(Kella, 2021) ZN Effect On Pore Size

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Microporous and Mesoporous Materials 323 (2021) 111216

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

Enhanced selectivity of benzene-toluene-ethyl benzene and xylene (BTEX)


in direct conversion of n-butanol to aromatics over Zn modified
HZSM5 catalysts
Tatinaidu Kella, Debaprasad Shee *
Department of Chemical Engineering, Indian Institute of Technology Hyderabad, Kandi, Sangareddy 502 285, Telangana, India

A R T I C L E I N F O A B S T R A C T

Keywords: The sustainable aromatics production from renewable source is indispensable to circumvallate the dependence
n-butanol on fossil fuel for commodity chemicals. Direct conversion of n-butanol to aromatics using xZn-HZSM5 (0–10 wt
Aromatics %, SiO2/Al2O3 = 23, 55 and 280) catalysts was proposed in the present investigation. Characterization studies
BTEX
revealed the formation of different Zn species depending on the Zn loading. Furthermore, XPS analysis confirmed
Zn-HZSM5
XPS
strong interaction of Zn species with the electronegative oxygen atom of zeolite framework. Zn deposition
Coke analysis enhanced the selectivities of aromatics and BTEX. The presence of Zn species suppressed the hydride transfer
reaction and promoted the dehydrogenation reaction resulting in higher selectivities of aromatics and BTEX. The
rising temperature and declining in WHSV enhanced the selectivities of aromatics and BTEX. The high pressure
had adverse effect on the selectivities of aromatics. The maximum selectivity of total aromatics (74.83%) and
BTEX (68.75%) were achieved at 723 K, 1 bar pressure and 0.75 h− 1 of WHSV. The coke analysis revealed the
formation of polynuclear aromatic coke at high pressure. The alkyl substitute aromatics and polynuclear aro­
matics were formed predominately on pure HZSM5 (55) than 5Zn-HZSM5 (55). A plausible reaction mechanism
was suggested considering the products distribution.

1. Introduction climate problems are the strong motivation behind the exploration of
alternate sustainable renewable resources to produce the aromatics,
Dependency on fossil fuels is increasing as the demands for organic especially BTEX. Renewable sources like solar, wind, thermal, tidal and
chemicals and fuels are rising continuously because of the improvement geothermal energies can be used to produce electricity and heat energy
of quality of life, which leads to faster depletion of fossil fuels and causes but are incompetent to produce liquid fuels and organic chemicals.
harmful effect to the environment. Especially, the demand for the BTEX Biomass is the only carbon-neutral renewable resource that can be used
(benzene, toluene, ethyl benzene and xylenes) has increased in recent to produce liquid transportation fuels and organic chemicals in sus­
years [1,2]. BTEX in aromatics family are the most important feedstocks tainable manner [13,14]. Biomass derived alcohols (methanol, ethanol,
to the petrochemical industry for the production of array of commodity propanol and butanol) are one such alternative for the production of
chemicals, which are the key ingredients for the manufacture of auto­ aromatics using solid acid catalysts such as zeolites (HZSM, SAPO, H-β,
motive parts, building materials, perfumery, medicine, etc. [3]. For Mordenite etc.), and metal incorporated zeolites [3,15–17]. It has been
example, toluene and xylenes by blending with gasoline are used as reported in the published literature that the transformation of methanol
octane enhancers [4]. Another important chemical p-xylene is used for and ethanol to aromatics resulted in the formation of undesired higher
the production of terephthalic acid, which can be polymerized to pro­ aromatics. Alcohols with higher carbon to oxygen ratio, produce hy­
duce polyester and polyethylene terephthalate (PETE) [5–8]. Conven­ drocarbon with higher theoretical yield [18]. From the structural point
tionally, these BTEX are produced via petroleum routes like naphtha of view, butylenes obtained from n-butanol dehydration are easily oli­
reforming and oil cracking, and non-petroleum routes like aromatiza­ gomerized and aromatized compared to methanol and ethanol derived
tion of paraffins (methane, ethane, propane, butanes) derived from coal olefins because of their lower activation energy [18]. The catalytic ac­
and natural gas [9–12]. Depletion of fossil fuels and growing global tivity for the production of liquid range hydrocarbons is higher for

* Corresponding author.
E-mail address: dshee@che.iith.ac.in (D. Shee).

https://doi.org/10.1016/j.micromeso.2021.111216
Received 3 February 2021; Received in revised form 6 May 2021; Accepted 31 May 2021
Available online 5 June 2021
1387-1811/© 2021 Elsevier Inc. All rights reserved.
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

n-butanol compared to methanol, ethanol and propanol [18]. n-butanol formation of toluene and xylene and also inhibited the formation of light
is considered as a fuel, and potentially important diesel and gasoline alkanes. From the above discussion, it is evident that metal incorporated
additive, to reduce the emissions and soot [19]. Moreover, n-butanol zeolites are effective for the production of aromatics from light alkanes
also has been used as a solvent and feedstock for the production of useful and alcohols. However, conversion of n-butanol to aromatics was not
derivatives [20,21]. In recent years, butanol isomers have drawn studied over metal incorporated zeolite catalysts.
numerous attention to produce butylenes, gasoline range hydrocarbons Based on the above discussion, it is also evident that incorporation of
and aromatics [3,22–25]. n-butanol is being produced commercially by suitable metal/metal oxide can improve the selectivity of aromatics and
propylene hydroformylation to aldehydes followed by hydrogenation BTEX. In the present investigation, Zn incorporated HZSM5 catalysts of
[26]. Traditionally, bio-butanol is being produced from biomass through different Zn loading and silica alumina ratios (23, 55 and 280) were
ABE (Acetone-Butanol-Ethanol) fermentation of sugars by using clos­ studied for the direct conversion of n-butanol to aromatics (BTA). The
tridium microorganism [27]. promotional effect on the selectivities of total aromatics and BTEX was
Product distribution in the alcohol conversion to aromatics is discussed. All the catalysts were well characterized by a number of
affected mostly by the nature of acidic sites and textural properties of the techniques such as BET, XRD, NH3-TPD, pyridine FTIR, UV–vis and XPS
catalyst. Hence, development of efficient catalysts and choosing the and the physico-chemical properties of these catalysts were correlated
appropriate process conditions are the key factors, which influence the with the selectivity of different aromatics formed under wide range of
aromatic product distribution [22,25,28]. So far, numerous studies on process conditions (temperature, pressure and WHSV). The alcohols to
the conversion of alcohols to various range of hydrocarbons have been aromatics reaction suffers from catalyst deactivation because of the
carried out using different types of zeolites such as MFI zeolites, H-β, deposition of carbonaceous species. The spent Zn-HZSM5 catalysts were
faujasite, offretite, and mordenite zeolites [3,22,25,29,30]. Among the well characterized by various characterizations to delineate character­
MFI zeolites family, the HZSM5 catalyst has shown superior perfor­ istics of coke deposited on the catalyst surface. A plausible reaction
mance in the conversion of different alcohols (methanol, ethanol, mechanism was suggested based on the product distribution derived
propanol and butanol) to olefins, gasoline range hydrocarbons and ar­ under broad range of process conditions.
omatics due to their excellent shape selectivity, tunable acidity
(Brønsted and Lewis) and pore size (0.55 nm) matching with kinetic 2. Experimental
diameter of BTX [15,18,25,31]. However, the two major challenges
faced during alcohol aromatization were selectivity of BTEX and catalyst 2.1. Materials
deactivation. The conversion of n-butanol to hydrocarbons over HZSM5
catalyst achieved nearly 38% selectivity for aromatics (selectivity for The ammonium form of HZSM5 [SiO2/Al2O3 ratio = 23, 55 and 280]
BTEX around 27%) along with the high proportion of light alkanes like was purchased from Zeolyst International, USA. n-butanol [C4H9OH,
propane and butanes [22]. To improve the selectivity for BTEX, the Purity ≥ 99.0%] was purchased from Merck Life Science Pvt. Ltd. and
formation of light alkanes and higher aromatics is required to suppress. zinc nitrate hexahydrate [Zn(NO3)2.6H2O, Purity ≥ 98.0%] was pro­
Apart from the pore size, catalyst acidity and dehydrogenation activity cured from Sigma-Aldrich. HZSM5 zeolite was heat treated at 823 K for
are two important factors to be considered for enhancing the aromati­ 6 h and all the other chemicals were used without further purification.
zation [3,22]. Brønsted acid sites within the zeolites are involved in the
oligomerization of olefins, which are then aromatized to form aromatic 2.2. Catalyst preparation
hydrocarbons. Moreover, the strength and density of acid sites have
significant influence on the selectivity of total aromatics and BTEX [25]. Several zinc (Zn) incorporated HZSM5 catalysts of different Zn
On the other hand, the MFI zeolites supported metal or metal oxides are loading were prepared by incipient wetness impregnation method. The
effective for the enhancement of dehydrogenation activity, which results Zn precursor solution was prepared by adding a calculated amount of
increase in the selectivity of aromatics and BTEX. Numerous literatures zinc nitrate hexahydrate precursor corresponding to the particular Zn
are available on the conversion of alcohols and light alkanes and alkenes loading to the incipient volume of millipore water. The required amount
to aromatics over different metal (Zn, Ga, Re, Mo, Pt) modified zeolites of pretreated HZSM5 was added to the precursor solution and stirred for
[3,17,32–38]. These metal or metal oxides enhance the dehydrogena­ 1 h to ensure intimate mixing. The pretreatment of HZSM5 zeolite was
tion and aromatization of alkanes. These metal cations limit the pref­ performed by adding incipient volume of millipore water followed by
erential elimination of hydrogen by transferring to olefins (hydride heat treatment similar to catalyst preparation. The wet paste was kept in
transfer) resulting in the formation of light alkanes. Recently, Lok et al. a desiccator for drying at room temperature for overnight (about 12 h).
reviewed the promotional effect of Ga and Zn on the production of The room temperature dried material was dried further in a hot air oven
bio-aromatics during thermochemical conversion of biomass and waste for 6 h at 383K. Finally, the dried material was calcined in the static air
in the temperature range of 673–773 K and space velocity of 0.5–1.0 h− 1 at 823 K for 6 h using a heating rate of 1 K/min. The calcined catalysts
[17]. It was reported that the extraframework Ga cations and close were denoted as xZn-HZSM5 (x = 0, 1, 3,5,7 and 10 wt%, SiO2/Al2O3 =
proximity Brønsted acid sites were active for the aromatization, whereas 23, 55 and 280), where x represents wt% of Zn metal.
ZnOH+ species seem to be the active sites for the aromatization reaction
over Zn/HZSM catalyst. The aromatization of various alkanes and al­ 2.3. Catalyst characterization
kenes such as methane, ethane, ethylene, propane, butane, pentane, and
hexane has been reported over Zn/HZSM5 catalysts [33,34,37–43]. The N2 adsorption-desorption studies were performed over xZn-
Wang et al. investigated methanol to aromatic (MTA) reaction in flu­ HZSM5 catalysts at 77 K using physisorption analyzer (Micromeritics
idized bed reactor over Zn/HZSM5 catalyst [36]. p-xylene was reported ASAP2020). Before the measurements, all the catalysts were heat
as the primary product in MTA reaction. Other xylenes were formed by treated at 473 K for 5 h under vacuum (5 × 10− 5 mmHg) in order to
isomerization of p-xylenes. Moreover, benzene, toluene and trimethyl eliminate adsorbed moisture and residual impurities. The Brunauer-
benzene were formed by dealkylation, alkylation or disproportionation Emmett-Teller (BET) equation and adsorption isotherm data in the
of xylene. The single step conversion of isobutyl alcohol to aromatic relative pressure (P/P0) range of 0.05–0.30 were used to determine the
hydrocarbons over different zeolites and metals modified zeolites was surface area of the catalysts. The pore size distribution (PSD) was
reported by Yu et al. [3]. The Zn modified HZSM5 (Si/Al ratio 34.3) determined by Barrett-Joyner-Halenda (BJH) method and average pore
produced highest aromatic yields (~60 wt%) among all the supported volume was considered as the volume of liquid nitrogen adsorbed at P/
metal species (Zn, Ga, Mo, La, Ni, Ag, and Pt) at 0.1 MPa, 723 K, WHSV P0 = 1.0 ca. The micropore area and volume were determined by t-plot
of 3.88 h− 1 [3]. Zn modified HZSM5 exhibited improvement in the method.

2
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

The powder X-ray diffraction (powder XRD) patterns of xZn-HZSM5 at this temperature for 10 min. Then the carrier gas was switched over to
catalysts were acquired by using Rigaku Smart Lab X-ray Diffractometer air (30 ml/min) and held at this temperature for 30 min to completely
equipped with a Cu-Kα radiation (λ = 1.5418 Å) source and D/teX ultra burn the deposited coke on the catalyst. The amount of coke was
250 detector and operated at 45 kV and 30 mA. The XRD patterns were considered as the percentage loss in weight.
acquired in the 2θ range of 10 to 90◦ with a scanning speed of 2◦ /min.
The relative crystallinity of the xZnHZSM5 catalysts was determined by
the integral areas of the peaks in the 2θ range of 22.0–25.0◦ with the 2.4. Reaction studies
pure HZSM5 as reference [44]. The crystallize size of the HZSM5 was
determined using Scherrer equation. The butanol to aromatics (BTA) reaction was performed in a high
FTIR spectra (Fourier Transform Infrared) under ambient condition pressure and temperature fixed bed reactor over xZn-HZSM5 catalysts
of the fresh and spent xZn-HZSM5 catalysts were recorded using a under different reaction conditions. The catalyst bed was made with the
Bruker TENSOR 37 FTIR spectrometer in transmission mode in the physical mixture of measured amount of catalyst and quartz particles.
wavenumber range of 400–4000 cm− 1 with a resolution of 4 cm− 1 and The catalyst bed supported by two layers of quartz wool was placed
256 no of scans. Prior to the analysis, all the catalyst samples were mixed inside a stainless steel reactor (1-inch ID × 8-inch length) and the reactor
with spectroscopic grade KBr powder and transformed into a trans­ loaded with catalyst was placed inside a vertically placed tubular
parent circular wafer. All the spectra were taken against the KBr back­ furnace. The catalyst bed temperature was measured and controlled by a
ground acquired separately under similar conditions. K type thermocouple placed immediately above the bed and PID tem­
The IR spectra of pyridine adsorbed xZn-HZSM5 catalysts were perature controller. The catalyst was first pretreated at the reaction
recorded in the above mentioned spectrometer under similar conditions. temperature in flowing of nitrogen (40 mL/min) for 1 h. The mass
A few drops of pyridine were added to the powder catalyst sample and flowrate of nitrogen was maintained by a thermal mass flow controller
then left at room temperature for 8 h. Then the pyridine soaked sample (Bronkhorst, EL-Flow). On completion of pretreatment, n-butanol was
was kept in hot air oven maintained at 383 K for 1 h to eliminate pumped at a certain flowrate (0.1 mL/min) using HPLC pump (Lab
physisorbed pyridine. A transparent circular wafer was made in similar alliance, Series-I) into the reactor through a preheater maintained at
way as mentioned above for acquiring the IR spectra. 473 K. The vaporized butanol in preheater was carried by nitrogen into
The temperature programmed NH3 desorption (NH3-TPD) studies of the reactor. The product gas mixture releasing from the exit of the
xZn-HZSM5 catalysts were performed in Micromeritics Chemisorption reactor was routed through a vertical condenser, where low boiling
analyzer (Autochem 2920 II, Micromeritics, USA). Before the analysis, organic compounds were condensed. The condensed liquid products
all the catalysts were pretreated at 473 K in flowing of Ar for 1 h to were separated from non-condensable gas mixture in a high pressure
eliminate adsorbed moisture and any other residual impurities. On gas-liquid separator. The liquid products were collected at regular in­
completion of pretreatment, the catalyst temperature was lowered to terval of time and the non-condensable gaseous products were sent for
373 K in flowing of Ar and 10 vol%NH3/Ar gas mixture was fed at a flow analysis to the online Gas chromatograph (GC). The permanent gases
rate of 20 mL/min. The pretreated catalyst was exposed to NH3/Ar gas (CH4, N2 and H2) were quantified using a micro GC (Agilent, Model no:
mixture for 1 h and subsequently the gas flow was changed to pure Ar. 490C) equipped with thermal conductivity detector (TCD) connected to
The Ar gas flow was maintained for another 1 h to remove physisorbed PPU (pora plot U) and MSSA (CP-molsieve) columns. The gaseous hy­
NH3 from catalyst surface and entire flow path. The catalyst was heated drocarbons (C1–C5 alkanes and alkenes) were quantified by online GC
to 1073 K using a heating rate of 10 K/min. The desorbed NH3 was (Agilent series, Model no. 7890B) equipped with flame ionization de­
monitored using a thermal conductivity detector (TCD). tector (FID) and GS-GasPro capillary column (Agilent J&W, 30 m × 0.32
UV–vis spectra of the xZn-HZSM5 (both fresh and spent) catalysts mm). The liquid products were analyzed offline by Gas chromatograph
were recorded by using UV–vis NIR PerkinElmer spectrophotometer equipped with FID and DB-5HT column (Phenomenex, 30 m × 0.32 mm
(Model No.: Lambda 1050) equipped with Harrick scientific diffuse × 0.10 μm). GC FID operating conditions for both gas and liquid anal­
reflectance accessory (Praying Mantis, Model No.: DRK-4-P72). The ysis, the injector and detector temperatures were kept at 523 K, and oven
spectra of all the samples were collected in the wavelength region of temperature varying from 313 K to 513 K. The phenanthrene was used
200–800 nm using BaSO4 as reference. as the external standard for the quantification purpose. Methane was
X-ray photoelectron spectroscopy (XPS) measurement of selected considered as the tie component to determine complete gas phase
xZn-HZSM5 catalysts was performed on Thermo Scientific™ K-Alpha™ compositions. The conversion of n-butanol over all the tested catalysts
XPS spectrometer (UK) using a micro focused monochromatized AlKα and reactions conditions was achieved 100% and results reported were
radiation. The powder catalyst was dispersed on a sample holder using considered beyond 5 h TOS. Material and carbon balance was verified
double-sided tape before introducing into the treatment chamber. The for all the conducted experiments and the errors were in the range of
sample pretreatment compartment was maintained under high vacuum ±5% for atmospheric pressure experimental runs. The carbon and ma­
(ca. 10− 8 Torr) before shifting to the analysis chamber. XPS survey scan terial balance for high-pressure reactions were nearly 76–93% as shown
spectra were acquired using 200 eV analyzer electron pass, while the in Table 2. All the liquid and gaseous products were identified using gas
high resolution XPS region scan spectra were obtained at 50 eV pass chromatography mass spectroscopy (GC-MS, Shimazdu GC-MS QP2010
energy within a short period of time to avoid or minimize the photo- Ultra). The selectivity of the products and WHSV were calculated by the
reduction. The binding energies calibration was performed using following equations (1) and (2) [22].
adventitious carbon (C 1s) peak appeared at 284.6 eV [45].
Selectivity ​ of ​ Cn ​ hydrocarbon, ​ % = 100
The amount of carbonaceous species (coke) deposited on the spent
HZSM5 and 5Zn-HZSM5 catalysts was quantified using thermogravi­ ×[
n × (moles ​ of ​ Cn ​ hydrocabon ​ produced)
] (1)
metric analyzer (TGA, Perkin Elmer Pyris 1 TGA). About 6–10 mg of 4 × (moles ​ of ​ butanol ​ reacted)
spent catalyst sample was loaded in the platinum crucible and kept in­
Where ‘n’ refers to the number of carbon atoms present in the particular
side the furnace. The sample was then heated to 1073 K using a heating
hydrocarbon.
rate of 10 K/min in flowing of nitrogen (30 ml/min) carrier gas and held

3
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Total ​ massflowrate ​ of ​ butanol ​ and ​ N2 ​ gas ​ (g/h)


Weight ​ hourly ​ space ​ velocity ​ (WHSV), ​ h− 1 = (2)
weight ​ of ​ the ​ catalyst ​ (g)

at low Zn loading (upto 3 wt%). At higher Zn loading, ZnO was also


deposited in micropores of xZn-HZSM5 (55) in addition to the meso­
3. Results and discussion pores and or macro pores. The decrease in surface area was because of
the increase in surface coverage and blockage of pores by deposited ZnO.
3.1. Catalyst characterization A slight increase in pore diameter for higher Zn loading catalyst in­
dicates the blockage of micropores.
The textural properties of the xZn-HZSM5 (x = 0 to 10 wt%; SiO2/ The XRD patterns of xZn-HZSM5 (55) catalysts were depicted in
Al2O3 = 55) catalysts were determined by N2 adsorption/desorption Fig. 1. All the catalysts showed the characteristic diffraction peaks at 2θ
studies and the results were tabulated in Table 1. The BET surface area of of 8.01◦ 8.9◦ and 23.1◦ 23.9◦ 24.4◦ of MFI crystal structure. However,
xZn-HZSM5 (55) decreases with increasing the Zn loading. However, the intensity of these peaks for Zn-HZSM5 (55) catalysts decreases as the
decrease of total pore volume was insignificant for these catalysts. The Zn loading increased in comparison to HZSM5 (55). The characteristics
micropore analysis of these catalysts was performed using t-plot method. diffraction peaks corresponding to ZnO were not noticed for any of these
The micropore area (Am) was increased by about 35% on deposition of 1 xZn-HZSM5 (55) catalysts indicating the higher dispersion of ZnO.
wt% ZnO and remained constant on further increase in ZnO loading upto However, the diffraction peak at 2θ of 23.1◦ shifted to lower angle in Zn-
3 wt%. The initial increase in micropore area might be because of the HZSM5 catalysts. The shift of diffraction peak around this region is
transformation of relatively lower size mesopores to micropores by the attributed to the different Zn content and skeleton dealumination of
deposition of ZnO. Further increase in the Zn loading (at and above 5 wt zeolite lattice [46]. The XRD patterns of 5Zn-HZSM5 of different
%) decreases the micropore area and cumulative decrease of micropores silica-alumina ratio (23, 55 and 280) were also shown in Fig. S4. Similar
was about 13% with the increase in ZnO loading from 1 to 10 wt%. features were noticed in the XRD patterns of 5Zn-HZSM5 of different
However, the external surface area (AE), which includes mesopores and silica-alumina ratio. The intensity of peaks characteristics of HZSM5
macropore/crystal void space is gradually decreased (loss of 48%) with zeolite (2θ = 8.01◦ , 8.9◦ ) was noticed to decrease with the increase in
increase in Zn loading from 0 to 10 wt%. An insignificant variation of ZnO loading and silica-alumina ratio for a particular ZnO loading
micropore volume was noticed with the increase in Zn loading. Again, resulting in decrease in relative crystallinity of xZn-HZSM5 catalysts
the relative decrease of external surface was much low above 5 wt%. (Table S2). However, the relative crystallinity was noticed to be affected
Furthermore, an insignificant variation of micropore volume was more with the increase silica-alumina ratio of HZSM (Table S2).
noticed with the increase in Zn loading. These results indicates that the The FTIR spectra of xZn-HZSM5 (55) catalysts in the hydroxyl region
ZnO was preferentially deposited in the mesopores and or macro pores exhibit three distinct IR bands at 3660, 3611 and 3437 cm− 1 (Fig. 2).

Table 1
Textural properties of HZSM5 (55) and xZn-HZSM5 (55) catalysts.
Catalyst SABETa (m2/g) AEb (m2/g) VTc (cm3/g) Vmd (cm3/g) DAAe (nm) Amf (m2/g)

HZSM5 (55) 373 195 0.29 0.12 5.9 162


1Zn-HZSM5 (55) 360 141 0.29 0.11 5.86 219
3Zn- HZSM5 (55) 343 125 0.28 0.11 6.3 218
5Zn- HZSM5 (55) 309 109 0.24 0.10 6.15 200
7Zn- HZSM5 (55) 307 104 0.25 0.10 6.90 203
10Zn- HZSM5 (55) 291 101 0.25 0.10 7.74 190
a
BET surface area.

b
External surface area derived from t-Plot method.

c
Total pore volume derived from BJH method.

d
Micropore volume derived from t-Plot method.

e
BJH adsorption average pore diameter (4V/A).

f
Micropore surface area derived from t-Plot method.

Table 2
Product distribution over HZSM5 (55) and 5Zn-HZSM5 (55) catalysts for conversion of butanol to aromatics, Reaction conditions: Temperature = 723 K; WHSV = 0.75
h− 1.
C1–C2 C3–C4 C5–C12 Aromatics BTEX Other higher aromatics Carbon balance (% error)

5Zn-HZSM5_1 bar 8.18 17.68 0.15 74.08 68.21 5.87 2.88


5Zn-HZSM5_5 bar 7.78 26.01 0.48 65.72 57.45 8.27 2.43
5Zn-HZSM5_10 bar 7.13 27.74 0.77 64.36 54.36 10.00 4.16
5Zn-HZSM5_20 bar 7.57 26.05 0.98 65.40 52.51 12.89 14.28
5Zn-HZSM5_30 bar 8.64 22.97 0.74 67.65 52.92 14.74 24.40
HZSM5_20 bar 15.14 30.81 0.47 53.58 37.85 15.73 17.37

4
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Fig. 1. XRD patterns of xZn-HZSM5 (55) (x = 0 to 10 wt%) catalysts.

Fig. 2. FTIR spectra of xZn-HZSM5 (55) (x = 0 to 10 wt%) catalysts in the hydroxyl region.

5
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Fig. 3. UV–vis spectra of xZn-HZSM5 (55) (x = 0 to 10 wt%) catalysts.

Fig. 4. Pyridine FTIR spectra xZn-HZSM5 (55) (x = 0 to 10 wt%) catalysts.

The IR bands at 3660, 3611 and 3437 cm− 1 were assigned to the extra groups in hydroxyl nests, respectively [22,25,47,48]. After impregna­
framework alumina with Lewis acid character, bridged hydroxyl group tion of zinc into HZSM5 (55), the band intensity at 3610 cm− 1 (Brønsted
(Al–OH–Si) with Brønsted acid character and hydrogen bonded Si–OH acid groups) was decreased, which was attributed to the consumption of

6
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

protons by Zn species. It was also observed that the band intensities at catalysts. The presence of Zn species alters the amount and strength of
3660 and 3437 cm− 1 considerably changed after the introduction of Zn medium and weak acid sites. Depending on the Zn loading new acid sites
species [40]. The UV–vis spectra of HZSM5 (55) and Zn-HZSM5 (55) of weak and medium strength were also formed as shown in Fig. 5.
catalysts as shown in Fig. 3 reveal a weak band at ~365 nm for The XPS spectra of selected Zn-HZSM5 (55) catalysts such as 1Zn-
Zn-HZSM5 (55) catalysts and was appeared because of the formation of HZSM5 (55), 5Zn-HZSM5 (55) and 10Zn-HZSM5 (55) were acquired
ZnO crystal on the surface of the catalyst [3]. to understand the distribution of different Zn species (Fig. 6). Pure ZnO
Fig. 4 shows the pyridine FTIR spectra of the HZSM5 (55) and Zn- (2p1/2 and 2p3/2) showed peaks at BE of 1021.9 and 1045.0 eV [45,53].
HZSM5 (55) catalysts. The adsorption of pyridine molecule on Lewis The peak appeared at BE 1022.2 eV for the Zn-HZSM5 (55) catalysts are
and Brønsted acid sites is resulted IR bands at 1450 cm− 1 and 1546 ascribed to -O-Zn-OH species. This species prevails as a portion of zinc
cm− 1, respectively. The synergistic effect between Lewis and Brønsted oxide clusters integrated with the zeolite lattice. However, the size of
acid sites during pyridine adsorption is characterized by the evolution of this zinc oxide cluster is not as large as the clusters of bulk zinc oxide
IR band at 1490 cm− 1. The exchange of zeolite proton with the Zn2+ with BE of 1021.9 eV [53]. A high BE of 1023.0 (Zn 2p3/2) for
species generates ZnOH + species and reveal strong IR band at 1620 1Zn-HZSM5 (55) indicates presence of Zn species situated at the cation
cm− 1 [16,49,50]. Moreover, the IR band intensity at 1620 cm− 1 in­ exchanged sites of zeolite. The hypochromic shift of BE is because of the
creases with the rise in Zn loading suggesting enhancement of the con­ higher electronegativity of the lattice oxygen than the O2− ligand in bulk
centration of ZnOH + species. The IR band intensity at 1546 cm− 1 ZnO [53]. The XPS spectra of 5Zn-HZSM5 (55) and 10Zn-HZSM5 (55)
(Brønsted acid sites) decreases as the Zn loading increased suggesting exhibit higher BE peaks at 1049.5 eV and 1026.5–1027.5eV corre­
increase in exchange of zeolite protons with Zn2+ species (Fig. 4) [51]. sponding to ZnOH+ species (2p1/2 and 2p3/2). Again, the shift of peaks at
In contrast, the amount of Lewis acid sites increases with increasing Zn higher BE indicates strong interaction with the electronegative oxygen
loading. The increase in Lewis acid sites suggests that a fraction of Zn in zeolite framework [45].
metals are present in cationic form [51].
The NH3-TPD profile of xZn-HZSM5 catalysts revealed desorption 3.2. Conversion of butanol to aromatics
peaks in three regions (273–473, 473–673 and above 673 K) corre­
sponding to weak, medium and strong acid sites (Figs. 5 and S5) [40,52]. The conversion of butanol to aromatics (BTA) reactions were per­
The pure HZSM5 showed two NH3 desorption peaks corresponding to formed under wide range of process parameters such as temperature
the weak (446 K) and medium acid sites (625 K). The NH3 desorption (573–723 K), WHSV (0.75–14.99 h− 1) and pressure (1–30 bar) over pure
peaks corresponding to the weak acid site was gradually shifted to HZSM5 (55) and xZn-HZSM5 (x = 1–10%, SiO2/Al2O3 = 55) of varying
higher temperature with the decrease in silica-alumina ratio suggesting Zn loading. The various products formed both in the gas and liquid
increase in strength of the acid sites (Fig. S5). A weak NH3 desorption phases were identified by GC-MS and quantified by gas chromatography
peak was also noticed at 768 K because of the occurrence of strong acid as shown in Fig. S1. The hydrocarbon products identified in the liquid
sites. The incorporation of Zn species on HZSM5 affects greatly the phase were benzene, toluene, xylene isomers, ethyl benzene, propyl
distribution and strength of acid sites. The NH3 desorption peak attrib­ benzene, trimethyl benzenes and higher aromatics (>C9+) and cyclic
uted to the strong acid sites (768 K) gradually shifted to the lower paraffins (C6–C12), and in gas phase were hydrogen, methane, ethane,
temperature as the Zn loading increases. A few new strong acid sites ethylene, propane, propylene, butane isomers, butylene isomers,
were appeared at around 790 K and above for higher Zn loading pentane and pentene isomers. Further these hydrocarbons were

Fig. 5. NH3-TPD profile of xZn-HZSM5 (55) catalysts (x = 0 to 10 wt%).

7
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Fig. 6. XPS spectra of 1Zn-HZSM5 (55), 5Zn-HZSM5 (55) and 10Zn-HZSM5 (55) catalysts.

classified as C1–C2, C3–C4, C5–C12 and aromatics for ease of discussion. parameter H2/(H2+propane + butane) was calculated for pure HZSM5
The variation of selectivities of these products under wide range of (55) and Zn-HZSM5 (55) catalysts. It was observed that H2/
process parameters are discussed in the following sections. (H2+propane + butane) value was increased for Zn-HZSM5 (55) cata­
lysts (Fig. S8). This result clearly indicates that hydride transfer reaction
3.3. Effect of zinc loading was prominent over pure HZSM5 (55) compare to Zn-HZSM5 (55) cat­
alysts under the studied reaction conditions. The hydride transfer re­
The BTA reaction was performed over pure HZSM5 (55) and xZn- action decreases the amount of olefins, which are primary precursor for
HZSM5 (55) (x = 1 to 10 wt%) catalysts at 673 K, WHSV of 0.75 h− 1 the aromatics. The hydride transfer reaction was suppressed in presence
and 1 atm pressure (Fig. 7). The selectivity of total aromatics and BTEX of Zn leading to increase in dehydrogenation of naphthene to produce
was increased with the increase in Zn loading upto 5% and decreased various aromatics. The higher yield of hydrogen in Zn-HZSM5 (55)
thereafter. The maximum selectivities of total aromatics and BTEX ob­ catalysts further support this statement. The nature and strength of acid
tained for 5Zn-HZSM5 (55) were 72% and 62%, respectively. This result sites also influence the various reactions leading to the formation of
indicates enhancement of the aromatization of C3–C4 olefins as the aromatics such as cyclization, oligomerization, hydride transfer reaction
selectivity of C3–C4 olefins follows the reverse trend of aromatics and dehydrogenation of naphthene [25]. The Brønsted acid sites pro­
selectivity. The lower selectivity of total aromatics (~37.5%) over pure mote the cyclization, oligomerization, and hydride transfer reaction of
HZSM5 (55) indicates that presence of Zn enhanced the aromatization olefins, whereas dehydrogenation of naphthene occurs on Lewis acid
(cyclization followed by dehydrogenation) reaction of olefins leading to sites. The strong Brønsted acid sites result in the formation of higher
the formation of higher amount of aromatics in Zn-HZSM5 (55). More­ hydrocarbons through oligomerization reaction, whereas medium
over, the higher selectivity of C3–C4 hydrocarbons for pure HZSM5 (55) strength Brønsted acid sites are appropriate for oligomerization reaction
confirms the enhance rate of aromatization reactions in presence of Zn. to limits the formation of higher aromatic hydrocarbons. The L/B ratio
The selectivity to propane and butane isomers over HZSM5 (55) was of Zn-HZSM5 (55) catalyst increases with the increasing Zn loading
nearly 32% and 20.5%, respectively, which were decreased to 5.25% suggesting the formation of additional Lewis acid sites and decreases the
and 11.8%, respectively over 5Zn-HZSM5 (55) catalyst as shown in Brønsted acid sites (specially higher strength) as discussed earlier. The
Fig. S7. The formation of various aromatics may also proceed by hydride decline of hydride transfer reaction was because of the decrease in
transfer reaction of olefins and naphthene together with cyclization and strong Brønsted acid sites and increase in Lewis acid sites, which
dehydrogenation reaction. In order to shed more light in this regard, a enhance the dehydrogenation of naphthene to aromatics leading to the

8
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Fig. 7. Effect of Zn loading over product distribution, Reaction conditions: Temperature = 673 K; WHSV = 0.75 h− 1; Pressure = 1 atm.

Fig. 8. Effect of Zn loading on individual aromatics selectivity, Reaction conditions: Temperature = 673 K; WHSV = 0.75 h− 1; Pressure = 1 atm.

formation of more hydrogen and less C3–C4 hydrocarbons. Therefore, constant with the increase in Zn loading, whereas selectivities of toluene
the decrease in selectivity of total aromatics above 5 wt% of ZnO loading and C8 aromatics (Xylenes and ethyl benzene) were increased upto 5% of
was because of the decrease in Brønsted acid sites, which were respon­ Zn loading (Fig. 8). Further increase in Zn loading decreases the selec­
sible for the oligomerization reaction followed by cyclization of olefins tivities of toluene and C8 aromatics resulting in decrease in BTEX se­
leading to the formation of napthene, a precursor for the aromatics. lectivities. The formation of C10 and higher aromatics was significantly
In BTEX fraction, the benzene selectivity was virtually remained lower compare to the pure HZSM5 (55). However, selectivity to C9

9
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Fig. 9. Effect of Zn loading on C8 aromatic hydrocarbons selectivity, Reaction conditions: Temperature = 673 K; WHSV = 0.75 h− 1; Pressure = 1 atm.

aromatics was increased with the increase in Zn loading (Fig. 8). The rise which confirms the enhancement of selectivity to C9 aromatics. Thus,
in selectivity to C9 aromatics might be because of the enhance alkylation the products stream comprised of different aromatics including C8 aro­
reaction of C8 aromatics over Lewis acid sites. As discussed earlier, the matics. These result indicate that self and cross oligomerization of
amount of Lewis acid sites enhances with the increase in Zn loading, various alkenes followed by cyclization and dehydrogenation reaction

Fig. 10. The variation of product distribution at different temperatures, Reaction conditions: WHSV = 0.75 h− 1; Pressure = 1 atm; 5Zn-HZSM5 (55) catalyst.

10
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

results in the formation of various aromatics. Formation of different multiple oligomerization and cracking of oligomeric intermediates
lower alkenes was attributed to the cracking reactions over Brønsted leading to the formation of unsaturated C6 oligomer, precursor of ben­
acid sites. Cracking reaction was more prominent over pure HZSM5 (55) zene. The unsaturated C6 oligomers further transform to benzene via
compared to Zn-HZSM5 (55) catalysts because of the higher Brønsted aromatization reaction. However, the selectivities of toluene and xy­
acidity of HZSM5 (55) as shown in Fig. 4. Total higher aromatics (C9+) lenes were varied with the Zn loading, which are well correlated with
selectivity was higher over HZSM5 than Zn-HZSM5 catalysts, and spe­ the published literature [3]. The toluene and xylenes selectivities over
cifically, C9 aromatics selectivity was increased on Zn-HZSM5 (55) HZSM5 (55) were about 12.6% and 8.75% and were increased to nearly
catalysts. The selectivity for C9+ hydrocarbons was higher over HZSM5 29% and 26.6%, respectively over 5Zn-HZSM5 (55). Further increase in
(55) because of the higher external surface area and pore volume. From Zn loading slightly decreases the selectivities of toluene and xylenes. The
BET analysis, external surface area, micropore area and pore volume of slight decrease in selectivities of toluene and xylenes might be because
the HZSM5 (55) were 195 m2/g, 162 m2/g and 0.26 cm3/g, respectively. of the decrease in Brønsted acid sites (Fig. S8) for higher Zn loading
External active sites promote the formation of higher hydrocarbons, catalyst.
whereas internal active sites favor the formation of light aromatics of
matching kinetic diameter with pores dimension. HZSM5 (55) has more
3.4. Effect of temperature
external active sites, which promote the formation of large molecular
size aromatics. In contrast, the external surface area, micropore area and
The influence of temperature on selectivity to various aromatics was
pore volume were decreased to 109 m2/g (29% reduction), 200 m2/g
investigated in the wide temperature range of 573–773 K and WHSV of
(15.6% reduction) and 0.24 cm3/g, respectively for the 5Zn-HZSM5 (55)
0.75 h− 1 using 5Zn-HZSM5 (55) catalyst under atmospheric pressure.
catalyst. The external surface area was affected more than the micropore
The selectivity of total aromatics was enhanced with the increasing
area. Hence, the 5Zn-HZSM5 (55) catalyst favors the formation of more
temperature from 573 K to 748 K, and was decreased on further increase
lighter aromatics (BTEX) [54,55]. The methane and ethane selectivities
in temperature to 773 K. The selectivities to C5–C12 range hydrocarbons
were increased over 5Zn-HZSM5 (55) than HZSM5 (55) catalyst because
and C3–C4 hydrocarbons were decreased as the temperature was
of the enhancement of the interconversion of the different aromatics by
increased in the studied temperature range. The selectivities to BTEX
hydrodealkylation/transalkylation and disproportionation reactions
and total aromatics were nearly 20.55% and 33.28%, respectively,
(Fig. S6) [56,57].
whereas the selectivity for C3–C4 and C5–C12 hydrocarbons at 573 K
The variation of selectivities of C8 aromatics fraction consisting of
were 50.70% and 15.27%, respectively. As the temperature increased to
ethylbenzene, p-xylene, o-xylene, and m-xylene with Zn loading is
773 K, the selectivities to C3–C4 and C5–C12 hydrocarbons were gradu­
shown in Fig. 9. The deposition of Zn and its amount have insignificant
ally decreased to 10% and 0%, respectively as shown in Fig. 10. The
influence on the selectivities of benzene and m-xylene. During BTA re­
maximum selectivities to BTEX (68.75%) and total aromatics (74.83%)
action, the benzene was formed through multiple oligomerizations of
were achieved at 723 K. The selectivity to higher aromatics was initially
olefins and cracking of oligomeric intermediates [3]. The multiple
increased to 17.48% at 598 K and then was gradually decreased to
oligomerization and cracking of oligomeric intermediates occur over
4.41% at 773 K because of the enhance cracking reaction at high tem­
strong Brønsted acid sites. Deposition of small quantity of ZnO reduces
perature (Fig. 11). The gradual increase in selectivities to C1–C2 hy­
the concentration of strong Brønsted acid sites, which suppress the
drocarbons indicates the increase of cracking of higher aromatics side

Fig. 11. The variation of aromatics product distribution at different temperatures, Reaction conditions: WHSV = 0.75 h− 1; Pressure = 1 atm; 5Zn-HZSM5
(55) catalyst.

11
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Fig. 12. The influence of temperature on BTEX selectivity, Reaction conditions: WHSV = 0.75 h− 1; Pressure = 1 atm; 5Zn-HZSM5 (55) catalyst.

chain with the increasing temperature. The selectivity to C3–C4 hydro­ can be concluded that dehydrogenation of naphthene was the primary
carbons was declined with the increase in temperature. The decrease in reaction for the formation of aromatics at high temperature. However,
selectivity of C4 paraffins in C3–C4 fraction might be because of the the selectivity to propane was approximately remained constant over
decrease in hydride transfer reaction at high temperature (Fig. S9). It the studied temperature range.

Fig. 13. Effect of WHSV on the product distribution, Reaction conditions: Temperature = 723 K; Pressure = 1 atm; 5Zn-HZSM5 (55) catalyst.

12
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

The selectivities to xylenes were increased from 11.33% to nearly 3.5. Effect of WHSV
26.63% as the temperature increased from 573 K to 673 K. Further in­
crease in the temperature, the xylenes selectivities were decreased to Effect of WHSV on selectivities of BTEX and aromatics using 5Zn-
17.07% at 773 K (Fig. 12). The decline in xylenes selectivities beyond HZSM5 (55) catalyst was studied with different WHSV of 14.96 h− 1,
673 K was attributed to the hydrodealkylation of xylenes to toluene and 7.48 h− 1, 2.99 h− 1, 1.50 h− 1, 0.75 h− 1 and 0.50 h− 1 at 723 K and 1 atm
methane. Moreover, the benzene selectivity was gradually increased pressure. The selectivity to total aromatics was increased from 29.19%
with the increasing temperature because of the enhanced hydro­ to 74% as the WHSV decreased to 0.50 h− 1. Similarly, the selectivities to
dealkylation of toluene to benzene and methane (Fig. 12). BTEX was also increased from 25.88% to 66.83% within the WHSV
The toluene selectivity was increased from 6.8% to 32.67%, as the range of 14.96 h− 1 to 0.5 h− 1. The maximum BTEX selectivities of 68%
temperature increased from 573 K to 723 K and thereafter the toluene was achieved at WHSV of 0.75 h− 1 (Fig. 13). The C5–C12 and C3–C4
selectivity was decreased to 30.58% at 773 K. Above 723 K, the hydrocarbons selectivities were decreased from 12% and 54.25% to
hydrodealkylation reaction of toluene to benzene and methane was 0.14% and 17.12%, respectively. However, C1–C2 selectivity was
dominated resulting in decrease in selectivity. Hence, the hydro­ increased from 4.47% to 8.71% as the WHSV decreased from 14.96 h− 1
dealkylation of xylenes and toluene increases the selectivity to benzene to 0.5 h− 1 as shown in Fig. 13. It was clearly evident that low WHSV
as the temperature raised from 573 K to 773 K (Fig. 12). The selectivity favors the formation of more aromatics and subsequently suppresses the
for BTEX and total aromatics remained nearly constant upto 773 K. formation of C3–C4 and C5–C12 hydrocarbons. Prolong exposure (higher
However, the variation of individual selectivity of benzene, toluene and contact time) of olefins to acid sites leads to successive oligomerization,
xylenes was attributed to the hydrodealkylation of toluene, xylenes to cyclization and dehydrogenation reactions resulting in the formation of
form benzene and methane and the hydrodealkylation of C9 aromatics various aromatics. The other higher aromatics cumulative selectivity
(Ethyl toluene, trimethyl benzene) could be the another reason for in­ was also increased from 3.32% to 7.18% with the decrease in WHSV
crease in selectivity of BTEX as shown in Fig. S10. For example, C9 ar­ from 14.96 h− 1 to 0.5 h− 1 (Fig. 14). More specifically, all the individual
omatics, ethyl toluene might have undergone hydrodealkylation higher aromatics selectivities were enhanced with the decrease in
reaction to produce toluene and ethane at higher temperatures. Due to WHSV, which was attributed to the oligomerization of various olefins
these pronounced hydrodealkylation reactions various aromatics were followed by aromatization (Fig. S12). The selectivities for methane and
formed at higher temperatures leading to the increase in selectivities for ethane at WHSV of 14.96 h− 1 were 0.45% and 0.15%, respectively and
methane and ethane. From Fig. S11, the selectivities to methane and were increased to 4.18% and 3.56%, respectively for the WHSV of 0.75
ethane were increased to 8.62% and 8.83%, respectively at 773 K from h− 1 as shown in Fig. S13. This result might be due to the hydro­
0.13% to 0.08%, respectively at 573 K. The sharp rise in selectivity of dealkylation of higher aromatics and interconversion of the light aro­
methane and ethane above 673 K was attributed to severe hydro­ matics on the strong acid sites at higher contact time [58].
dealkylation reactions of various alkyl substituted aromatics. The Formation of butylenes and propylene were observed to be signifi­
interconversion of various aromatics might also proceed by dispropor­ cant at high WHSV and declined with the decrease of WHSV [22]. The
tionation and transalkylation reactions. However, the hydro­ propylene was formed by dehydroformylation of butanal, which was
dealkylation might be more pronounced than disproportionation and formed by dehydrogenation of n-butanol. The dehydrogenation of
transalkylation reactions over the 5Zn-HZSM5 (55) catalyst in the n-butanol and dehydroformylation of butanal were favorable route for
studied temperature range [57–59]. n-butanol transformation at lower temperature. However, these re­
actions may contribute significantly to n-butanol transformation at high

Fig. 14. Effect of WHSV on the aromatics product distribution, Reaction conditions: Temperature = 723 K; Pressure = 1 atm; 5Zn-HZSM5 (55) catalyst.

13
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Fig. 15. Effect of WHSV on BTEX hydrocarbons selectivity, Reaction conditions: Temperature = 723 K; Pressure = 1 atm; 5Zn-HZSM5 (55) catalyst.

Fig. 16. Effect of pressure on the aromatic product distribution, Reaction conditions: Temperature = 723 K; WHSV = 0.75 h− 1; 5Zn-HZSM5 (55) catalyst.

temperature resulting in higher selectivity of propylene. The selectiv­ respectively at WHSV of 0.5 h− 1 (Fig. S14). In contrast, the selectivities
ities for propylene and butylenes at WHSV of 14.96 h− 1 were 18.31% of light aromatics, benzene, toluene and xylenes were increased from
and 29.72%, respectively and was decreased to nearly 1% and 1.36%, 2.19%, 11.93% and 11.20% to 12.4%, 30.36%, and 17.30%,

14
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

respectively as WHSV decrease from 14.96 h− 1 to 0.5 h− 1 (Fig. 15). n-butanol. Depending on the reaction conditions (temperature, pressure
These results clearly indicates that the oligomerization of olefins and and WHSV) and strength of acid sites, the hydrocarbon species either
aromatization reactions are favorable at relative low WHSV meaning desorb from the catalyst surface or further converted to higher
high contact time. substituted aromatics and polynuclear aromatics, which may deposit
Another interesting observation was that the selectivity of propane inside the porous channel and or external surface of the catalyst. For
was increased with the decrease of WHSV (Fig. S14). This result in­ example, the low reaction temperature favors the formation of oligo­
dicates that hydride transfer reaction favors at lower WHSV. The mers and polymers as primary carbonaceous species. However, high
decreasing trend of values of H2/(H2+C3+C4) further confirms the reaction temperature and pressure promotes the formation of poly­
increasing contribution of hydride transfer reaction as the WHSV is aromatics (naphthalene) and highly substituted aromatics (trimethyl
decreased. Hence, the selectivity for butylenes isomers and propylene benzene, ethyl toluene etc.). These higher aromatics were formed over
was decreased and the aromatics selectivity was increased with the acid sites through various complex reactions, which include condensa­
decrease in WHSV (Fig. S15). tion, rearrangement, hydrogen transfer reaction and dehydrogenation
reactions. Thus, enhanced selectivities of these higher aromatics at
3.6. Effect of pressure higher pressures were attributed to the condensation and rearrangement
reactions [56,60]. Furthermore, a comparison on formation of higher
The influence of pressure on total aromatics and BTEX selectivities aromatics (C9–C13) over HZSM5 (55) and 5Zn-HZSM (55) was made as
were studied at different pressures (1, 5, 10, 20 and 30 bar), 723 K, and shown in Fig. S18. In case of 5Zn-HZSM5 (55), the selectivity of all the
WHSV of 0.75 h− 1. The selectivity for BTEX was decreased as the aromatics increases with the increase in reaction pressure (Fig. S22). The
pressure increased from 1 bar to 30 bar as shown in Fig. S16. In contrast, selectivity of C9 aromatics at 20 bar was higher for 5Zn-HZSM5 (55)
the selectivity to total aromatics was decreased as the pressure increased than the pure HZSM5 (55). However, the selectivities of C10–C13 were
upto 10 bar and was marginally increased with the increasing pressure higher for HZSM5 (55) than the 5Zn-HZSM5 (55). These results indicate
beyond 10 bar. The marginal increase in the selectivity to total aro­ pure HZSM5 (55) favors the formation of higher aromatics. From
matics was due to the formation of higher aromatics. The selectivity for Fig. 17, The selectivity of light aromatics benzene, toluene and xylenes
higher aromatics at 1 bar was 5.87% and was increased to 14.74% at 30 were 11.93%, 32.26% and 23.59% at 1 bar pressure and was decreased
bar pressure (Fig. 16). to 8.78%, 23.89% and 19.04%, respectively at 20 bar pressure. How­
These higher aromatics were primarily consisted of ethyl toluene, ever, no significant change in selectivities of these aromatics was noticed
trimethyl benzenes and naphthalenes (naphthalene, methyl, ethyl, at 30 bar pressure.
propyl - naphthalenes). The selectivity of these aromatics at 1 bar was The variation of selectivities of BTEX was monitored with time on
3.22%, 1.04% and 1.12%, respectively and was increased to 5.18%, stream (TOS) at different pressure using 5Zn-HZSM5 (55) (Fig. 18). At
1.65% and 4.75%, respectively at 30 bar pressure (Fig. S17). These, ambient pressure, the selectivities of BTEX were increased upto first 2 h
higher aromatics were formed on the external surface of the 5Zn-HZSM5 of TOS and remained constant at higher TOS. In contrary, the selectiv­
(55) catalyst. The formation of various carbonaceous deposits on cata­ ities to BTEX were continued to decrease with the increase in TOS at 10
lyst was explained by hydrocarbon pool model [60]. This model suggests bar and above pressure. However, a small fluctuation in selectivities was
the involvement of organic active species and inorganic active sites (acid observed at relatively lower pressure of 5 bar. The high pressure favors
sites and metal center) during the formation of hydrocarbon pool from the formation of higher aromatics of large kinetic diameter, which were

Fig. 17. Effect of pressure on BTEX hydrocarbons selectivity, Reaction conditions: Temperature = 723 K; WHSV = 0.75 h− 1; 5Zn-HZSM5 (55) catalyst.

15
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Fig. 18. Effect of pressure on BTEX selectivity over 6 h TOS, Reaction conditions: Temperature = 723 K; WHSV = 0.75 h− 1; 5Zn-HZSM5 (55) catalyst.

deposited on the external surface of the HZSM5 (55) causing catalyst methane and ethane was insignificant with pressure indicating minor or
deactivation by blockage of pores. The light aromatics, BTEX of smaller no effect on hydrodealkylation intermediate reaction. The propane
kinetic diameters were selectively formed inside the porous channel of selectivity was found to be increased upto 10 bar pressure and was
HZSM5 (55). The gradual decrease in selectivities of BTEX (upto 20 bar) decreased with the further increase in pressure. The hydride transfer
further confirms the limited access of pores to light aromatics. The reaction for the formation of aromatics favors at high pressure over
decline in BTEX selectivities was more prominent at higher pressures, dehydrogenation reactions leading to the formation of various aro­
particularly at 10 bar, 20 bar and 30 bar over 5Zn-HZSM5 (55) catalyst matics. However, alkylation reactions dominates at high pressure
after 1h of TOS. resulting in decrease of propane selectivity [57]. Butanes selectivity was
The BTEX selectivities were 63.99%, 62.89% and 62.44% at 1 h also decreased at higher pressure (Fig. S22). The selectivity for C3–C4
decreased to 54.35%, 50.56% and 48.05% after 6 h TOS, at 10 bar, 20 hydrocarbons was increased with the increase in pressure upto 10 bar
bar and 30 bar, respectively. The other higher aromatics (>C8) selec­ and decreased with a further increment in the pressure. In summary, the
tivity after 1 h of TOS were 12.34%, 15.90% and 22.45% at 10 bar, 20 maximum selectivities of total aromatics and BTEX achieved were
bar and 30 bar, respectively and were decreased to nearly 10%, 12.29% 74.83% and 68.75% (B: 11.10%, T: 32.67%, E: 0.54%, X: 24.44) at 723
and 13.79% respectively at 6 h TOS (Fig. S19). There were two impor­ K, 1 bar and 0.75 h− 1 WHSV for 5Zn-HZSM5 (55) catalyst.
tant findings clearly evident from here. First, the formation of higher
aromatics (>C8) at the beginning of the BTA reaction was higher due to 3.7. Effect of silica-alumina ratio
higher availability of active sites and later decreased gradually as the
reaction progressed. Second was the enhancement of the formation of The BTA reaction studied over 5Zn-HZSM5 catalysts of different
higher aromatics (>C8) at the beginning as the reaction pressure in­ silica-alumina ratio (23, 55 and 280) was studied under optimized re­
creases. Thus, higher aromatics formed at the beginning might be action conditions as mentioned above (Fig. S23). The highest selectivity
strongly adsorbed on the active sites for a longer time or trapped in the of total aromatics and BTEX was achieved for 5Zn-HZSM5 (55). The
catalyst pores resulting in catalyst deactivation and decrease in the lower selectivity total aromatics and BTEX over 5Zn-HZSM5 (23) was
product selectivity. The carbon and material balance errors were because of the presence of strong acid sites (Fig. S5) which promotes
increased particularly at higher pressure reactions (Table 2) and was formation of higher hydrocarbons leading to deactivation of the catalyst
attributed to the deposition of higher aromatics. [25]. The HZSM5 (23) is associated with lower external surface area
At higher pressure, the internal conversion of aromatics to heavy than the other zeolites in the present investigation [25]. It was
aromatics could be possible as mentioned earlier [57]. The higher aro­ demonstrated before that ZnO deposits preferencially on external sur­
matics (>C8) selectivity over pure HZSM5 (55) after 1h of TOS (25.86%) face area. Thus, deposition of ZnO in micropores of HZSM5 (23) is
was much higher than over 5Zn-HZSM5 (55) (12.29%) at 20 bar pres­ inevitable because of low external surface area, which may cause
sure because of the higher Brønsted acidity, external surface area in pure blockage of pores. The pore blockage phenomena may also contribute
HZSM5 (Figs. S20a and b). The higher aromatics of large kinetic diam­ towards poor performance of 5Zn-HZSM5 (23) catalyst. The signifi­
eter were readily formed on the external surface [54,55,57,61]. How­ cantly lower performance of 5Zn-HZSM5 (280) catalyst was due to the
ever, the individual selectivities of BTEX and other aromatics were insufficient concentration of Brønsted acid sites. In addition, the higher
decreased with the increasing TOS, due to the deposition of higher ar­ crystallite size and poor crystallinity might be another reason for poor
omatics. From Fig. S21, it was evident that the variation of selectivity of performance of this catalyst (Table S2).

16
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Fig. 19. TGA analysis of spent HZSM5 (55) and 5Zn-HZSM5 (55) catalysts, Reaction conditions: Temperature = 723 K; WHSV = 0.75 h− 1.

3.8. Coke analysis (Fig. 19). The derivative thermogravimetry (DTG) curves of the spent
catalysts obtained at different reaction pressures were generated to
The amount of carbonaceous species (coke) deposited in the spent know the temperature for decomposition/removal of different types of
catalysts (5Zn-HZSM5 (55) and HZSM5 (55)) was determined by ther­ coke as shown in Fig. S24.
mogravimetric analysis (TGA) as shown in Fig. 19. From Fig. 19, the The weight losses occurred during TGA analysis were at four
total weight loss was increased with the increase in reaction pressure different temperature region as reported in the literature: 293–438,
from 1 bar to 30 bar for 5Zn-HZSM5 (55) catalysts. The total weight 438–531, 531–673, and 673–873 K [60]. The first stage (<438 K) loss of
losses observed were 7.19, 8.01, 8.07, 9.71 and 9.72 wt% at 1, 5, 10, 20 weight corresponds to the removal of moisture and any other physically
and 30 bar pressure, respectively. The total weight loss of 9.60 wt% was adsorbed species. The weight loss occurred in the moderate tempera­
obtained for spent HZSM5 (55) (20 bar), which was fairly close to the tures (438–673 K) is because of the removal of volatile species named as
weight loss obtained for spent 5Zn-HZSM5 (55) catalysts at 20 bar and “soft coke”, whereas the weight loss occurred between 673 and 873 K is
30 bar pressures. The thermogravimetry curve shows stage wise loss of because of the matured coke called “hard coke”. The soft cokes are
weight indicating deposition of different types of coke on the catalysts aliphatic in nature and soluble in organic solvents. Whereas hard coke

Fig. 20. Spent HZSM5 (55) and 5Zn-HZSM5 (55) catalysts, A) 5Zn-HZSM5 (55)_1 bar, B) 5Zn-HZSM5 (55)_5 bar, C) 5Zn-HZSM5 (55)_10 bar, D) 5Zn-HZSM5 (55)_20
bar, E) 5Zn-HZSM5 (55)_30 bar, and F) HZSM5 (55)_20 bar, Reaction conditions: Temperature = 723 K; WHSV = 0.75 h− 1.

17
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Table 3
Textural properties of spent HZSM5 (55) and 5Zn-HZSM5 (55) catalysts.
Catalyst SABETa (m2/g) AEb (m2/g) VTc (cm3/g) Vmd (cm.3/g) Ame (m2/g)

Fresh 5Zn-HZSM5 309 110 0.24 0.10 199


5Zn-HZSM5_1 bar 252 95 0.23 0.08 157
5Zn- HZSM5_5 bar 240 87 0.22 0.08 153
5Zn- HZSM5_10 bar 226 88 0.21 0.07 138
5Zn- HZSM5_20 bar 210 79 0.21 0.07 131
5Zn- HZSM5_30 bar 192 69 0.19 0.06 123
HZSM5_20 bar 223 53 0.21 0.07 170
HZSM5 (Fresh) 373 195 0.29 0.12 162
a
BET surface area.

b
External surface area derived from t-Plot method.

c
Total pore volume derived from BJH method.

d
Micropore volume derived from t-Plot method.

e
Micropore surface area derived from t-Plot method.

are primarily of polyaromatic in nature and insoluble in organic sol­ aromatics (Fig. S27). The ratio of internal to external coke over spent
vents. The weight losses observed in different temperature ranges for 5Zn-HZSM5 (55) at 1, 5, 10, 20, 30 bar and spent HZSM5 (55) at 20 bar
spent HZSM5 (55) catalyst (20 bar) follow the trend: 293–438 < were 0.53, 0.45, 0.87, 0.61, 0.99 and 0.15 respectively.
<438–673 < 673–873. In contrast, for 5Zn-HZSM5 (55) (20 bar) the In addition, the spent catalysts were characterized by FTIR and
weight losses observed were more in the temperature range of 438–673 UV–vis spectroscopy. The IR spectra in the wavenumber region of
K and significantly less in the temperature region of 673–873 K. At 1300–1700 cm− 1 and 2800-3100 cm− 1 are important to characterize the
ambient pressure, both the soft and hard coke were formed over coke in spent catalysts (Figs. S28a and b). In the first region, the IR band
5Zn-HZSM5 (55) catalyst. As the reaction pressure increased to 30 bar appeared at 1390 cm− 1 is assigned to the terminal –CH3 groups, whereas
(for 5Zn-HZSM5 (55)), the more weight loss in the medium temperature IR band at 1460 cm− 1 is attributed to the combination of aliphatics
range corresponding to soft coke was observed and the weight loss peak (bending and scissoring) and alkyl aromatics [64]. Thus, the assignment
corresponding to hard coke shifted from 673 K to 773 K. This result of the later IR band is not free from ambiguity. The IR bands appeared at
indicates the formation of more polynuclear aromatic species as hard 1620, 1515, 1400 and 1370 cm− 1 are assigned to the νC = C vibrations
coke at higher pressure. High external surface area and high Brønsted of unsaturated hydrocarbons, νC-C, νC-C in ring aromatic vibration,
acidity in case of HZSM5 (55), enhance the formation of more hard coke δaCH3 and δsCH3 of aliphatic hydrocarbon. In the second region, the IR
at high pressure as mentioned earlier. These results indicate that HZSM5 band at 2870 cm− 1 corresponds to CH3 groups, and the IR bands ~2900
(55) promotes the formation of hard coke, whereas 5Zn-HZSM5 pro­ and 2935 cm− 1 are assigned to – CH2 and –CH3 groups [65,66]. The IR
motes the formation of soft coke (volatiles) at 20 bar reaction pressure, band appeared at 2960 cm− 1 is due to the –CH3 groups [64]. As observed
723 K temperature and 0.75 h− 1 WHSV. The coke deposited on spent from the figures, the intensity of the IR band for aliphatic and alkyl
HZSM5 (55) and 5Zn-HZSM5 (55) catalyst was extracted using the aromatics at 1460 cm− 1 is increased with the increase in reaction
procedure described by Zhang et al. [60]. The details of the coke pressure from 1 bar to 30 bar. So, it is clearly evident that at higher
extraction procedure are available in the supplementary information. As pressures more aliphatic and alkyl aromatics hydrocarbons were
shown in Fig. 20, higher amount of hard coke was deposited on spent deposited on the 5Zn-HZSM5 (55) catalysts. The UV–vis spectra of spent
HZSM5 as deep black color solution together with higher amount of catalysts at different reaction pressure are shown in Fig. S30. All the
insoluble black residue was obtained compared to spent 5ZnHZSM5 (55) spent catalyst UV–vis spectra revealed broad band suggesting hetero­
catalyst. The CH2Cl2 organic solvent extracted phase was analyzed by geneity of the coke. The band appeared at 365–375 nm is attributed to
GC for both the HZSM5 (55) and 5Zn-HZSM5 (55) catalysts (Figs. S26a the conjugated double bond and poly condensed aromatics with multi­
and b). The alkyl-substituted aromatics and polynuclear aromatics ple aromatic rings [64]. The appearance of band at 560 and 700 nm
(naphthalene and alkyl naphthalenes) were formed on both the cata­ further confirms the formation of polycondensed aromatics. The in­
lysts. However, relative amount of alkyl substituted aromatics and tensity of bands characteristics of polycondensed aromatics increases
polynuclear aromatics were higher in spent HZSM5 (55) than with the increase in reaction pressure confirming the formation of pol­
5Zn-HZSM5 (55). ycondensed aromatics at higher reaction pressure.
Moreover, pore size and surface area measurement revealed a
decrease in the micropore volume and external surface area of spent 3.9. Reaction mechanism
catalysts with an increase in reaction pressure (Table 3) suggesting
deposition of carbonaceous species. The drastic drop in the external The BTA reactions was performed over xZn-HZSM5 catalysts under
surface area was observed for spent HZSM5 (55) (20 bar), which was wide range of process parameters. Various aliphatic hydrocarbons (C1 to
because of the deposition of higher aromatics on the external surface C12) and aromatic hydrocarbons (C6 to C12) were formed as products
compared to Zn-HZSM5 (55) spent catalysts. The internal (inside pores) under the studied reaction conditions. Water and hydrogen was also
and external (external surface) cokes were estimated based on the formed. However, CO was not observed in the gaseous products mixture.
method reported in the literature [62–64] and details are also available The absence of CO might be because of its rapid transformation to
as supplementary information. A small increase in amount of internal methane, which was identified as gaseous product. Based on the product
coke than external coke at higher pressure attributed to the formation of distribution obtained under various reaction conditions, a plausible re­
higher molecular aromatics on the surface of the catalyst as discussed action mechanism was proposed for the direct conversion of n-butanol
earlier. The high amount of external coke observed for spent HZSM5 to various aromatics over xZn-HZSM5 catalysts as shown in Scheme 1.
(55) at 20 bar reaction pressure was because of the formation of heavy The direct conversion of n-butanol to various aromatics was

18
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

Scheme 1. Reaction pathway for the aromatization of n-butanol to aromatics over Zn-HZSM5 catalysts.

proceeded via the formation of olefins (butylenes and propylene), which increase in Zn loading. In case of Zn-HZSM5 (55) catalysts, the dehy­
were subsequently transformed to different aromatics following various drogenation of naphthene was dominant reaction pathway for the for­
reactions as discussed below. n-butanol underwent direct dehydration mation of aromatics. Thus, the suppression of hydride transfer reaction
reaction and or via the formation of di-butyl ether to produce butylenes and enhancement of dehydrogenation of naphthene in presence of Zn
over acid sites [22,25]. However, contribution of these dehydration were the key factors for the higher selectivity of BTEX. The distribution
pathways was not discernible. Nonetheless, the di-butyl-ether was not of aromatics, specifically BTEX was also greatly influenced by various
identified in the present investigation as this product is generally formed reaction parameters. Isomerization, disproportionation, transakylation
at lower temperature (below 550 K). The butylenes including all of its and hydrodealkylation reactions were important for their inter trans­
isomers were formed via isomerization reaction over acid sites. In formation as shown in Scheme 1.
another route, the n-butanol was first transformed to butanal via
dehydrogenation reaction, which was then transformed to propylene 4. Conclusions
and propane via dehydroformylation and decarbonylation reaction,
respectively. The dehydrogenation of n-butanol proceeds over metallic The direct conversion of n-butanol to aromatics with enhanced
center of the Zn-HZSM5 catalysts and to some extent over extra frame­ selectivity to BTEX was performed over Zn impregnated HZSM5 (SiO2/
work alumina of zeolites [22,25]. The various olefins formed were oli­ Al2O3 = 23, 55 and 280) catalysts under different process parameters.
gomerized to form C5–C12 olefins over acidic sites. The ethylene was The incorporation of Zn in HZSM5 has a great influence on the textural
formed because of the cracking of alkyl chains of the aromatics at high properties and, nature and strength of acidity of Zn-HZSM5 catalysts.
reaction temperature. The self metathesis of 1-butylene or propylene Deposition of Zn decreases the external surface area significantly than
was another possibility for the formation of ethylene [22]. the micropore area. The NH3-TPD studies revealed the generation of
The hydrogenation of various olefins (C2–C12) led to the formation of medium strength acidity on deposition of Zn in HZSM5. The concen­
corresponding paraffins, C2–C12. The C6–C12 olefins transformed to cy­ tration of Brønsted acid sites decreases, whereas Lewis acid sites in­
clic paraffins (naphthene) via cyclization reaction and dehydrogenation creases as the Zn loading increases. The incorporation of Zn in HZSM5
of these cyclic paraffins led to the formation of aromatics. Moreover, the has strong influence on the selectivity of total aromatics and BTEX. The
hydride transfer reaction of C3–C4 and cyclic paraffins (naphthene) was selectivity of total aromatics and BTEX increases with the increase in Zn
another pathway for the formation of C3–C4 paraffins and aromatics. loading till 5 wt% and slightly decreases thereafter. The presence of Zn
The contribution of hydride transfer reaction was declined with the species suppresses the hydride transfer reaction and promotes the

19
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

dehydrogenation reaction of naphthene leading to the enhancement of production of renewable p -xylene, ACS Catal. 2 (2012) 935–939, https://doi.org/
10.1021/cs300011a.
the selectivity of aromatics and BTEX. Moreover, the enhancement of
[8] S. Fegade, B. Tande, A. Kuba, W. Seames, E. Kozliak, Novel two-step process for the
Lewis acidity increases the alkylation of C8 aromatics resulting in the production of renewable aromatic hydrocarbons from triacylglycerides, Ind. Eng.
enhancement of C9 aromatic selectivity. Specifically, in BTEX fraction, Chem. Res. (2015), https://doi.org/10.1021/acs.iecr.5b01932.
the variation of benzene and ethyl benzene selectivities were insignifi­ [9] A. Bhan, W.N. Delgass, Propane aromatization over HZSM-5 and Ga/HZSM-5
catalysts, Catal. Rev. Sci. Eng. 50 (2008) 19–151, https://doi.org/10.1080/
cant, whereas the selectivity of toluene and xylenes selectivities was 01614940701804745.
increased significantly upto 5 wt% of Zn loading. The higher selectivity [10] G. Caeiro, R.H. Carvalho, X. Wang, M.A.N.D.A. Lemos, F. Lemos, M. Guisnet,
of C8 aromatics might be because of the enhanced alkylation of F. Ramôa Ribeiro, Activation of C2–C4 alkanes over acid and bifunctional zeolite
catalysts, J. Mol. Catal. A Chem. 255 (2006) 131–158, https://doi.org/10.1016/j.
respective lighter aromatics. The reaction parameters have also signifi­ molcata.2006.03.068.
cant effect on the aromatics product distribution. Low WHSV favors the [11] H. Borwitzky, G. Schomburg, Separation and identification of polynuclear aromatic
aromatization reaction resulting in higher selectivity of total aromatics compounds in coal tar by using glass capillary chromatography including
combined gas chromatography-mass spectrometry, J. Chromatogr. A 170 (1979)
and BTEX. Moreover, low WHSV promotes hydride transfer reaction for 99–124, https://doi.org/10.1016/S0021-9673(00)84242-9.
the production of aromatics. The selectivity of aromatics and BTEX in­ [12] J.H. Lunsford, Catalytic conversion of methane to more useful chemicals and fuels:
crease with the rise in temperature till 748K. Hydrodealkylation was a challenge for the 21st century, Catal. Today 63 (2000) 165–174, https://doi.org/
10.1016/S0920-5861(00)00456-9.
favorable reaction for the production of light aromatics at high tem­ [13] S.K. Maity, Opportunities, recent trends and challenges of integrated biorefinery:
perature. However, the reaction pressure has adverse effect. The selec­ Part I, Renew. Sustain. Energy Rev. 43 (2015) 1427–1445, https://doi.org/
tivities of total aromatics and BTEX decreases with the increase in 10.1016/j.rser.2014.08.075.
[14] S.K. Maity, Opportunities, recent trends and challenges of integrated biorefinery:
reaction pressure. The high reaction pressure results in the formation of
Part II, Renew. Sustain. Energy Rev. 43 (2015) 1446–1466, https://doi.org/
higher aromatics causing deactivation of the catalysts. The maximum 10.1016/j.rser.2014.08.075.
selectivities of total aromatics and BTEX achieved were 74.83% and [15] M. Inaba, K. Murata, M. Saito, I. Takahara, Ethanol conversion to aromatic
68.75% (B: 11.10%, T: 32.67%, E: 0.54%, X: 24.44) at 723 K, 1 bar and hydrocarbons over several zeolite catalysts, React. Kinet. Catal. Lett. 88 (2006)
135–142, https://doi.org/10.1556/RKCL.88.2006.1.18.
o.75 h− 1 WHSV for 5Zn-HZSM5 (55) catalyst. Detail coke analysis [16] K. Ji, J. Xun, P. Liu, Q. Song, J. Gao, K. Zhang, J. Li, The study of methanol
revealed that high reaction pressure led to the formation of coke of aromatization on transition metal modified ZSM-5 catalyst, Chin. J. Chem. Eng. 26
polynuclear aromatic in nature. The alkyl substitute aromatics and (2018) 1949–1953, https://doi.org/10.1016/j.cjche.2018.03.024.
[17] C.M. Lok, J. Van Doorn, G. Aranda Almansa, Promoted ZSM-5 catalysts for the
polynuclear aromatics were formed predominately on pure HZSM5 (55) production of bio-aromatics, a review, Renew. Sustain. Energy Rev. 113 (2019)
than 5Zn-HZSM5 (55). A plausible reaction mechanism was proposed 109248, https://doi.org/10.1016/j.rser.2019.109248.
based on the product distribution obtained at different reaction [18] K.K. Ramasamy, Y. Wang, Catalyst activity comparison of alcohols over zeolites,
J. Energy Chem. 22 (2013) 65–71, https://doi.org/10.1016/S2095-4956(13)
conditions. 60008-X.
[19] W.R. da S. Trindade, R.G. dos Santos, Review on the characteristics of butanol, its
CRediT authorship contribution statement production and use as fuel in internal combustion engines, Renew. Sustain. Energy
Rev. 69 (2017) 642–651, https://doi.org/10.1016/j.rser.2016.11.213.
[20] M. Mascal, Chemicals from biobutanol: technologies and markets, Biofuels
Tatinaidu Kella: Investigation, Formal analysis, Writing – original Bioprod. Biorefin. 6 (2012) 483–493, https://doi.org/10.1002/bbb.1328.
draft, preparation. Debaprasad Shee: Supervision, Conceptualization, [21] J. Mora Vargas, L.H. Tofaneli Morelato, J. Orduna Ortega, M. Boscolo, G. Metzker,
Upgrading 1-butanol to unsaturated, carbonyl and aromatic compounds: a new
Resources, Writing – review & editing. synthesis approach to produce important organic building blocks, Green Chem. 22
(2020) 2365–2369, https://doi.org/10.1039/d0gc00254b.
[22] V.C.S. Palla, D. Shee, S.K. Maity, Conversion of n-butanol to gasoline range
Declaration of competing interest hydrocarbons, butylenes and aromatics, Appl. Catal. A Gen. 526 (2016) 28–36,
https://doi.org/10.1016/j.apcata.2016.07.026.
The authors declare that they have no known competing financial [23] M. John, K. Alexopoulos, M.F. Reyniers, G.B. Marin, Mechanistic insights into the
formation of butene isomers from 1-butanol in H-ZSM-5: DFT based microkinetic
interests or personal relationships that could have appeared to influence modelling, Catal. Sci. Technol. 7 (2017) 1055–1072, https://doi.org/10.1039/
the work reported in this paper. c6cy02474b.
[24] M. John, K. Alexopoulos, M.F. Reyniers, G.B. Marin, First-principles kinetic study
on the effect of the zeolite framework on 1-butanol dehydration, ACS Catal. 6
Appendix A. Supplementary data (2016) 4081–4094, https://doi.org/10.1021/acscatal.6b00708.
[25] V.C.S. Palla, D. Shee, S.K. Maity, Production of aromatics from n-butanol over
Supplementary data to this article can be found online at https://doi. HZSM-5, H-β, and γ-Al2O3 : role of silica/alumina mole ratio and effect of pressure,
ACS Sustain. Chem. Eng. 8 (2020) 15230–15242, https://doi.org/10.1021/
org/10.1016/j.micromeso.2021.111216.
acssuschemeng.0c04888.
[26] Y. Jiang, J. Liu, W. Jiang, Y. Yang, S. Yang, Current status and prospects of
References industrial bio-production of n-butanol in China, Biotechnol. Adv. 33 (2015)
1493–1501, https://doi.org/10.1016/j.biotechadv.2014.10.007.
[27] B. Ndaba, I. Chiyanzu, S. Marx, n-Butanol derived from biochemical and chemical
[1] P.S. Rezaei, H. Shafaghat, W.M.A.W. Daud, Production of green aromatics and
routes: a review, Biotechnol. Rep. 8 (2015) 1–9, https://doi.org/10.1016/j.
olefins by catalytic cracking of oxygenate compounds derived from biomass
btre.2015.08.001.
pyrolysis: a review, Appl. Catal. A Gen. 469 (2014) 490–511, https://doi.org/
[28] J. Sun, Y. Wang, Recent advances in catalytic conversion of ethanol to chemicals,
10.1016/j.apcata.2013.09.036.
ACS Catal. 4 (2014) 1078–1090, https://doi.org/10.1021/cs4011343.
[2] G. Fogassy, N. Thegarid, G. Toussaint, A.C. van Veen, Y. Schuurman, C. Mirodatos,
[29] P. Dejaifve, A. Auroux, P.C. Gravelle, J.C. Védrine, Z. Gabelica, E.G. Derouane,
Biomass derived feedstock co-processing with vacuum gas oil for second-
Methanol conversion on acidic ZSM-5, offretite, and mordenite zeolites: a
generation fuel production in FCC units, Appl. Catal. B Environ. 96 (2010)
comparative study of the formation and stability of coke deposits, J. Catal. 70
476–485, https://doi.org/10.1016/j.apcatb.2010.03.008.
(1981) 123–136, https://doi.org/10.1016/0021-9517(81)90322-5.
[3] Z. Liu, L. Xu, L. Yu, S. Xie, W. Xin, S. Huang, S. Zhang, Transformation of isobutyl
[30] J.M. Fougerit, N.S. Gnep, M. Guisnet, Selective transformation of methanol into
alcohol to aromatics over zeolite-based catalysts, ACS Catal. 2 (2012) 1203–1210,
light olefins over a mordenite catalyst: reaction scheme and mechanism,
https://doi.org/10.1021/cs300048u.
Microporous Mesoporous Mater. 29 (1999) 79–89, https://doi.org/10.1016/
[4] M.D. Boot, M. Tian, E.J.M. Hensen, S. Mani Sarathy, Impact of fuel molecular
S1387-1811(98)00322-9.
structure on auto-ignition behavior – design rules for future high performance
[31] Y. Gao, B. Zheng, G. Wu, F. Ma, C. Liu, Effect of the Si/Al ratio on the performance
gasolines, Prog. Energy Combust. Sci. 60 (2017) 1–25, https://doi.org/10.1016/j.
of hierarchical ZSM-5 zeolites for methanol aromatization, RSC Adv. 6 (2016)
pecs.2016.12.001.
83581–83588, https://doi.org/10.1039/c6ra17084f.
[5] A. Maneffa, P. Priecel, J.A. Lopez-Sanchez, Biomass-derived renewable aromatics:
[32] M. Conte, J.A. Lopez-Sanchez, Q. He, D.J. Morgan, Y. Ryabenkova, J.K. Bartley, A.
selective routes and outlook for p-xylene commercialisation, ChemSusChem 9
F. Carley, S.H. Taylor, C.J. Kiely, K. Khalid, G.J. Hutchings, Modified zeolite ZSM-5
(2016) 2736–2748, https://doi.org/10.1002/cssc.201600605.
for the methanol to aromatics reaction, Catal. Sci. Technol. 2 (2012) 105–112,
[6] C.A. Mullen, A.A. Boateng, Production of aromatic hydrocarbons via catalytic
https://doi.org/10.1039/c1cy00299f.
pyrolysis of biomass over Fe-modified HZSM-5 zeolites, ACS Sustain. Chem. Eng. 3
[33] T. Liang, H. Toghiani, Y. Xiang, Transient kinetic study of ethane and ethylene
(2015) 1623–1631, https://doi.org/10.1021/acssuschemeng.5b00335.
aromatization over zinc-exchanged HZSM-5 catalyst, Ind. Eng. Chem. Res. 57
[7] C.L. Williams, C. Chang, P. Do, N. Nikbin, S. Caratzoulas, D.G. Vlachos, R.F. Lobo,
(2018) 15301–15309, https://doi.org/10.1021/acs.iecr.8b03735.
W. Fan, P.J. Dauenhauer, Cycloaddition of biomass-derived furans for catalytic

20
T. Kella and D. Shee Microporous and Mesoporous Materials 323 (2021) 111216

[34] A. Mehdad, R.F. Lobo, Ethane and ethylene aromatization on zinc-containing [50] E.-M. El-Malki, R.A. van Santen, W.M.H. Sachtler, Introduction of Zn, Ga, and Fe
zeolites, Catal. Sci. Technol. 7 (2017) 3562–3572, https://doi.org/10.1039/ into HZSM-5 cavities by sublimation: identification of acid sites, J. Phys. Chem. B
c7cy00890b. 103 (1999) 4611–4622, https://doi.org/10.1021/jp990116l.
[35] Y. Ono, H. Kitagawa, Y. Sendoda, Transformation of but-1-ene into aromatic [51] H. Coqueblin, A. Richard, D. Uzio, L. Pinard, Y. Pouilloux, F. Epron, Effect of the
hydrocarbons over ZSM-5 zeolites, J. Chem. Soc. Faraday Trans. 1 Phys. Chem. metal promoter on the performances of H-ZSM5 in ethylene aromatization, Catal.
Condens. Phases 83 (1987) 2913, https://doi.org/10.1039/f19878302913. Today 289 (2017) 62–69, https://doi.org/10.1016/j.cattod.2016.08.006.
[36] T. Wang, X. Tang, X. Huang, W. Qian, Y. Cui, X. Hui, W. Yang, F. Wei, Conversion [52] K. Wang, M. Dong, X. Niu, J. Li, Z. Qin, W. Fan, J. Wang, Highly active and stable
of methanol to aromatics in fluidized bed reactor, Catal. Today 233 (2014) 8–13, Zn/ZSM-5 zeolite catalyst for the conversion of methanol to aromatics: effect of
https://doi.org/10.1016/j.cattod.2014.02.007. support morphology, Catal. Sci. Technol. 8 (2018) 5646–5656, https://doi.org/
[37] L.M. Lubango, M.S. Scurrell, Light alkanes aromatization to BTX over Zn-ZSM-5 10.1039/c8cy01734d.
catalysts: enhancements in BTX selectivity by means of a second transition metal [53] A.A. Gabrienko, S.S. Arzumanov, A.V. Toktarev, I.G. Danilova, I.P. Prosvirin, V.
ion, Appl. Catal. A Gen. 235 (2002) 265–272, https://doi.org/10.1016/S0926- V. Kriventsov, V.I. Zaikovskii, D. Freude, A.G. Stepanov, Different efficiency of Zn2
860X(02)00271-5. +
and ZnO species for methane activation on Zn-modified zeolite, ACS Catal. 7
[38] S. Tamiyakul, T. Sooknoi, L.L. Lobban, S. Jongpatiwut, Generation of reductive Zn (2017) 1818–1830, https://doi.org/10.1021/acscatal.6b03036.
species over Zn/HZSM–5 catalysts for n–pentane aromatization, Appl. Catal. A [54] N. Wang, Y. Hou, W. Sun, D. Cai, Z. Chen, L. Liu, B. Ge, L. Hu, W. Qian, F. Wei,
Gen. 525 (2016) 190–196, https://doi.org/10.1016/j.apcata.2016.07.020. Modulation of b-axis thickness within MFI zeolite: correlation with variation of
[39] J.A. Biscardi, G.D. Meitzner, E. Iglesia, Structure and density of active Zn species in product diffusion and coke distribution in the methanol-to-hydrocarbons
Zn/H-ZSM5 propane aromatization catalysts, J. Catal. 179 (1998) 192–202, conversion, Appl. Catal. B Environ. 243 (2019) 721–733, https://doi.org/10.1016/
https://doi.org/10.1006/jcat.1998.2177. j.apcatb.2018.11.023.
[40] J. Liu, A.S.N. La Hong, N. He, G. Liu, C. Liang, X. Zhang, H. Guo, The crucial role of [55] P. He, J. Jarvis, S. Meng, A. Wang, S. Kou, R. Gatip, M. Yung, L. Liu, H. Song, Co-
reaction pressure in the reaction paths for i-butane conversion over Zn/HZSM-5, aromatization of methane with olefins: the role of inner pore and external surface
Chem. Eng. J. 218 (2013) 1–8, https://doi.org/10.1016/j.cej.2012.12.037. catalytic sites, Appl. Catal. B Environ. 234 (2018) 234–246, https://doi.org/
[41] N. Viswanadham, A.R. Pradhan, N. Ray, S.C. Vishnoi, U. Shanker, T.S.R. Prasada 10.1016/j.apcatb.2018.04.034.
Rao, Reaction pathways for the aromatization of paraffins in the presence of H- [56] N.R. Meshram, Selective toluene disproportionation over ZSM-5 zeolites, J. Chem.
ZSM-5 and Zn/H-ZSM-5, Appl. Catal. A Gen. 137 (1996) 225–233, https://doi.org/ Technol. Biotechnol. 37 (1987) 111–122, https://doi.org/10.1002/
10.1016/0926-860X(95)00287-1. jctb.280370206.
[42] B.S. Liu, Y. Zhang, J.F. Liu, M. Tian, F.M. Zhang, C.T. Au, A.S.-C. Cheung, [57] N.R. Meshram, S.B. Kulkarni, P. Ratnasamy, Transalkylation of toluene with C9
Characteristic and mechanism of methane dehydroaromatization over Zn-Based/ aromatic hydrocarbons over ZSM-5 zeolites, J. Chem. Technol. Biotechnol. Chem.
HZSM-5 catalysts under conditions of atmospheric pressure and supersonic jet Technol. 34 (1984) 119–126, https://doi.org/10.1002/jctb.5040340307. A.
expansion, J. Phys. Chem. C 115 (2011) 16954–16962, https://doi.org/10.1021/ [58] T.-C. Tsai, W.-H. Chen, S.-B. Liu, C.-H. Tsai, I. Wang, Metal zeolites for
jp2027065. transalkylation of toluene and heavy aromatics, Catal. Today 73 (2002) 39–47,
[43] P. He, J.S. Jarvis, S. Meng, Q. Li, G.M. Bernard, L. Liu, X. Mao, Z. Jiang, H. Zeng, V. https://doi.org/10.1016/S0920-5861(01)00516-8.
K. Michaelis, H. Song, Co-aromatization of methane with propane over Zn/HZSM- [59] T. Tsai, Disproportionation and transalkylation of alkylbenzenes over zeolite
5: the methane reaction pathway and the effect of Zn distribution, Appl. Catal. B catalysts, Appl. Catal. A Gen. 181 (1999) 355–398, https://doi.org/10.1016/
Environ. 250 (2019) 99–111, https://doi.org/10.1016/j.apcatb.2019.03.011. S0926-860X(98)00396-2.
[44] M. Xin, E. Xing, Y. Ouyang, X. Gao, G. Xu, Y. Luo, X. Shu, Insight into interactions [60] H. Zhang, S. Shao, R. Xiao, D. Shen, J. Zeng, Characterization of coke deposition in
among P, Zn and ZSM-5 during bi-component modification on ZSM-5, New J. the catalytic fast pyrolysis of biomass derivates, Energy Fuels 28 (2014) 52–57,
Chem. 44 (2020) 20785–20796, https://doi.org/10.1039/D0NJ04136J. https://doi.org/10.1021/ef401458y.
[45] J. Liu, N. He, W. Zhou, L. Lin, G. Liu, C. Liu, J. Wang, Q. Xin, G. Xiong, H. Guo, [61] C. Yang, M. Qiu, S. Hu, X. Chen, G. Zeng, Z. Liu, Y. Sun, Stable and efficient
Isobutane aromatization over a complete Lewis acid Zn/HZSM-5 zeolite catalyst: aromatic yield from methanol over alkali treated hierarchical Zn-containing HZSM-
performance and mechanism, Catal. Sci. Technol. 8 (2018) 4018–4029, https:// 5 zeolites, Microporous Mesoporous Mater. 231 (2016) 110–116, https://doi.org/
doi.org/10.1039/c8cy00917a. 10.1016/j.micromeso.2016.05.021.
[46] Q.T. Cheng, B.X. Shen, H. Sun, J.G. Zhao, J.C. Liu, Methanol promoted naphtha [62] M. Guisnet, P. Magnoux, Organic chemistry of coke formation, Appl. Catal. A Gen.
catalytic pyrolysis to light olefins on Zn-modified high-silicon HZSM-5 zeolite 212 (2001) 83–96, https://doi.org/10.1016/S0926-860X(00)00845-0.
catalysts, RSC Adv. 9 (2019), https://doi.org/10.1039/c9ra02793a, 20818–20828. [63] D. Bibby, Coke formation in zeolite ZSM-5, J. Catal. 97 (1986) 493–502, https://
[47] F.L. Bleken, S. Chavan, U. Olsbye, M. Boltz, F. Ocampo, B. Louis, Conversion of doi.org/10.1016/0021-9517(86)90020-5.
methanol into light olefins over ZSM-5 zeolite: strategy to enhance propene [64] P. Castaño, G. Elordi, M. Olazar, A.T. Aguayo, B. Pawelec, J. Bilbao, Insights into
selectivity, Appl. Catal. A Gen. (2012) 447–448, https://doi.org/10.1016/j. the coke deposited on HZSM-5, Hβ and HY zeolites during the cracking of
apcata.2012.09.025, 178–185. polyethylene, Appl. Catal. B Environ. 104 (2011) 91–100, https://doi.org/
[48] Y. Kolyagin, V. Ordomsky, Y. Khimyak, A. Rebrov, F. Fajula, I. Ivanova, Initial 10.1016/j.apcatb.2011.02.024.
stages of propane activation over Zn/MFI catalyst studied by in situ NMR and IR [65] L. Palumbo, F. Bonino, P. Beato, M. Bjørgen, A. Zecchina, S. Bordiga, Conversion of
spectroscopic techniques, J. Catal. 238 (2006) 122–133, https://doi.org/10.1016/ methanol to hydrocarbons: spectroscopic characterization of carbonaceous species
j.jcat.2005.11.037. formed over H-ZSM-5, J. Phys. Chem. C 112 (2008) 9710–9716, https://doi.org/
[49] M. Akçay, The surface acidity and characterization of Fe-montmorillonite probed 10.1021/jp800762v.
by in situ FT-IR spectroscopy of adsorbed pyridine, Appl. Catal. A Gen. 294 (2005) [66] H.G. Karge, W. Nießen, H. Bludau, In-situ FTIR measurements of diffusion in
156–160, https://doi.org/10.1016/j.apcata.2005.07.019. coking zeolite catalysts, Appl. Catal. A Gen. 146 (1996) 339–349, https://doi.org/
10.1016/S0926-860X(96)00175-5.

21

You might also like