystallinity is summarized in Table 1. As can be seen in
Table 1, the relative crystallinity of the different zeolite catalysts is
high. Comparison of the relative crystallinity for the different samples
Fig. 1. XRD patterns of binder and synthesized zeolite samples (H-[B]-ZSM-5
calculated from PXRD and IR shows the values are almost the same.
zeolite powder (BZP), extruded (BZE) and steam-treated extruded (BZEH)).

3.1.2. Catalyst morphology


peaks are very relatively weak [40,47]. The typical peaks of ZSM-5 The FE-SEM images of BZP, BZE and BZEH samples are shown in
orthorhombic structure (JCPDS: 42–23) (2θ = 7.8, 8.7, 23.1, 23.8 and Fig. 2. As can be seen in the SEM micrographs, in all samples, ag-
24.3°) were detected in the patterns of all zeolite samples [40]. Fur- glomeration of the quasi-spherical H-[B]-ZSM-5 particles with the
thermore, the comparison between the XRD pattern of BZP sample with average size of around 1 μm is obvious. This kind of morphology has
that of the two other catalysts showed a slight decrease in the peak been reported in the literature [7,18]. Furthermore, the external surface
intensity for BZE and BZEH samples at 2θ = 22–25°, indicating that, the of BZP particles is quite clear and smooth which confirms the small
binding with Al2O3 and followed by steaming did not significantly amount of meso-cavities in the catalyst structure. On the other hand, in
change the structure of the H-[B]-ZSM-5 zeolite. The XRD spectra of the case of BZE, its surface is uneven because of the presence of binder
BZE and BZEH samples shows well-crystalline MFI characteristic and around and between the H-[B]-ZSM-5 particles like a blanket. However,
some characteristic peaks of γ-Al2O3 (2θ = 46.0°, 66.9°), suggesting after steam treatment, the surface of the BZEH became clean and the
that during the process of catalyst preparation the zeolite structure edges of its particles seemed to have melted and became very irregular,
keeps well as well as the boehmite is transformed into γ-Al2O3. How- causing higher degrees of agglomeration. The formation of such ag-
ever, the binder characteristic peaks in extruded samples almost dis- glomerates could happen through binding of the surface hydroxyl
appear in the baseline of the XRD spectra, due to their low intensity. groups (Si–OH) of neighboring crystallites, which can be facilitated at
Moreover, comparison between the XRD patterns of sample BZP with the high temperatures of the steam treatment, and in the presence of
catalysts BZE and BZEH showed that alumina prominent presence alumina binder [49]. The elemental composition and distribution of
caused an intensity decrease of the zeolite reflections in the patterns of each element were determined by Energy dispersive X-ray (EDX) and
the sample BZE and BZEH compared to BZP one. In order to determine SEM mapping analysis. The images of the elemental distribution of the
the influence of the combined extrusion-steam treatments on the degree prepared catalysts are depicted in Fig. S4, shows the distribution of Si
of relative crystallinity of H-[B]-ZSM-5 catalysts, calculations were and Al particles in the cross-section of the BZE and BZEH zeolites and
performed based on procedure A described in ASTM D5758-01. In this confirms the presence of agglomerated particles in localized areas. As
method, calculation includes a comparison of the integrated peak areas shown in the SEM images and SEM mapping, better distribution and
in the range of 2θ = 22° to 25° of different H-[B]-ZSM-5 catalysts re- less agglomeration of the particles was observed on the BZP catalyst.
lative to those of a highly crystalline sample (BZP sample). As can be The elemental compositions of the different catalysts synthesized in the
observed in Table 1, the relative crystallinity of the steam-treated cat- present study have been listed in Table 1. According to the literature
alyst (with 91% relative crystallinity) is partly lower than that of the [7], the extraction of aluminum is more favorable than silicon from the
un-treated sample. On the other hand, extrusion and the combined zeolite body in mild steaming post-treatment. As listed in Table 1, the
extrusion-steam treatments did not take down the crystallinity of the molar ratio of Si/Al slightly decreases after steaming treatment. Results
treated catalysts. showed the molar ratio of Si/Al of two samples BZE and BZEH is high
The FTIR spectra of all samples were recorded in the range of compared to the powder one.
400–4000 cm−1. Fig. S3 displays the infrared spectra in the region of
the silicate vibrations (400-1400 cm−1). As shown in Fig. S3, all of 3.1.3. Textural properties
catalysts displayed special adsorption bands of ZSM-5 structure near N2 adsorption/desorption isotherms and corresponding BJH pore
450, 550, 800, 1120, and 1225 cm−1. The bands near 450 and size distribution (PSD) curve of all samples are illustrated in Fig. 3 and
550 cm−1 are related to the Si–O vibration of SiO4 internal tetrahedral Fig. 4, respectively, and their textural properties are summarized in
units and the characteristic vibration of double five ring in MFI zeolite, Table 1. As listed in Table 1, the BET surface area and total pore volume
respectively. The peaks around 800 and 1120 cm−1 come from the of BZP sample are 375.3 m2/g and 0.212 cm3/g, respectively, while the
symmetric stretching of the external linkage and the internal asym- mesoporous surface area and pore volume are relatively low, 106 m2/g
metric stretching of Si–O-T linkage, respectively. The band at and 0.08 cm3/g, respectively. It is clearly shown that the surface area of
1225 cm−1 is attributed to the external asymmetric stretching vibration the treated catalysts does not change significantly. So that, decreasing
and verifies the presence of structures containing four chains of 5- of surface area is less than 2% after the combined extrusion-steaming
member rings [7,11]. To confirm the incorporation of boron into the treatment. This stability is attributed to the low Al content of the H-[B]-

4
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

Table 1
Textural properties of the fresh, deactivated and regenerated H-[B]-ZSM-5 zeolite powder (BZP), extruded (BZE) and steam-treated extruded (BZEH) samples.
Sample Elemental compositions (wt.%)a IR XRD Relative Crystallinity (%) Surface area (m2/g) Pore volume (cm3/g)
Crystallinity (%)
Si Al SBETb SMicroc SMesod VTotale VMicrof VMesog

BZP 46.47 0.69 100 100 375.3 269.3 106 0.212 0.132 0.080
BZE 24.22 5.49 95.8 96 386.6 173.9 212.7 0.247 0.073 0.174
BZEH 22.54 5.03 88.8 91 368.6 196.4 172.2 0.230 0.085 0.145
Deactivated BZP – – – – 250.5 185.6 64.9 0.137 0.085 0.052
Deactivated BZE – – – – 205.1 150.1 55.0 0.109 0.069 0.040
Deactivated BZEH – – – – 258.4 162.7 95.7 0.147 0.072 0.075
Reduced BZP – – – – 370.2 173.6 196.6 0.229 0.081 0.148
Reduced BZE – – – – 350.7 170.8 179.9 0.245 0.078 0.167
Reduced BZEH – – – – 368.2 162.7 205.5 0.251 0.069 0.182

a
Elemental compositions were determined by EDX analysis.
b
Total surface areas were obtained by the BET method using adsorption data in P/P0 ranging from 0.05 to 0.25.
c
Micropore surface area evaluated by t-plot method.
d
Mesopore surface area calculated using SBET-SMicro.
e
Total pore volume at P/P0 = 0.99.
f
Micropore volume calculated by t-plot method.
g
Mesopore volume calculated using VTotal-VMicro.

ZSM-5 sample (high Si/Al ratio) that makes this H-[B]-ZSM-5 resistant the steaming treatment at 500 °C for 12 h, the mesopore surface area
to hydrothermal deactivation. and volume of the extrudates decrease, while the micropore surface
As shown in Fig. 3, the powder H-[B]-ZSM-5 zeolite represents a area and volume increase. This is tentatively attributed to the collapse
combination of types I and IV isotherm, according to the IUPAC clas- and densification of some inter particle mesopores between the binder
sification, with a small hysteresis loop (type H4) at higher relative and zeolite upon the steaming treatment that led to the creation of new
pressures (P/Po = 0.5–0.95), implying the micropores are the major micropores in the hydrothermally treated extrudates [52]. In addition,
component of sample BZP. This hysteresis loop may have attributed to the researchers believe that during the steaming process, pores are
capillary filling and condensation of N2 in homogeneous slit-shaped blocked using extra framework aluminum atoms in the zeolite channels.
inter crystalline pores formed by the aggregation of nano-sized H-[B]- During the dealumination process, the framework of the zeolite was
ZSM-5 crystals [50]. On the other hand, although, BZP catalyst has damaged which it's severity depends on the operation conditions of this
more microporosity, but, when its crystals get close to each other the process. As can be seen in Table 1, the relative crystallinity of the BZEH
intercrystaline voids can act as a pore, which leads to condensation and sample after steaming treatment has been decreased compared to BZE
gives such type of hysteresis at higher relative pressure (very small catalyst. These results confirm significantly framework damage under
compared to the extrusion-steam treated samples). Besides, the shape of the selected steaming conditions, because of that the production of
nitrogen adsorption-desorption isotherm of BZP sample at relative extra framework aluminum atoms is high which those remain in the
pressures lower than P/Po = 0.2 seems very rare, verifying its micro- zeolite channels and block the pores, which reduces the amounts of
porous structure without considerable mesoporosity. This point can be surface area and pore volume [40]. These observations are agreed well
further verified by the BJH pore size distribution curve of BZP sample with the PXRD and FE-SEM results.
(Fig. 4), where there is a tiny peak around 2.5 nm, indicating a few
mesopores in the powder H-[B]-ZSM-5 zeolite.
However, it can be observed that after the addition of alumina 3.1.4. Acidity
binder, the N2 adsorption/desorption isotherms of the extruded sam- The acidity of catalysts plays a vital role in determining the catalytic
ples, no matter before and after steaming, are transformed to type-IV, performance in the MTP reaction, since, zeolites have been recognized
with a pronounced hysteresis loop at a relative pressure higher than 0.5 as solid-acid catalysts. Hence, the acidic properties of H-[B]-ZSM-5
(Fig. 3), indicating the existence of both micropores and mesopores in powder and extruded samples were investigated by NH3-TPD tech-
the extruded samples. Also, the isotherms show a sub-step in the re- nique, and their NH3-TPD profiles are given in Fig. 5. Moreover, the
lative pressure around P/Po = 0.2–0.4 which is attributed to fluid-to- acidities of all catalysts determined from TPD peak areas as well as the
crystalline like phase transition of the absorbed nitrogen in the micro- strength of acid sites corresponding to desorption peaks temperatures
porous structure [51]. This difference could be attributed to the orga- are tabulated in Table 2.
nization of zeolite and binder particles upon agglomeration by extru- As can be seen in Fig. 5, two evident desorption peaks are observed
sion, which creates an additional inter particle mesoporosity [36,40]. in the NH3-TPD profiles of the catalysts: a low-temperature peak (LTd1)
This was further verified by the corresponding pore size distributions at around 180–195 °C and a high temperature peak (HTd2) at around
(PSDs) calculated by BJH model based on adsorption curves. As shown 370–390 °C assigning the weak and the strong acid sites, respectively,
in Fig. 4, both extruded and steam treated extruded catalysts reveal a which are the characteristic peaks of zeolite with MFI structure. Al-
bimodal pore size distribution, with a narrow peak centered at ~2.5 nm though NH3-TPD analysis cannot distinguish between Brønsted and
and a broadened distribution centered at ~9 nm, respectively. How- Lewis acidity, the low temperature peak could be assigned to ammonia
ever, after steam treatment the pore size distribution of BZEH sample desorption from non-framework Lewis acid sites and weaker Brønsted
became narrower. Indeed, compared to powder catalyst (BZP), both acid sites, whereas the high temperature peak could be attributed to
extruded samples (BZE and BZEH), demonstrated a smaller value in desorption of ammonia from stronger Brønsted and Lewis acid sites
micropore surface area and micropore volume, due to dilution of the H- [11,13,17]. As seen in NH3-TPD curves, there is a significant third peak
[B]-ZSM-5 crystal and blocking some of its well-developed micropore in the range of 530–870 °C for the treated catalysts, while the higher
structure during extrusion [35,40], whereas, opposite changes were temperature peak above 530 °C is also shown in the NH3-TPD profile of
observed for external area and mesopore volume due to binder addi- BZP sample but less intense. This can be explained by the fact that both
tion. Moreover, comparison of BZE and BZEH samples reveals that upon BZE and BZEH catalysts have been prepared as extruded form using
alumina sol as a binder. In this case, some aluminum species that

5
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

Fig. 2. FE-SEM images of H-[B]-ZSM-5 zeolite powder (BZP), extrudes (BZE) and steam-treated extruded (BZEH) samples.

belong to alumina sol strongly adsorb NH3, which is probably due to 345 °C and 615 °C, respectively [53]. This shows that the presence of
NH3 adsorbed on aluminum atoms that only bind to one Al [53,54]. As the alumina binder in the extruded zeolites (BZE and BZEH) induced
clearly seen in Fig. 5, the area of the third peak of samples BZE and additional acidity. The total acidity of the extruded samples is ap-
BZEH is greater than catalyst BZP. According to experimental results, proximately 20% higher than that of sample BZP (quantity of binder
the acidity amount of the third peak is increased in the order of used in the extrusion process). Also, the maximum temperature of the
BZEH > BZE ≫ BZP. The acidity profile of ALS-Cal. sample revealed third peak shift to the higher temperature at the different extruded
the existence of two mainly moderate and strong acid sites, located at specimens (698 °C and 724 °C for samples BZE and BZEH, respectively).

6
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

Fig. 3. N2 adsorption/desorption isotherms of the fresh H-[B]-ZSM-5 zeolite powder (BZP), extruded (BZE) and steam-treated extruded (BZEH) samples.

Fig. 4. BJH pore size distribution curves of the fresh H-[B]-ZSM-5 zeolite
powder (BZP), extruded (BZE) and steam-treated extruded (BZEH) samples.
Fig. 5. NH3-TPD profiles of the fresh H-[B]-ZSM-5 zeolite powder, extruded and
The large asymmetric peak at the ammonia-TPD profiles of samples, dealuminated extruded samples.
BZE and BZEH can be assigned to desorption from strong Lewis acid
sites, while the small peak at the ammonia-TPD profile of sample, BZP than Al–OH–Si. It is extensively believed that weak acid sites have little
at the higher temperature (729 °C) is assigned to desorption from strong activity in methanol conversion to light olefins, whereas strong acid
Brønsted acid sites [54]. The strong Lewis acidity of the extrudates sites on the H-ZSM-5 catalyst surface are indispensable for the hydro-
slightly increased after the steaming process, which is related to the carbon production [11,17]. Indeed, the first step of the MTP reaction
aluminum extraction from the framework of the zeolite [54]. This ob- (methanol to DME conversion), alkylation and methylation reactions
servation suggests the effective interaction of the binder with the fra- were performed by weak acid sites. Also, these sites can prevent many
mework aluminum of the ZSM-5 catalyst. These Brønsted and Lewis side reactions such as hydrogen transfer reactions in MTP process. But,
acid sites were related to framework Al atoms and to extra-framework the strong acid sites are known as main active acid sites for olefin
Al atoms, respectively. production in MTH reaction. Therefore, in order to improve catalyst
According to literature, for boron-containing ZSM-5, B incorporated stability and propylene selectivity in MTP reaction, it is necessary to
into the framework positions (Si–OH–B) and silanol groups (Si–OH) at adjust and control the values of weak and strong acid sites [3,11,18].
the external surface or at lattice defects are responsible for generation As shown in Fig. 5 and Table 2, the powder H-[B]-ZSM-5 catalyst
of weak Brønsted acidity, whereas Al present in the zeolite framework (BZP sample) has a moderate amount of weak (0.142 mmol NH3/g) and
(Al–OH–Si) gives rise to strong Brønsted acidity [55,56]. Based on NH3- strong (0.106 mmol NH3/g) acid sites. Also, the NH3-TPD test results
TPD and 1H MAS NMR studies, Scholle et al. [57] additionally con- clearly indicate that not only did the agglomeration of powder zeolite
cluded that the Si–OH–B sites are more acidic than Si–OH but weaker with alumina sol binder slightly decrease the concentration of strong

7
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

Table 2 As noted earlier, even though the NH3-TPD results show a decrease
NH3-TPD data of the fresh and regenerated H-[B]-ZSM-5 zeolite powder, ex- in the acidity of the steam treated catalyst, BZEH has higher acid
truded and steam-treated extruded samples. strength than the BZE sample. This is mainly owing to the formation of
Sample Distribution and concentration of acid sites Peak temperature (°C) new Al–OH–Si groups (i.e. strong Brønsted acid sites) upon hydro-
(mmol NH3/g) thermal treatment during steaming process, when Al-containing spices,
transferred from alumina sol binder, interact with the terminal Si–OH
Region I Region II Total Strong/ LTP1 HTP2
groups (weak Brønsted acid site) at the zeolite/binder interface.
Weak Strong Weak
Additionally, the framework defects resulted from hydrolysis of the
BZP 0.142 0.106 0.248 0.75 186 387 framework B are compensated for by Al migrated from the binder under
BZE 0.154 0.099 0.253 0.64 184 364 steaming conditions since the loss of crystallinity of BZEH is not sig-
BZEH 0.151 0.084 0.235 0.56 189 369
nificant, its strong acidity is reduced, and there is an increase in its
Reduced 0.049 0.059 0.108 1.21 195 384
BZP strong acid strength. In other words, the isomorphous substitution of Al
Reduced 0.102 0.045 0.147 0.44 197 360 for B under hydrothermal conditions resulted in the formation of new
BZE strong acid sites. According to literature, the Brønsted acid strength is
Reduced 0.120 0.065 0.185 0.54 192 346 increased in the order of Al ≫ B > Si [46,56].
BZEH

3.2. Catalytic performance


acid sites, but also it increased the concentration of weak acid sites of
the extruded catalyst (BZE sample), so the strong/weak ratio decreased The catalytic activities of all H-[B]-ZSM-5 powder and extruded
from 0.75 to 0.64. Furthermore, the desorption maximums of both catalysts were evaluated in the conversion of methanol to propylene
strong and weak acidities in TPD profile were slightly shifted to lower (MTP) reaction in a continuous flow fixed bed reactor under the similar
temperatures due to the addition of alumina sol binder, indicating that operation conditions (atmospheric pressure and 480 °C) using feed
extruded catalyst has the lower acidic strength than the powder catalyst containing a mixture of 50 wt% methanol in water with methanol
(see Fig. 5 and Table 2). It should be noted that Al in zeolite framework WHSV of 0.9 h−1. In this reaction, ethylene, propylene and butylenes
causes strong acid sites. So, the decrease in the density of strong acid were found as main products while dimethyl ether, methane, C1–C4
sites of extruded BZE sample could be firstly attributed to the adequate paraffinic hydrocarbons and C5+ were also accompanied as by-pro-
dilution of the strong acid sites of H-[B]-ZSM-5 with binder [35], and ducts. All H-[B]-ZSM-5 catalysts performed catalytic conversion to light
secondly related to the partial blockage of zeolite channels by the olefins but they showed different properties in product selectivity and
binder [58], which was again confirmed by its textural properties listed lifetime. The methanol conversion and DME formation as a function of
in Table 1. Thirdly, the migration of some soluble Al species from the time on stream (TOS) over the various H-[B]-ZSM-5 catalysts are shown
binder and subsequent insertion into the framework of the zeolite in Fig. 6.
during calcination leads to a partial neutralization of Brønsted sites by As can be seen in Fig. 6, with methanol/water feeding on the three
aluminum cations and consequently causes a decrease in the strength type of catalysts, the reaction proceeds from the initial stage to the
and amount of strong acid sites [42,45,59]. Lastly, the decrease in the steady stage and then catalyst deactivation. The H-[B]-ZSM-5 catalysts
Brønsted acidity of the extruded zeolite may also be due to deal- with white color turned to gray during the reaction and finally to black
umination of the zeolite during its calcination at higher temperature color when they were deactivated. At the initial stage of the reaction,
because of the water vapor produced by the dehydration of the binders. due to the availability of the majority of acid sites and methanol's small
The decrease in the acidic strength of the extruded sample mainly is due size, all catalysts exhibited nearly full methanol conversion, indicating
to the preferential neutralization of the most acidic bridging hydroxyl the high initial activity of all samples [17,50]. On the other hand, it was
groups in H-[B]-ZSM-5 when ion-exchanged by alumina species from well known that methanol dehydration formed DME first and then
the binder [52,60]. Moreover, early work investigating the stability of converted to light olefins. Thus, product olefins were formed together
boron-substituted H-ZSM-5 showed that the insertion of Al could be
facilitated by the presence of a readily hydrolysable heteroatom in the
framework of B-[H]-ZSM-5 zeolite [46,55].
Besides the decrease in the concentration of Brønsted acid sites,
additional weak and strong acid sites can be imported on the external
surface of zeolite by the alumina-sol binder which is mostly due to the
substitution of framework Si for Al from the binder at high temperature
during calcinations [42,45,61]. This may be the reason why both the
weak and the total acid sites densities of extruded sample increase to
some extent.
However, a comparison of BZE and BZEH samples reveals that after
steaming process, the densities of both the weak and the strong acid
sites of steam treated extruded catalyst decreased, whereas the
strengths of the aforementioned sites increased (see Fig. 5 and Table 2).
The dehydroxylation and agglomeration of the extra-framework Al
species (weak Lewis) and alumina binder, as well as the removal of
intercrystalline Brønsted acid (formed during the extrusion process),
along with the preferential removal of more labile Al located at lower T-
O-T angle due to dealumination of ZSM-5 framework, are the three
factors contributing to the decrease in the amount of weak acid sites of
the extruded zeolite after the steam treatment [52]. In contrast, the Fig. 6. Methanol conversion and DME selectivity as a function of time on
stream over the H-[B]-ZSM-5 powder, extruded and steam-treated extruded
decrease in the strong acid sites of hydrothermally extruded sample is
samples. (Reaction conditions: T = 480 °C, P = 1 atm, WHSV = 0.9 h-1, feed:
expected to be the result of the removal of framework aluminum, when
50 wt% methanol in water, compositions are averages over useful catalyst
the Al–O–Si bonds are attacked by steam during the treatment [61]. lifetime.)

8
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

with residual DME which could not convert to products. When the re- contains light olefins, which is consisted of ethylene, propylene, and
actions were launched, H-[B]-ZSM-5 showed approximately 100% total butylene (C=2 − C=4 ). The last group contains heavier olefins,
methanol conversion without any value of DME. Nevertheless, with the starting from pentene, as well as higher paraffins and aromatics such as
increasing time on stream, all the tested catalysts were gradually lost benzene, toluene, and xylene (BTX), which are denoted by C5+. At the
their catalytic activity (the methanol conversion started to drop to- beginning of MTP reaction, the powder catalyst delivers the lowest
gether with increasing DME formation) with different deactivation propylene selectivity (38.64%) but the highest ethylene selectivity
rates, which is generally attributed to the different coverage of acid (19.52%) which results in the lowest propylene/ethylene (P/E) ratio
sites and blockage of the pore mouth by carbon deposit [1,10], which is (1.97). Moreover, it also displays an especially high C1–C4 saturated
further discussed in the following sections. In literature, it has been hydrocarbons selectivity (8.44%) which may be related to poor diffu-
reported that deactivation appeared with DME formation. To obtain a sion rate that leads to quick coke deposition [25]. In comparison with
quantitative estimate of the catalytic stability, the ‘‘catalytic lifetime’’ is the powder catalyst, both extruded samples demonstrate noticeably
defined as the running time at which the conversion rate of oxygenates higher selectivity to propylene as well as butylenes and C=2 − C=4 olefins,
(methanol and dimethyl ether) exceeds 90% (i.e., t90%), which is while their selectivity toward ethylene and C1–C4 paraffinic hydro-
marked with a dashed line in Fig. 6. From an industrial point of view, a carbons is relatively lower. However, after steaming treatment, the
conversion less than 90% is unacceptable for the MTP process and the obtained BZEH catalyst showed better catalytic performance than BZE
catalyst should be regenerated or substituted [28]. catalyst, where the ethylene selectivity of BZEH (13.36%) is somewhat
It is well believed that catalyst deactivation in MTP reaction is lower than that of BZE (15.35%), although there is almost no significant
generally attributed to the coke deposition on the catalyst surface, changes in propylene selectivity (40.31% vs. 41.56%). Hence, BZEH
which is caused by the side reactions that also take place on the acid represents the highest initial P/E ratio (3.11) among the three catalysts
sites. In any fixed bed catalyst operation, the portion of the catalyst bed (Table 3).
that is in the downstream of the reaction front is deactivated by coking. According to the industrial perspective, the MTP reaction was per-
As the reaction progresses with increased time, the deactivation front formed over various zeolite catalysts to reach the point of deactivation.
moves through the catalyst bed, until most of the catalyst bed is de- Also, the output of the products during the reaction time is very im-
activated and the conversion drops rapidly after breakthrough of me- portant in industrial units. Fig. S5 and Table 4 show the detailed pro-
thanol (and DME) [5]. As clearly seen in Fig. 6, the deactivation be- duct distribution in the conversion of methanol to light olefins over
havior of extruded samples is rather different from that of the powder three different H-[B]-ZSM-5 catalysts. It should be noted that the pro-
H-[B]-ZSM-5 sample. In the case of BZP sample, the conversion of duct distribution represents average values over the time when me-
methanol slowly declines to 90% after 580 h on stream and, thereafter, thanol conversion is above 90% (in useful lifetime of catalyst). As
deactivation rate is relatively fast. Whereas, methanol conversion over shown in Fig. S5 and Table 4, all the three catalysts exhibited almost
extruded catalysts before and after steaming drops below 90% after complete methanol conversion, but their product selectivity somewhat
520 h and 750 h on stream, respectively. Furthermore, compared to the differed. The propylene selectivity is almost similar (44–45%), while
powder catalyst, a slower drop in methanol conversion is observed for the selectivity to ethylene (7–12%) and butane (20–24%) is different.
extruded samples (BZE and BZEH samples) after catalyst breakthrough. The steam treated extruded catalyst (BZEH sample) exhibited the
However, the untreated-extruded sample exhibits shorter lifetime than lowest and highest selectivity to C=2 (7.78%) and C=4 (23.66%), respec-
that of the powder H-[B]-ZSM-5 catalyst in MTP reaction which is tively. In an MTP process, all alkenes except propylene, especially
unfavorable for the current industrial practice using multi-stage fixed- C4~C6, are recycled back to the reactor inlet for a dominant propylene
bed reactor. So, the lifetime of the catalysts increases in the following production [62]. It is verified that the reactivity of C=4 species is sig-
order: BZE (520 h) < BZP (580 h) < BZEH (750 h). Specifically, the nificantly high. These species react with ethylene (metathesis reaction)
treated-extruded sample exhibits the longest lifetime among these three and produce propylene. Based on this theory, the BZEH catalyst is more
types of catalysts, indicating that it may be a remarkably effective suitable for the MTP process than other samples. Furthermore, the
catalyst for the industrial MTP reaction. It is interesting to note that a steam treated extruded catalyst showed a further increase in the pro-
similar improvement in the catalyst lifetime was previously reported by pylene to ethylene ratio (P/E, 6.41 vs. 4.43 for the powder H-[B]-ZSM-5
Zhang et al. using a series of hydrothermally-treated extruded H-[B]- catalyst). The C5+ hydrocarbons, methane and alkanes are known as
ZSM-5 zeolites during the methanol-to-propylene reaction [40]. by-products which are produced from the side reactions and finally
Another state indicates that the beginning of deactivation of H-[B]- converted to coke. So, when the selectivity to these hydrocarbons is
ZSM-5 catalysts in the MTP process can be shown by the DME appeared high, the coke formation and catalyst deactivation will be accelerated.
in the reactor outlet products. As can be seen clearly in Fig. 6, at the According to this, the selectivity of methane (1.09 vs. 2.43%) and C1–C4
initial stage of MTP reaction, it was found that DME formation is ap- paraffinic hydrocarbons (4.37 vs. 5.94%) is the lowest over the BZEH
proximately zero. That means the catalyst has the highest activity to sample, whereas the selectivity of C5+ hydrocarbons (18.93 vs.
undergo methanol to DME reaction and DME to olefins reaction. After a 16.20%) is the highest (see Fig. S5 and Table 4). Although the amount
while, partial deactivation of the catalyst begins when a small amount of the produced C5+ hydrocarbons over BZEH sample is more than the
of DME in the product stream is produced. Subsequently, DME pro- two other catalysts, the results showed that the performance and sta-
duction began to increase at a considerable rate. As shown in Fig. 6, the bility of this catalyst are better and more appropriate (see Fig. 6 and
rate of increase in DME production over powder H-[B]-ZSM-5 catalyst Fig. S5 and Tables 3–4).The amount of stronger acid sites of BZP and
(BZP) is much faster than that over the two treated samples (BZE and BZE samples is higher than those of BZEH catalyst (see Table 2), it can
BZEH). Thus, the BZP catalyst was deactivated quickly. Fig. 6 shows be observed that high amount of coke was deposited on strong acid sites
that among the three different catalysts, BZEH catalyst has the lowest of BZP and BZE samples. This was further verified by the determination
rate of deactivation. of coke content calculated and measured using TGA analysis. Thus, the
Table 3 displays full descriptions of initial products distributions in strong acid sites of BZP and BZE samples, that the DME to olefins re-
the conversion of methanol-to-propylene over the H-[B]-ZSM-5 powder, action is occurred by them, were blocked by the produced coke and the
extruded and steam treated extruded catalysts when the catalysts catalyst was deactivated quickly.
yielded nearly full conversion and were presumably free from coke One of the most important points regarding fixed reactor-based MTP
(after 24 h TOS). At this case, the methanol conversion is approximately process, like Lurgi's technology, is the proper selection of high-perfor-
100% and the catalyst is fresh. The obtained products are classified into mance catalysts with long lifetime and products with the steady dis-
three groups. The first group contains lower paraffins, including me- tribution. So, the changes in product selectivity with time on stream
thane, ethane, propane, and butanes (C1–C4). The second group (TOS) over all samples are investigated in detail. As clearly seen in

9
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

Table 3
Initial products distributions of MTP reactiona over the H-[B]-ZSM-5 powder, extruded and steam-treated extruded samples.
Sample Conversion (%) Selectivity (C-wt.%) P/E ratio

C1-4 C=
2 C=
3 Total C=
4 C=
2 − C=
4 C5+

BZP 99.89 8.44 19.52 38.64 18.76 76.92 14.64 1.97


BZE 99.79 6.49 15.35 40.32 21.82 77.49 16.02 2.63
BZEH 98.55 6.51 13.36 41.55 23.27 78.19 15.29 3.11

a
Reaction conditions: T = 480 °C, P = 1 atm, WHSV = 0.9 h−1, feed: 50 wt% methanol in water, the results obtained after 24 h of reaction.

Fig. 7, all catalysts exhibit regular catalytic performance with TOS for large molecules like heavier olefins and aromatics, from the interior
target product distribution, including propylene and butylenes nearly center to the surface mouth of the channel in the zeolite crystal. These
steady formation, a gradual decrease of ethylene and a gradual increase coke precursors heavily deposit inside the micropores of zeolite and its
of C5+ hydrocarbons [39,40]. external crystal surface, causing complete or partial blockage of some of
With increasing reaction time, propylene selectivity over BZP the channel openings and thereby shortening the catalyst lifetime. On
sample gradually increases from an initial value of 38.64% to a max- the contrary, BZEH maintains a stable formation of propylene for a
imum of 51.47% and then rapidly decreases. Butylene shows a similar longer period of time than BZE and BZP catalysts, which is mainly at-
tendency to propylene in a rather alleviated extent. In contrast, ethy- tributed to the generation of the new active site during steaming
lene exhibits a completely opposite behavior along with the prolonged treatment [46,56]. Besides, although BZE and BZEH catalysts exhibit
reaction time on stream, decreasing slowly from 19.52% at the begin- very similar propylene selectivity at the steady stage, a further con-
ning of the test to 6.02% after 400 h TOS and then remains almost tinues decrease in ethylene selectivity is observed for BZEH. So it shows
unchanged even within the time period of catalyst deactivation a higher continuous increase in P/E ratio during the MTP reaction, as
(Fig. 7(a)). C1–C4 paraffinic hydrocarbons also exhibits a decreasing shown in Fig. 7(h).
selectivity from 8.44% to 4.67% after 570 h TOS, but close to the de- The prominent difference in the catalytic performance of powder
activation stage their formation increases greatly to 11.88%. Further- and hydrothermally treated extruded catalysts can be explained by the
more, the formation of C5+ hydrocarbons exhibits obvious increase difference in the level of porosity and acidity as well as the strength of
from 14.64% at the initial to 42.79% at the end. According to “dual the acid sites and accessibility of reactants to these active sites [11,17].
cycle” mechanism, aromatic-based cycle and olefin-based cycle are fa- For the solid-acid catalytic reaction, the acidity of the catalyst is one of
vorable to form ethylene and propylene, respectively [17,63]. It was the most important factors affecting the reactivity in the MTP reaction.
also reported that the generation of ethylene needs stronger acid sites It is accepted that the strong acidity is regarded as the dominant active
compared to propylene and butylene. At the initial stage, the high sites for the MTP reaction, but they also can promote the formation of
amount of strong acid sites makes the aromatics-based cycle propaga- coke species through the secondary reactions, which deactivates the
tion preferred, so, the ethylene, C1–C4 and C5+ species are the dominate catalysts. The weak acidity, in addition to the fact that the catalyst
products. But as the reaction proceeds, the active acid sites of catalyst participates in the conversion of methanol to dimethyl ether, can also
are gradually reduced, because the coke deposit could partially cover catalyze the alkylation and methylation reactions, which highly influ-
the acid sites of zeolite and block its pore channels. Based on the pre- ences propylene selectivity according to the dual-cycle concept. More-
vious literature, at the catalyst deactivation stage, the propylene se- over, the weak acid sites efficiently hinder various hydrogen-transfer
lectivity rapidly decreases but the formation of C5+ species grows fast reactions producing saturated hydrocarbons and aromatics (C5+),
due to a significant decrease in the number of acid sites by coke cov- which are the precursors for coke formation [17,22,63,64]. As a result,
erage or pore mouth blockage [9,17,20,64]. in comparison to the powder catalyst, BZEH sample with a higher
In contrast to the powder catalyst, extruded samples do not show amount of weak acid sites and a lower amount of strong acid sites re-
abrupt changes in the distribution of the hydrocarbons with time on presents an appropriate strong/week acid ratio (0.56) for the propa-
stream and the selectivity to propylene is moderately increasing in gation of the olefin-based cycle in the MTP reaction, leading to high
time. Indeed, the selectivity to ethylene is slightly decreasing. However, propylene selectivity for a longer period of time. The synergetic effect
despite these significant changes in the products distribution, the MTP of extruding and steaming of H-[B]-ZSM-5 zeolite also leads to weak-
lifetime of BZE sample is lower than that of BZP catalyst, which is ening of the acid sites strength for BZEH sample, which is beneficial to
predominately attributed to decrease in the strength and amount of alleviate the secondary reactions (alkylation, cracking, hydrogen
strong acid sites of powder zeolite during the extrusion process as well transfer and cyclization) and suppress the aromatic-based cycle. Con-
as its lower ratio of strong/weak acid sites, as discussed in section 3.1.3. sequently, the coke precursor condensation steps are attenuated, and
Also, the deterioration of micropores caused by alumina sol binder is therefore deactivation is minimized [11,23,40].
another notable factor. As shown in Table 1, BZE sample exhibits the Another possible explanation of substantial change in the catalytic
least micropore volume compared to other catalysts, as a result of the behavior of BZEH in comparison with BZP can be the presence of sig-
blockage of the micropore mouth. The partial blockage of micropores in nificant mesoporosity in this hydrothermally extruded sample. In case
this alumina-bound catalyst can significantly reduce the diffusivity of of BZEH sample, coke is formed exclusively at the external surface and/

Table 4
Average products distributions of MTP reactiona over the H-[B]-ZSM-5 powder, extruded and steam-treated extruded samples.
Sample Conversion (%) Selectivity (C-wt.%) P/E ratio

C1-4 C=
2 C=
3 Total C=
4 C= =
2 − C4 C5+

BZP 99.14 5.94 11.52 45.05 20.93 77.51 16.2 4.43


BZE 98.57 5.26 9.66 44.22 22.59 76.48 18.09 5.18
BZEH 98.55 4.37 7.78 44.96 23.66 76.39 18.93 6.41

a
Reaction conditions: T = 480 °C, P = 1 atm, WHSV = 0.9 h−1, feed: 50 wt% methanol in water, compositions are averages over useful catalyst lifetime.

10
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

Fig. 7. Product selectivity as a function of time on stream over the H-[B]-ZSM-5 powder, extruded and steam-treated extruded samples. (Reaction conditions:
T = 480 °C, P = 1 atm, WHSV = 0.9 h-1, feed: 50 wt% methanol in water, compositions are averages over useful catalyst lifetime.)

11
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

or mesopores, while for the solely microporous powder zeolite (BZP), resulting in higher deposition of catalytic coke on the zeolite, BZE, and
the coke is more heavily deposited inside the micropores [50]. To in turn, fast deactivation of the catalyst.
support these speculations, the deactivated and regenerated catalysts Apart from pore size, zeolite acidity is also influential on the amount
were evaluated using TGA/DTG, BET and NH3-TPD methods. of coke formation. Since catalytic coke is formed over zeolite acid sites,
its yield is dependent on strength distribution and density of acid sites.
Catalytic coke content of zeolite is expected to be increased by an in-
3.3. Characterization of the deactivated and regenerated catalysts
crease in strength and number of acid sites. According to the NH3-TPD
data (Table 2) and profile (Fig. 5), the sample BZE has relatively more
In order to investigate more the catalysts deactivations during the
strong active acid sites compared to others (BZP and BZEH), conse-
MTP reaction, the deactivated catalysts were characterized with respect
quently, analysis of the reaction kinetics indicates that an increased
to their surface areas and coke contents. It is known that the catalyst of
Brønsted acid site density facilitates the formation of coke species and
the MTP reaction is regenerated periodically in industrial units. As on-
enhances their formation rate. Researchers have confirmed that with
stream time increases, the activity of MTP catalyst becomes lower due
increasing Brønsted acid site density, the formation of heavy hydro-
to partial blocking of the active sites of the catalyst by coke. If the
carbons (Table 3) and subsequently the amount of coke deposition in-
conversion of methanol is less than 90%, the catalyst has to be re-
creases.
generated with a mixture of air and nitrogen to burn off the coke. In this
As noted in section 3.1.2, H-[B]-ZSM-5 zeolites have both micro and
work, the deactivated catalysts were regenerated in flowing of active
meso cavities in their structure. Hence, deactivation of these catalysts
gas (mixture of air and N2 including different O2 content form 1% up to
can happen by coke formation/deposition inside the micropore (in-
21%) in a fixed-bed reactor at 480 °C for 60 h. After that, the re-
ternal coke) and/or on the mesopore surface (external coke) that leads
generated catalysts were characterized using N2 adsorption-desorption
to progressive pore blockage, coverage and subsequent loss of active
and NH3-TPD methods.
sites. The coke deposition rate and amount of coke are variables de-
pending on the physico-chemical characteristics of the catalyst and the
3.3.1. TGA analysis operating conditions of the MTP process. The coke distributions as in-
The amounts of coke deposited on the deactivated BZP, BZE and ternal and external cokes were calculated using the results of TGA and
BZEH catalysts were evaluated by TGA measurements, and the average N2 adsorption/desorption analysis. The details of the calculations have
coke deposition rates were obtained by correlation analysis with the been reported in the literature [65–67]. MTP reactions mostly occur in
lifetime. The coking rate is an important evaluation for the catalytic micropore channels of ZSM-5 zeolite where the majority of active sites
performance [16]. TGA/DTG profiles of the deactivated BZP, BZE, and are situated, the internal coke can lead to direct coverage of the active
BZEH catalysts are illustrated in Fig. S6(a). Also, the coke content and sites and obstruction of diffusion thus imposing a detrimental effect on
the average coking rate of these deactivated catalysts are summarized the durability of the catalyst. Whereas, the formation of coke on the
in Table 5. Approximately 16%, 22%, and 14% weight loss of the BZP, external surface (external coke) of the catalyst may lead to plugging of
BZE and BZEH deactivated catalysts, respectively, between 400 °C and pore mouth if the coke is nonporous and impermissible for gas mole-
700 °C was observed, which was attributed to the coke burning of the cules involved. This would avoid the reactant from accessing the active
deactivated H-[B]-ZSM-5. These weight losses can be observed with a sites, adversely affecting the catalyst performance. Therefore, it infers
DTG exothermic peak, which was attributed to the combustion of the that the rate of formation of internal coke and structural properties of
coke deposited on the H-[B]-ZSM-5 catalysts surface. As can be seen in external coke hold the key to explaining the different deactivation be-
Table 5, BZP contains 15.86 wt% coke deposit after 618 h on stream, havior of the two catalysts. As shown in Table 5 and Fig. S6(b), (c), the
while for BZEH the coke content is 14.37 wt% after 774 h TOS. The amount of coke deposited on the external surface and its corresponding
average coke deposition rate (Rcoke) was calculated using the method in coke deposition rate was estimated to be about 74, 94 and 98% of total
Ref. [65]. Rcoke on BZEH sample is 0.367 mg/gcat.h which is lower than coke of BZP, BZE and BZEH samples, respectively, while the coke
0.512 mg/gcat.h of BZP and 0.716 mg/gcat.h of BZE catalysts. The best formed inside the micropore was estimated to be about 26, 6 and 2% of
catalytic performance of BZEH catalyst is mainly due to its large ex- total coke, respectively. The rate of internal coke formation determines
ternal surface area and low strong/weak acid sites ratio, which are the catalyst resistance to deactivation. As the results show, in the case
helpful for attenuating coke deposition in the micropore channels even of the treated samples especially BZEH, coke strongly preferred to form
at a high coking level. Consequently, the slight coke deposition and the on the external surface, while during MTP reaction over BZP catalyst,
lesser susceptibility to form more coke deposits can contribute to the coke is deposited both on the external surface and inside the micropore.
slowing down of the deactivation of steam treated extruded catalyst On the other hand, the treated catalysts contain higher external coke
(BZEH) that ensures the long-term stable production of propylene from deposition rate and lower internal coke content than the untreated
methanol [40,41,66]. sample (see Fig. S6(b), (c)). These results are evidence for increasing
The acidity and the pore structure of zeolites are the two significant mass transfer in both BZE and BZEH samples, which is consistent with
catalyst properties which greatly affect the amount of coke formation. the previous reports [65–67]. Results showed that BZP zeolite rapidly
Coke formation over the catalyst BZE was remarkably higher than that deactivates due to high internal coke formation, while BZEH catalyst
over BZP and BZEH. The zeolite BZE which contained a larger pore size deactivates slowly due to the absence of significant internal coke.
(Table 1) produced higher amount of coke compared to BZP and BZEH. Steaming treatment under mild operation conditions (500 °C, 12 h)
As clearly shown in Table 1, the high mesopore volume of the catalyst, causes a relatively slight defect in the zeolite framework and partially
BZE leads to creating a high degree of polymerization reactions and extracts framework Al which is attributed to stronger Brønsted active
formation of polyaromatic compounds which act as coke precursors,

Table 5
Coke content and average coking rate over the H-[B]-ZSM-5 powder, extruded and steam-treated extruded samples.
Sample Time on stream Total weight loss H2O content Coke content Coke on external surface Coke inside micropore Average coke deposition rate
(h) (%) (%) (%) (gcoke/gcat.) (gcoke/gcat.) (mg.g−1
Cat .h
−1
)

BZP 618 16.77 0.91 15.86 0.1176 0.041 0.512


BZE 586 22.92 1.33 21.59 0.2029 0.013 0.716
BZEH 774 15.78 1.41 14.37 0.1407 0.003 0.367

12
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

Fig. 8. N2 adsorption/desorption isotherms and BJH pore size distribution curves for the completely deactivated (a) and the regenerated (b) different catalysts of the
H-[B]-ZSM-5 zeolite powder (BZP), extruded (BZE) and steam-treated extruded (BZEH).

acid sites. It also removes amorphous materials which exist around the deactivated samples exhibited the characteristic features of type-I iso-
open cavities [65]. Because of these advantages, during MTP process therm according to IUPAC classification. The pore size distribution
over BZEH catalyst, side reaction reduces and the internal/external curves of the deactivated catalysts showed that the mesopore volume is
coke deposition improves. For these reasons, the steam treated extruded significantly reduced. Also, as listed in Table 1, after the reaction,
catalyst (BZEH sample) having the appropriate catalytic lifetime was surface area and pore volume of the deactivated catalysts were sig-
chosen as the best catalyst for the MTP reaction. nificantly decreased. In case of BZEH sample, the surface area and pore
volume reduced from 368 to 258 m2/g and 0.23 to 0.147 cm3/g for the
3.3.2. N2 adsorption-desorption analysis totally deactivated zeolite. The percentage values of accessible surface
N2 adsorption/desorption isotherms along with their pore size dis- area and/or pore volume were also calculated by comparing those va-
tribution curves of the deactivated and regenerated catalysts are shown lues of the fresh and deactivated catalyst. The accessible surface area
in Fig. 8(a) and Fig. 8(b), respectively. Moreover, a summary of all reduced to almost 66%, 53% and 70% by N2 molecules for the com-
textural properties (surface area and pore volume) is listed in Table 1. pletely deactivated BZP, BZE and BZEH samples, respectively. The
Comparison between the isotherms obtained for the deactivated cata- pore/channel blockage with coke precursors might be a main reason for
lysts (Fig. 8(a)) and their fresh versions (Fig. 3) showed that all declining the surface area and pore volume after deactivation. Table 1

13
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

shows the decreasing Vmicro values (25.4%, 5.5%, and 15.3% for BZP, propylene selectivity to non-treated one, not only exhibits the longest
BZE and BZEH sample, respectively) which are lower than the Vmeso catalytic lifetime among these three types of catalysts, but also shows a
values (35%, 77%, and 49.3%). These results confirmed that coke for- further continues decrease in ethylene selectivity than the others. So, it
mation/deposition on the external surface was more serious. These maintains a stable formation of propylene for a longer period of time
results are in compliance with TGA analysis (see Table 5 and Fig. S6(b), and shows a higher continuous increase in P/E ratio than powder and
(c).These results could be ascribed to the coke precursor being easy to non-treated extruded catalysts. Moreover, this sample showed an ac-
continuously react to form coke in the micropores of BZP sample due to ceptable lifetime (750 h) in the MTP reaction compared to the powder
a large amount of internal Al, while the coke precursor was easy to H-[B]-ZSM-5 catalyst (580 h). Consequently, the combination of ex-
continuously react to form coke on the external surface of BZE and trusion and steaming treatments over the powder H-[B]-ZSM-5 catalyst
BZEH catalysts due to a large amount of external aluminum with little developed the best catalytic activity in the MTP reaction. This excellent
internal aluminum due to extrusion and steaming treatment. catalytic performance of hydrothermally treated extruded catalyst is
N2 adsorption-desorption of the deactivated catalysts changed to mainly due to its well-developed mesopores and low strong/weak acid
type-IV structure after regeneration process (Fig. 8(b)), which is quite sites ratio, which are helpful for attenuating coke deposition in the
similar to that of their fresh species. Also, according to Table 1, the micropore channels even at a high coking level.
surface area and pore volume of the regenerated BZP and BZEH samples
are approximately the same or slightly larger than those of their fresh Acknowledgments
ones. These results could be attributed to the formation of new cavities
using aggregate of tiny crystals and converting of some micropores to The authors would like to acknowledge the Petrochemical Research
mesopores types during the MTP reaction (hydrothermal treatment) and Technology Company for their support of the research.
and regeneration process at high temperature (500 °C). Furthermore,
these results indicated that decoking and regeneration processes have Appendix A. Supplementary data
been fully performed. In case of BZE sample, the surface area and pore
volume of the regenerated sample is slightly decreased compared to its Supplementary data to this article can be found online at https://
fresh species. The reason may be the agglomeration of the zeolite doi.org/10.1016/j.micromeso.2019.109699.
crystals/particles with alumina sol as a binder during extrusion process.
References
3.3.3. NH3-TPD analysis
The NH3-TPD profiles and acidity quantity of the regenerated cat- [1] Z. Hu, H. Zhang, L. Wang, H. Zhang, Y. Zhang, H. Xu, W. Shen, Y. Tang, Catal. Sci.
alysts are illustrated in Fig. S7 and Table 2. Similar to fresh samples, the Technol. 4 (2014) 2891–2895 https://doi.org/10.1039/C4CY00376D.
[2] F. Yaripour, Z. Shariatinia, S. Sahebdelfar, A. Irandoukht, J. Nat. Gas Sci. Eng. 22
NH3-TPD profiles of all regenerated catalysts indicated two main des- (2015) 260–269 https://doi.org/10.1016/j.jngse.2014.12.001.
orption peaks around 195 °C and 380 °C (Fig. S7), which correspond to [3] M. Rostamizadeh, A. Taeb, Synth. React. Inorg. Met. Org. Chem. 46 (2016) 665–671
NH3 desorption from weak acid sites (Lewis acidity) and strong acid https://doi.org/10.1080/15533174.2014.988825.
[4] H. Chen, M. Yang, W. Shang, Y. Tong, B. Liu, X. Han, J. Zhang, Q. Hao, M. Sun,
sites (Brønsted acidity), respectively. The acidity values (Table 2) show X. Ma, Ind. Eng. Chem. Res. 57 (2018) 10956–10966 https://doi.org/10.1021/acs.
that the hydrothermal treatment of H-[B]- ZSM-5, which occurs during iecr.8b00849.
the MTP reaction, creates a tremendous reduction of acidity especially [5] U.V. Mentzel, K.T. Højholt, M.S. Holm, R. Fehrmann, P. Beato, Appl. Catal. Gen.
417–418 (2012) 290–297 https://doi.org/10.1016/j.apcata.2012.01.003.
over BZP and BZE sample. Comparison of acidity of the fresh and re- [6] J. Lefevere, S. Mullens, V. Meynen, J.V. Noyen, Chem. Pap. 68 (2014) 1143–1153
generated samples shows that the concentration of strong acid sites https://doi.org/10.2478/s11696-014-0568-0.
(Brønsted acidity) decreases about 66%, 34% and 21% over BZP, BZE [7] F. Gorzin, J.T. Darian, F. Yaripour, S.M. Mousavi, J. Solid State Chem. 271 (2019)
8–22 https://doi.org/10.1016/j.jssc.2018.12.016.
and BZEH samples, respectively. In case of weak acid sites (Lewis
[8] N. Hadi, A. Niaei, R. Alizadeh, J. Raeisipour, C. R. Chim. 21 (2018) 523–540
acidity), decrease in concentration is about 45%, 55% and 23%, re- https://doi.org/10.1016/j.crci.2018.01.001.
spectively (Table 2). Furthermore, according to Table 2, the strength of [9] L. Zhang, Y. Song, G. Li, Q. Zhang, S. Zhang, J. Xu, F. Deng, Y. Gong, RSC Adv. 5
the weak and the strong acid sites of BZEH catalyst is lower than that of (2015) 61354–61363 https://doi.org/10.1039/C5RA09561A.
[10] H. Chen, Y. Wang, F. Meng, C. Sun, H. Li, Z. Wang, F. Gao, X. Wang, S. Wang,
BZP and BZE sample. Also, it can be seen that decreasing of the con- Microporous Mesoporous Mater. 244 (2017) 301–309 https://doi.org/10.1016/j.
centration of the Brønsted and Lewis acid sites is the smallest over micromeso.2017.02.014.
BZEH sample after performing the first cycle of the MTP reaction. [11] F. Yaripour, Z. Shariatinia, S. Sahebdelfar, A. Irandoukht, Microporous Mesoporous
Mater. 203 (2015) 41–53 https://doi.org/10.1016/j.micromeso.2014.10.024.
Consequently, it can be predicted that the catalytic performance of [12] J. Liu, C. Zhang, Z. Shen, W. Hua, Y. Tang, W. Shen, Y. Yue, H. Xu, Catal. Commun.
BZEH sample is better than that of the other catalyst in the second cycle 10 (2009) 1506–1509 https://doi.org/10.1016/j.catcom.2009.04.004.
of the MTP reaction. [13] M. Firoozi, M. Baghalha, M. Asadi, Catal. Commun. 10 (2009) 1582–1585 https://
doi.org/10.1016/j.catcom.2009.04.021.
[14] J. Ahmadpour, M. Taghizadeh, Synth. React. Inorg. Met. Org. Chem. 46 (2016)
4. Conclusions 1133–1141 https://doi.org/10.1080/15533174.2015.1004433.
[15] H. Chen, Y. Wang, C. Sun, F. Gao, L. Sun, C. Wang, Z. Wang, X. Wang, Catal.
Commun. 100 (2017) 107–111 https://doi.org/10.1016/j.catcom.2017.06.048.
In this work, the effects of steaming treatment on the physico- [16] H. Li, Y. Wang, C. Fan, C. Sun, X. Wang, C. Wang, X. Zhang, S. Wang, Appl. Catal.
chemical properties and catalytic performance of high-silica H-[B]- Gen. 551 (2018) 34–48 https://doi.org/10.1016/j.apcata.2017.12.007.
ZSM-5 extruded catalyst in the MTP reaction have been investigated, [17] J. Ahmadpour, M. Taghizadeh, C. R. Chim. 18 (2015) 834-847.https://doi.org/10.
1016/j.crci.2015.05.002.
and the results have been compared with those of the powder and non-
[18] F. Gorzin, J.T. Darian, F. Yaripour, S.M. Mousavi, RSC Adv. 8 (2018) 41131–41142
treated extruded H-[B]-ZSM-5. https://doi.org/10.1039/C8RA08624A.
According to the MTP conversion experiments under the similar [19] J. Li, H. Ma, Y. Chen, Z. Xu, C. Li, W. Ying, Chem. Commun. 54 (2018) 6032–6035
operation conditions, the extruded catalyst exhibits a unique ad- https://doi.org/10.1039/C8CC02042F.
[20] M. Rostamizadeh, F. Yaripour, H. Hazrati, J. Porous Mater. 25 (2018) 1287–1299
vantage, lowering ethylene and C1–C4 saturated hydrocarbons se- https://doi.org/10.1007/s10934-017-0539-2.
lectivity, and increasing the propylene and butylenes selectivity. [21] M. Rostamizadeh, F. Yaripour, J. Taiwan Inst. Chem. Eng. 71 (2017) 454–463
However, due to the loss of some strong acid sites and partial blockage https://doi.org/10.1016/j.jtice.2016.12.003.
[22] R. Feng, X. Yan, X. Hu, Z. Yan, J. Lin, Z. Li, K. Hou, M.J. Rood, Catal. Commun. 109
of micropores during the extrusion process, the lifetime of extruded (2018) 1–5 https://doi.org/10.1016/j.catcom.2018.02.005.
catalyst is somewhat shorter compared to that of powder type, which is [23] I. Yarulina, K. De Wispelaere, S. Bailleul, J. Goetze, M. Radersma, E. Abou-Hamad,
unfavorable for the current industrial practice using multi-stage fixed- I. Vollmer, M. Goesten, B. Mezari, E.J.M. Hensen, J.S. Martinez-Espin, M. Morten,
S. Mitchell, J. Perez-Ramirez, U. Olsbye, B.M. Weckhuysen, V. Van Speybroeck,
bed reactor. F. Kapteijn, J. Gascon, Nat. Chem. 10 (2018) 804–812 https://doi.org/10.1038/
On the contrary, steam-treated extruded catalyst, with similar

14
M.S. Beheshti, et al. Microporous and Mesoporous Materials 291 (2020) 109699

s41557-018-0081-0-ris. [46] C.D. Chang, S.D. Hellring, J.N. Miale, K.D. Schmitt, P.W. Brigandi, E.L. Wu, J.
[24] T.L. Cui, L.B. Lv, W.B. Zhang, X.H. Li, J.S. Chen, Catal. Sci. Technol. 6 (2016) Chem. Soc., Faraday Trans. 1: Phys. Chem. Condens. Phases 81 (1985) 2215–2224
5262–5266 https://doi.org/10.1039/C6CY00379F. https://doi.org/10.1039/F19858102215.
[25] J. Li, M. Liu, X. Guo, S. Xu, Y. Wei, Z. Liu, C. Song, ACS Appl. Mater. Interfaces 9 [47] R. Caicedo-Realpe, J. Pérez-Ramírez, Microporous Mesoporous Mater. 128 (2010)
(2017) 26096–26106 https://doi.org/10.1021/acsami.7b07806. 91–100 https://doi.org/10.1016/j.micromeso.2009.08.009.
[26] F. Gorzin, F. Yaripour, Res. Chem. Intermed. 45 (2019) 261–285 https://doi.org/ [48] H. Xin, A. Koekkoek, Q. Yang, R. van Santen, C. Li, E.J.M. Hensen, Chem. Commun.
10.1007/s11164-018-3601-z. 0 (2009) 7590–7592 https://doi.org/10.1039/B917038C.
[27] M. Jiao, S. Fan, J. Zhang, X. Su, T.-S. Zhao, Catal. Commun. 56 (2014) 153–156 [49] C.S. Triantafillidis, A.G. Vlessidis, N.P. Evmiridis, Ind. Eng. Chem. Res. 39 (2000)
https://doi.org/10.1016/j.catcom.2014.07.025. 307–319 https://doi.org/10.1021/ie990568k.
[28] M. Rostamizadeh, A. Taeb, J. Ind. Eng. Chem. 27 (2015) 297–306 https://doi.org/ [50] J. Ahmadpour, M. Taghizadeh, J. Nat. Gas Sci. Eng. 23 (2015) 184–194 https://doi.
10.1016/j.jiec.2015.01.004. org/10.1016/j.jngse.2015.01.035.
[29] N. Hadi, A. Niaei, S.R. Nabavi, R. Alizadeh, M. Navaei Shirazi, B. Izadkhah, J. [51] M. Rostamizadeh, F. Yaripour, H. Hazrati, J. Anal. Appl. Pyrolysis 132 (2018) 1–10
Taiwan Inst. Chem. Eng. 59 (2015) 173–185 https://doi.org/10.1016/j.jtice.2015. https://doi.org/10.1016/j.jaap.2018.04.003.
09.017. [52] L.C. Ong, Nature and Stability of Aluminum Species in HZSM-5: Changes upon
[30] P. Losch, G. Laugel, J.S. Martinez-Espin, S. Chavan, U. Olsbye, B. Louis, Top. Catal. Hydrothermal Treatment and Effect of Binder, PhD dissertation Technische
58 (2015) 826–832 https://doi.org/10.1007/s11244-015-0449-y. Universität München, 2009.
[31] M. Rostamizadeh, F. Yaripour, Fuel 181 (2016) 537–546 https://doi.org/10.1016/ [53] R. Bingre, R. Li, Q. Wang, P. Nguyen, T. Onfroy, B. Louis, Catalysts 9 (2019)
j.fuel.2016.05.019. 545–556 https://doi.org/10.3390/catal9060545.
[32] Y. Yang, C. Sun, J. Du, Y. Yue, W. Hua, C. Zhang, W. Shen, H. Xu, Catal. Commun. [54] F. Lónyi, J. Valyon, Microporous Mesoporous Mater. 47 (2001) 293–301 https://
24 (2012) 44–47 https://doi.org/10.1016/j.catcom.2012.03.013. doi.org/10.1016/S1387-1811(01)00389-4.
[33] A. Xu, H. Ma, H. Zhang, D. Weiyong, D. Fang, Pol. J. Chem. Technol. 15 (2013) [55] C.T.-W. Chu, G.H. Kuehl, R.M. Lago, C.D. Chang, J. Catal. 93 (1985) 451–458
95–101 https://doi.org/10.2478/pjct-2013-0075. https://doi.org/10.1016/0021-9517(85)90192-7.
[34] J.S.J. Hargreaves, A.L. Munnoch, Catal. Sci. Technol. 3 (2013) 1165–1171 https:// [56] G. Coudurier, J.C. Vedrine, Stud. Surf. Sci. Catal. 28 (1986) 643–652 https://doi.
doi.org/10.1039/C3CY20866D. org/10.1016/S0167-2991(09)60930-7.
[35] S.D. Kim, S.C. Baek, Y.-J. Lee, K.-W. Jun, M.J. Kim, I.S. Yoo, Appl. Catal. Gen. 309 [57] K.F.M.G.J. Scholle, A.P.M. Kentgens, W.S. Veeman, P. Frenken, G.P.M. Van der
(2006) 139–143 https://doi.org/10.1016/j.apcata.2006.05.008. Velden, J. Phys. Chem. 88 (1984) 5–8 https://doi.org/10.1021/j150645a003.
[36] N.-L. Michels, S. Mitchell, J. Pérez-Ramírez, ACS Catal. 4 (2014) 2409–2417 [58] X.-D. Chen, X.-G. Li, H. Li, J.-J. Han, W.-D. Xiao, Chem. Eng. Sci. 192 (2018)
https://doi.org/10.1021/cs500353b. 1081–1090 https://doi.org/10.1016/j.ces.2018.08.047.
[37] K.-Y. Lee, H.-K. Lee, S.-K. Ihm, Top. Catal. 53 (2015) 247–253 https://doi.org/10. [59] J.M. Fougerit, N.S. Gnep, M. Guisnet, P. Amigues, J.L. Duplan, F. Hugues, Stud.
1007/s11244-009-9412-0. Surf. Sci. Catal. 84 (1994) 1723–1730 https://doi.org/10.1016/S0167-2991(08)
[38] J. Freiding, B. Kraushaar-Czarnetzki, Appl. Catal. Gen. 391 (2011) 254–260 https:// 63725-8.
doi.org/10.1016/j.apcata.2010.05.035. [60] S. Mitchell, N.-L. Michels, J. Pérez-Ramírez, Chem. Soc. Rev. 42 (2013) 6094–6112
[39] Y.-J. Lee, Y.-W. Kim, N. Viswanadham, K.-W. Jun, J.W. Bae, Appl. Catal. Gen. 374 https://doi.org/10.1039/C3CS60076A.
(2010) 18–25 https://doi.org/10.1016/j.apcata.2009.11.019. [61] V.R. Choudhary, P. Devadas, A.K. Kinage, M. Guisnet, Appl. Catal. Gen. 162 (1997)
[40] S. Zhang, Y. Gong, L. Zhang, Y. Liu, T. Dou, J. Xu, F. Deng, Fuel Process. Technol. 223–233 https://doi.org/10.1016/S0926-860X(97)00100-2.
129 (2015) 130–138 https://doi.org/10.1016/j.fuproc.2014.09.006. [62] W. Wu, W. Guo, W. Xiao, M. Luo, Chem. Eng. Sci. 66 (2011) 4722–4732 https://doi.
[41] Y. Song, L.-L. Zhang, G.-D. Li, Y.-S. Shang, X.-M. Zhao, T. Ma, L.-M. Zhang, Y.- org/10.1016/j.ces.2011.06.036.
L. Zhai, Y.-J. Gong, J. Xu, F. Deng, Fuel Process. Technol. 168 (2017) 105–115 [63] M. Bjørgen, S. Svelle, F. Joensen, J. Nerlov, S. Kolboe, F. Bonino, L. Palumbo,
https://doi.org/10.1016/j.fuproc.2017.08.020. S. Bordiga, U. Olsbye, J. Catal. 249 (2007) 195–207 https://doi.org/10.1016/j.jcat.
[42] J. Li, M. Liu, X. Guo, C. Dai, S. Xu, Y. Wei, Z. Liu, C. Song, Ind. Eng. Chem. Res. 57 2007.04.006.
(2018) 8190–8199 https://doi.org/10.1021/acs.iecr.8b00513. [64] C. Chen, Q. Zhang, Z. Meng, C. Li, H. Shan, Appl. Petrochem. Res. 5 (2015) 277–284
[43] C.S. Triantafillidis, A.G. Vlessidis, L. Nalbandian, N.P. Evmiridis, Microporous https://doi.org/10.1007/s13203-015-0129-7.
Mesoporous Mater. 47 (2001) 369–388 https://doi.org/10.1016/S1387-1811(01) [65] X. Zhao, Y. Hong, L. Wang, D. Fan, N. Yan, X. Liu, P. Tian, X. Guo, Z. Liu, Chin. J.
00399-7. Catal. 39 (2018) 1418–1426 https://doi.org/10.1016/S1872-2067(18)63117-1.
[44] L. Jin, H. Hu, S. Zhu, B. Ma, Catal. Today 149 (2010) 207–211 https://doi.org/10. [66] Z. Wan, G.K. Li, C. Wang, H. Yang, D. Zhang, Appl. Catal. Gen. 549 (2018) 141–151
1016/j.cattod.2009.07.088. https://doi.org/10.1016/j.apcata.2017.09.035.
[45] D.S. Shihabi, W.E. Garwood, P. Chu, J.N. Miale, R.M. Lago, C.T.-W. Chu, [67] K. Lee, S. Lee, Y. Jun, M. Choi, J. Catal. 347 (2017) 222–230 https://doi.org/10.
C.D. Chang, J. Catal. 93 (1985) 471–474 https://doi.org/10.1016/0021-9517(85) 1016/j.jcat.2017.01.018.
90195-2.

15

You might also like