You are on page 1of 10

Fuel 139 (2015) 401–410

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Effect of Fe and Zn promoters on Mo/HZSM-5 catalyst for methane


dehydroaromatization
Victor Abdelsayed a,b,⇑, Dushyant Shekhawat a, Mark W. Smith a,b
a
National Energy Technology Laboratory, U.S. Department of Energy, 3610 Collins Ferry Rd., Morgantown, WV 26507, USA
b
URS Corporation, 3610 Collins Ferry Rd., Morgantown, WV 26507, USA

h i g h l i g h t s

 Fe–Mo catalyst showed a higher benzene yield compared to Mo catalyst.


 Presence of Fe induced the formation of carbon nanostructures observed on spent catalysts.
 Two Zn species were suggested on working Zn catalysts.
 Promoting Mo/ZSM-5 with both Fe and Zn decrease the catalyst stability due to coke formation.

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

Article history: The promotional effect of Fe in the presence and absence of Zn was studied over Mo/HZSM-5 catalysts for
Received 9 April 2014 the methane dehydroaromatization reaction. The reaction was conducted at 700 °C, 3000 scc/g/h and
Received in revised form 22 August 2014 atmospheric pressure. Adding Fe as the only promoter increased the benzene formation rate by 35% com-
Accepted 26 August 2014
pared to unpromoted Mo after 12 h of reaction time. With Zn as the only promoter the benzene formation
Available online 16 September 2014
rate was initially slightly higher (up to 10%); however, after 3 h it performed similarly to the unpromoted
Mo catalyst. When Mo/HZSM-5 was promoted with both Fe and Zn a significant drop (31%) in benzene
Keywords:
formation was observed. Temperature programmed oxidation of the spent catalysts suggested that unhy-
Methane utilization
Dehydroaromatization reactions
drogenated, high-temperature carbon plays a role in catalyst deactivation of Fe–Zn promoted Mo/HZSM-
Benzene 5 catalyst. Further, Zn-loaded catalysts had a higher formation rate of ethylene, with and without Fe,
Fe–Zn promoter compared to MoFe catalyst, possibly due to coke formation associated with Zn. Characterization of the
ZSM-5 catalysts suggested the presence of two types of Zn species. The first type is likely a loosely bound, unsta-
ble ZnO, which was reduced and vaporized during the reaction. This was supported by elemental micro-
analysis of spent catalysts, which showed that more than 50% of Zn was lost during the reaction testing.
The second type of Zn is more strongly bound to HZSM-5. This is consistent with the NH3-TPD results,
where significantly fewer Bronsted acid sites were observed after Zn addition to Mo catalyst. The cata-
lysts were also characterized by XRD, SEM, surface area and micropore analysis, and the correlation
between performance and catalyst properties is discussed.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction since the indirect route, also called the oxidative route, is associ-
ated with an additional step of converting methane into syngas fol-
The abundance and variety of methane resources, whether in lowed by either Fisher–Tropsch or alcohol synthesis steps, adding
natural gas, coal-bed gas, methane hydrates, biogas, or recently to the process complexity [1,3,6–9]. Nevertheless, direct conver-
shale gas make it a potential alternative feedstock for petrochem- sion of methane into C2+ chemicals is a significant challenge due
icals. Conversion of methane into higher value chemicals via direct to thermodynamic limitations, like high activation energy of meth-
or indirect methods has been, and still is, of major interest [1–5]. ane, and low conversions, even at reaction temperatures above
Direct methane conversion is considered the more ambitious route 700 °C [1–5].
Since first reported by Wang et al. [10], Mo/HZSM-5 has been
⇑ Corresponding author at: National Energy Technology Laboratory, U.S.
one of the most studied catalysts for methane dehydroaromatiza-
Department of Energy, 3610 Collins Ferry Rd., Morgantown, WV 26507, USA. tion (MDHA) reaction. Mo/HZSM-5 is generally accepted as a
Tel.: +1 3042855273; fax: +1 3042854850. bifunctional catalyst. First, carburized Mo activates the methane
E-mail address: abdelsav@netl.doe.gov (V. Abdelsayed). C–H bond and forms the first C–C intermediate species [5,11]. Next

http://dx.doi.org/10.1016/j.fuel.2014.08.064
0016-2361/Ó 2014 Elsevier Ltd. All rights reserved.
402 V. Abdelsayed et al. / Fuel 139 (2015) 401–410

the acid sites in HZSM-5 oligomerize these species into aromatic analysis, and the correlation between performance and catalyst
products [2,12]. The unique pore structural of ZSM-5 is essential properties is discussed.
during MDHA due to the similarity in ZSM-5 pore dimensions
and the dynamic pore diameter of benzene [13,14], which favors 2. Experimental
its high shape selectivity. However, coking problems have hin-
dered the commercial viability of this process. Many researchers 2.1. Catalyst synthesis
have focused on finding ways to suppress coke formation and to
improve aromatic yield, such as modifying the surface of zeolite The ammonium ZSM-5 zeolite with SiO2/Al2O3 mole ratio (SAR)
support [15–17], adding promoters [18–21], adding trace amounts of 55 was supplied by Zeolyst Inc. Ammonium heptamolybdate tet-
of oxidants [22–24], periodic methane/H2 switching [22,25], opti- rahydrate ((NH4)6Mo7O244H2O), zinc nitrate hexahydrate
mizing the number and density of acid sites on HZSM-5 by chang- (Zn(NO3)26H2O), and iron(II) chloride, anhydrous (FeCl2) were
ing the silica to alumina ratio (SAR) [26], and shifting the reaction purchased from Alfa Aesar and were used without any further
equilibrium by removing hydrogen from the reaction using mem- purification.
brane reactors [27,28]. The NH4-ZSM-5 powder was calcined at 500 °C in air for 3 h to
The effects of many promoters on Mo/HZSM-5 catalyst have been convert the powder from the ammonium form to its protonated
investigated in the literature [13,18,29] all with the goal of reducing form (HZSM-5). Conventional incipient wetness impregnation
coke formation, which leads to rapid catalyst deactivation and poor was used to prepare Mo catalysts. Typically, ammonium heptamo-
stability. Among these promoters is Fe, which is a well-known lybdate salt corresponding to 4 wt% Mo was dissolved in deionized
Fischer Tropsch catalyst that promotes chain growth of light hydro- water and was added dropwise to HZSM-5. The powder was mixed
carbons via C–C coupling mechanisms [30,31]. Balancing the Mo to using a mortar and pestle for 15 min before drying it at 75 °C over-
Fe ratio on HZSM-5 is essential to obtain a high aromatic yield. night. The powder was then calcined in air at 550 °C for 4 h. Simi-
Higher Fe/Mo ratio usually results in higher methane conversion larly, Zn and Fe promoted Mo catalysts were prepared by co-
but lower aromatic selectivity due to coke formation, as reported impregnation method. Table 1 lists all the catalysts and their nom-
by several research groups [18,32,33] The promoting role of Fe has inal metallic concentrations used in this study. These metal weight
been explained differently in the literature. Xu et al. [29,34,35] percentages correspond to the following atomic ratios between
reported an improvement in stability and benzene yield with Fe. Mo:Zn:Fe of 7.8:2.8:1.0.
They concluded that Fe initially suppresses benzene formation
while inducing the formation of carbon nanotubes (CNT), but later
2.2. Catalyst characterization
these CNT helped in preventing agglomeration of ZSM-5 particles,
and accelerated reactant and product diffusion process leading to
2.2.1. X-ray diffraction (XRD)
increased benzene yield. Masiero et al. [32] found that Fe is not sim-
Powder X-ray diffraction analysis was performed on a Panalyt-
ply a promoter, but it interacted structurally with Mo to form a new
ical X’pert Pro (PW3040) X-ray diffraction system utilizing Cu Ka
binary phase Fe2(MoO4)3 with different catalytic properties com-
radiation. Samples were placed on a zero diffraction Si holder
pared to unpromoted Mo/HZSM-5. Burns et al. [18] did not observe
and were scanned from 5° to 40° (2h). Analysis was carried out
an initial enhancement in benzene formation rate by adding Fe to
with Highscore Plus Analysis software equipped with a standard
Mo/HZSM-5, but did see improvement after a longer time on stream.
ICDD X-ray diffraction database supplied by Panalytical.
Kubota et al. [36] found that the addition of a promoter did not
increase benzene yield. They ascribed their results to the weakening
2.2.2. Surface area and micropore analysis
of the interaction between the acid sites of HZSM-5 and Mo oxide
Nitrogen adsorption experiments were performed at 77 K using
precursors due to the competitive consumption of zeolite protons
Micromeritics ASAP-2020 unit. Catalyst samples were vacuum-
by added metal atoms. This resulted in the formation of less dis-
degassed at 300 °C for 10 h to remove surface humidity and pre-
persed Mo carbide particles and a decreased number of acid sites.
adsorbed gases before exposure to adsorption gas. The surface area
Aboul-Gheit et al. [37] have justified the low catalytic activity of
was calculated from the N2 isotherm data using the Brunauer–
Mo with group VII metal (Fe, Co, Ni) promoters to the electronic
Emmett–Teller (BET) model [46]. The micropore volumes and
properties of unfilled d-orbitals where they are filled with six to
micropore areas were measured using t-plot analysis [47].
eight d-electrons influencing the degree of chemisorption and
strengthening the CH4 adsorption leading to coke formation.
Most of the existing Zn-based HZSM-5 catalysts are reported for 2.2.3. Ammonia adsorption-temperature programmed desorption
aromatization of light alkanes [38–40] with a few catalysts using (NH3-TPD)
Zn as a promoter under pure methane aromatization [29,38,41–43]. NH3-TPD experiments were carried out using a flow system
Xiong et al. [42,43] reported an improvement in catalyst stability with a thermal conductivity detector (Micromeritics Autochem
and benzene selectivity over Zn Mo- [41] and W- [41–43] 2910). Prior to each TPD run, the catalyst (0.2 g) was heated to
HZSM-5 at 800 °C. Similarly, Aboul-Gheit et al. [44] reported higher 500 °C for 30 min under pure He (50 sccm, ramp 5 °C/min) to
benzene selectivity upon substituting half the Mo concentration remove all moisture. The catalyst was then cooled to 150 °C and
with Zn on HZSM-5 at 700 °C. They correlate coke formation with exposed to 30 sccm 15% NH3 in He for 30 min. The catalyst was
the promoter electronegativity where highly electronegative purged with pure He gas to remove excess ammonia before the
promoters are more likely to react with H atoms on the surface
of HZSM-5 leading to more coke formation. In contrast, Xu et al. Table 1
Composition of catalyst used in this study.
[29,34] did not report any improvement on catalyst stability by
promoting Mo/HZSM-5 with Zn. Catalysta Mo (wt%) Zn (wt%) Fe (wt%)
Very little data has been reported on double promoted Mo/ Mo 4 0 0
HZSM-5 catalysts for methane aromatization [45] possibly due to MoZn 4 1 0
its complexity. The present work investigates the promotional MoZnFe 4 1 0.3
MoFe 4 0 0.3
effect of Fe and Zn on the stability and activity of Mo/HZSM-5
Fe 0 0 0.3
added separately and together on the catalyst. The catalysts were
a
also characterized by XRD, SEM, surface area and micropore Catalysts were supported on HZSM-5.
V. Abdelsayed et al. / Fuel 139 (2015) 401–410 403

temperature was ramped up to 600 °C at 5 °C/min ramp rate to diffraction peaks between 2h = 22–25° indicate that all catalysts
obtain NH3-TPD data profile. retained their ZSM-5 crystal structure after being loaded and cal-
cined with different metal oxides. The absence of any metal oxide
2.2.4. Electron microscopy and X-ray microanalysis peaks in the XRD patterns indicates good metal dispersion with
SEM micrographs and energy dispersive spectroscopic (EDS) small particle size on zeolite surfaces [9,48,49].
analysis were obtained utilizing a JEOL FE-7600 scanning electron It is generally accepted that the small angle X-ray diffraction
microscope (SEM) interfaced to a Thermo-Electron Noran System (SAXRD) patterns of microporous materials are very sensitive to
Seven (NSS) X-ray microanalysis system. The EDS detector utilized the presence of any particles inside their micropore channel struc-
in the X-ray microanalysis was a Thermo-Electron Ultradry Energy tures, where the intensity and the d-spacing of the diffraction peak
dispersive spectrometer, which was calibrated utilizing the Cu Ka changes accordingly [50]. In order to compare the XRD patterns of
line at 8.041 kV. these catalysts, the data were normalized with respect to the dif-
fraction peak at 2h  23.1°. The intensity of the low angle diffrac-
2.3. Catalyst evaluation tion peaks between 2h  8–9° was significantly lower with Zn-
loaded catalysts (MoZn, MoZnFe) compared to Mo and MoFe cata-
2.3.1. Experimental setup and reaction conditions lysts, which could indicate a strong interaction between Zn and the
All reactions were carried out in a continuous fixed-bed flow Al framework on the catalyst surface leading to lower crystallinity
reactor (6.25 mm ID) system from Autoclave Engineers, Inc. (model [32,48]. As a result of the impregnation method used during prep-
# 5010-2334) at 700 °C and at atmospheric pressure. The catalyst aration of these catalysts most of the Zn particles likely remained
powders were pressed and sieved between 20/+60 mesh particle on the zeolite external surfaces and did not diffuse into the pore
size. Typically, 1.0 g catalyst was heated in the reactor under structure of HZSM-5. In contrast, MoO3 can thermally diffuse
50 sccm N2 to 500 °C at a ramp rate of 5 °C/min followed by an inside ZSM-5 micropores and anchor to the Bronsted acid sites as
activation step to convert Mo oxide into its carbide form. This car- suggested by Xu et al. [29] and Borry et al. [51]. No effect on the
burization step [45] was conducted under a flow of 40 sccm H2 and diffraction patterns was observed between MoZn and MoZnFe cat-
10 sccm methane gas mixture. The catalyst temperature was alysts suggesting that Fe addition had no significant effect on the
increased from 500 to 600 °C and held for 2 h before increasing crystallinity of that catalyst compared to Zn. This could be due to
again to 700 °C under 50 sccm under methane gas with 10% Ar the difference in Zn and Fe concentrations.
(used as internal standard for GC) resulting in a WHSV of Fig. 1b shows the influence of metal loading on the micropore
3000 scc/g/h. Once the catalyst temperature reached 700 °C the structure monitored at low angle diffraction peaks between
reaction was conducted for 12 h. All gases used were purchased 2h  7–10°. A relatively large shift in all diffraction peaks accompa-
from Butler Gas Products Co. with UHP grade. nied with slightly lower intensities was observed after HZSM-5
was loaded with 4 wt% Mo. Accordingly, the d-spacing was
expected to be lower for smaller pore volumes resulting in a shift
2.3.2. Product analysis
proportional to the quantity of Mo in the pore structure of ZSM-5.
All reactants and products were detected on the same GC (Perkin
To examine the possible role of Zn and its impact on the Al
Elmer-Claurs 500) equipped with dual detectors (flame ionization
framework structure, the XRD pattern for MoFe catalysts was
(FID) and thermal conductivity (TCD) detectors). The transfer line
obtained. This catalyst had a similar Mo and Fe loading with no
between the reactor and the GC was heat traced to prevent product
Zn added. Similar intensities, but slightly higher d-spacing shifts,
condensation that might cause transfer line plugging. The product
were observed in MoFe as in Mo catalyst suggesting that Fe had
gas steam (H2, methane, Ar, ethane, ethylene) was detected using
an insignificant effect on Mo dispersion in micropores possibly
the TCD equipped with HayeSep N 60/80 , HayeSep T 60/80, Molec-
due to its low concentration. Fig. 1b shows a larger peak shift in
ular Sieve 5A 45/60, and Molecular Sieve 13x 45/60 packed col-
all Zn-free catalysts (Mo and MoFe), which could indicate that in
umns. Benzene and toluene were detected with the FID equipped
the absence of Zn more Mo can migrate into the channels of
with GS-GASPRO capillary column (30 m, 0.32 mm ID) from Agilent
Technologies. The GC used in this study is not capable of detecting
naphthalene; therefore, it had to be separated out from the analysis
gas steam. The GC data were processed using TotalChrom Worksta-
tion software. The GC was calibrated with the appropriate standard (a) (b)
gas mixtures before each run. The reproducibility of the catalysts
MoFe
performance were tested and the results were within +/ 5% error.
Normalized Intensity (a.u.)

2.3.3. Temperature programmed oxidation (TPO) In-situ MoZnFe


TPO runs were measured directly after methane aromatization
reaction. The reactor was ramped from 50 °C to 800 °C at 10 °C/
min under 100 sccm of air and the temperature was kept at MoZn
800 °C for 2 h. The TPO data were recorded using Pfeiffer Omnistar
mass spectrometer (MS) coupled with the reactor. The MS was cal-
ibrated and the CO2 concentration in the outlet flow was Mo
calculated.

3. Results and discussion HZSM-5

6 12 18 24 30 7 8 9 10
3.1. Catalyst characterization
2θ 2θ
3.1.1. X-ray diffraction (XRD) analysis Fig. 1. (a) X-ray diffraction patterns for freshly calcined HZSM-5 support, Mo,
The X-ray diffraction patterns for all freshly prepared catalysts MoZn, MoZnFe and MoFe catalysts. (b) Small angle X-ray diffraction (SAXRD)
together with HZSM-5 support are shown in Fig. 1a. High intensity patterns for these catalysts.
404 V. Abdelsayed et al. / Fuel 139 (2015) 401–410

ZSM-5. The presence of Zn on the external surfaces of HZSM-5 pos- 175


HZSM-5
sibly prevents some Mo from getting into ZSM-5 channels.
Mo
MoZn

Quantity Adsorbed (cm3/g STP)


3.1.2. Surface area and micropore analysis 150 MoZnFe
The BET surface area (SBET), microporous surface area (Smicro), MoFe
external surface area (Sextrer), microporous volume (Vmicro), and
total pore volume (Vtotal) of the studied catalysts are listed in 125
Table 2. The BET surface area and total pore volume decreased as
the total metal weight loading increased on HZSM-5. After loading
4 wt% of Mo on HZSM-5, the BET surface area decreased from 420 100
to 369 cm2/g possibly due to strong Mo interaction with HZSM-5
[49]. Addition of Zn in MoZn and MoZnFe catalysts significantly
affected the microporous texture properties (Smicro and Vmicro) of 75
these catalysts (Table 2 and Fig. 2). This is likely due to strong
interaction between Zn and external Bronsted acid sites on the cat-
alyst external surface blocking some of its pore structures [52]. 50
This change in surface properties was more pronounced in the 0.0 0.2 0.4 0.6 0.8 1.0
presence of both Zn and Fe in MoZnFe catalyst with a microporous Relative Pressure (p/pº)
area of (Smicro 195 m2/g) and a pore volume of (Vmicro 0.08 cm3/g)
compared to MoZn, which had 236 m2/g and 0.10 cm3/g, respec- Fig. 2. N2-adsorption and desorption isotherms for promoted Mo/HZSM-5
catalysts. Solid and open symbols are for N2 adsorption and desorption points,
tively. The reduction in XRD crystallinity of MoZnFe catalyst respectively.
observed in Section 3.1.1 agrees well with its surface properties.
Adding only Fe to Mo (MoFe catalyst) resulted in a decrease in
surface area and microporous volume (Vmicro) indicating that Fe 1.5
mostly deposited on the external surfaces and could have blocked
1.0
some of the micropores of HZSM-5 [35]. This result is consistent
0.5
with the similar d-spacing shift observed in the SAXRD data dis- MoFe
cussed in Section 3.1.1. Similarly, a significant difference was 0.0
1.5
observed in the surface area and pore volume between MoFe and
Normalized TCD Signal (a.u.)

1.0
MoZnFe indicating that the presence of Zn could result in further
0.5 MoZnFe
micropore blockage on HZSM-5.
Fig. 2 shows the N2-adsorption and desorption isotherms for 0.0
1.5
freshly calcined catalysts. All catalysts exhibit typical type I iso-
1.0
therms with a small hysteresis loop in the range of P/Po = 0.4–1.0
0.5 MoZn
revealing some mesoporosity in zeolites catalysts due to the
0.0
adsorption of N2 on the external surface of their crystallites and 1.5
capillary condensation in spaces between crystallites [35,53]. The
1.0
MoZnFe isotherm shows a small hysteresis starting at a higher rel-
0.5
ative pressure of 0.42 indicating blockage of micropores by metal Mo
loading on the external surfaces [52]. 0.0
1.5
1.0
3.1.3. Ammonia-temperature programmed desorption (NH3-TPD) 0.5 HZSM-5
analysis
0.0
The total number of acid sites on the catalysts studied was mea-
sured by the NH3-TPD technique. This characterization can provide 200 400 600
o
information about acid site strength and distribution on the cata- Temperature ( C)
lyst surface, as well as the state of metal dispersion, interaction
Fig. 3. NH3-TPD profiles of for freshly calcined (a) HZSM-5 support, (b) Mo, (c)
with HZSM-5 surface, and how the dispersion of the loaded metal MoZn, (d) MoZnFe and (e) MoFe catalysts.
is affected by the presence of other metals.
Fig. 3 shows the NH3-TPD profiles of HZSM-5 before and after
loading with different metals. Two desorption peaks were
desorption of NH3 from weak acid (mostly Lewis-acid) sites and
observed for pure HZSM-5 at 217 and 406 °C corresponding to
strong acid (mostly Bronsted-acid) sites, respectively [29,3
5,43,50,54]. The fraction of Al sites, which is equivalent to Bronsted
Table 2 acid sites anchored to metal catalyst supported on HZSM-5, was
Surface properties of freshly prepared promoted Mo/HZSM-5 catalysts. estimated based on the difference in Bronsted acid sites on the
Catalyst SBET Smicro Sexter Vmicro Vtotal HZSM-5 before and after metal loading (Table 3). Even with MoZ-
(m2/g)a (m2/g)b (m2/g)b (cm3/g)b (cm3/g)c nFe catalyst about 40% of Bronsted acid sites were still available
HZSM-5 420 273 147 0.12 0.22 suggesting that the Zn and Fe were mostly on the external surface
Mo 369 256 113 0.11 0.19 of the catalyst. According to Tessonnier et al. [55] HZSM-5 with
MoZn 337 236 101 0.10 0.18 similar Si/Al ratio has about 0.64 mmol/g of Bronsted acid sites,
MoZnFe 299 195 104 0.08 0.17 which is higher than the total metal loading in MoZnFe catalyst
MoFe 340 225 115 0.09 0.18
(0.62 mmol metal/g). These confirm that the catalysts were not
a
Calculated using BET method in the range up to P/Po = 0.2. overloaded with metals. Table 3 shows the peak desorption tem-
b
Calculated using t-plot method. peratures, moles of ammonia desorbed per gram of catalyst and
c
Total pore volume up to P/Po = 0.93.
the calculated number of acid sites per gram of catalyst. Loading
V. Abdelsayed et al. / Fuel 139 (2015) 401–410 405

Table 3
Quantitative analysis of NH3-TPD profiles of different Mo-promoted catalysts.

Catalyst Desorption Temp. (°C) MNH3 Fraction of Al anchored


(mmolNH3/gcat) by metal (%)
T1 T2
HZSM-5 217 406 0.63 0.00
Mo 221 403 0.78 46.72
MoZn 221 388 0.84 61.22
MoZnFe 225 398 0.87 58.71
MoFe 235 440 1.03 46.99

of 4 wt% Mo on HZSM-5 resulted in a decrease in the number of


Bronsted acid sites and an increase in the number of Lewis acid
sites. The decrease in Bronsted acid sites suggests that some of
the Mo oxide migrated via surface thermal diffusion mechanism
during heat treatment into ZSM-5 channels anchoring to the bridg-
ing hydroxyl groups of tetrahedral Al framework sites [11,29,56–
59]. These results agree well with what was observed in the XRD
results and micropore analysis in the previous sections, where
the Mo loading was accompanied by some micropore structural
changes of in ZSM-5. Meanwhile, the increase in the Lewis acid
sites upon addition of Mo is also indicative of highly dispersed
MoO3 on ZSM-5. Table 3 shows that the Lewis acid peak shifts to
a slightly higher temperature upon metal loading suggesting stron-
ger Lewis acid sites were created by the metal oxides [32].
Almost no difference in surface acidity is observed between
MoZn and MoZnFe catalysts. Both of these catalysts experienced
a drastic reduction of Bronsted acid sites and an increase in Lewis
acid sites suggesting that the Zn is already occupying most of the
remaining Bronsted acid sites, [42] which is not anchored to Mo,
or exchanging with Al in zeolite framework [60]. Similar results
have been reported for Zn promoted W, where the addition of Zn
resulted not only in the elimination of most Bronsted acid sites,
but also in generating a kind of new medium–strong surface Bron-
sted acid sites [42].
The NH3-TPD profiles of MoFe catalyst show a similar Bronsted
acid intensity compared to Mo catalyst indicating that Fe did not
compete as much as Zn did with Mo for Bronsted acid sites. Simi-
larly, Xu et al. [29] showed that the addition of Fe did not nega-
tively influence the Bronsted acid sites distribution since they are
supported on the external surfaces. Table 3 shows that the MoFe
catalyst desorbs the highest amount of ammonia (MNH3;
1.03 mmol/gcatalyst) corresponding to the total Bronsted and Lewis
acid sites on the catalyst. However, the Bronsted acid sites that are
not anchored in Mo and MoFe are very similar. This could indicate
that the Fe only increased the number of Lewis acid sites with
Fig. 4. (a) Methane conversion, (b) benzene and toluene formation rate and (c)
insignificant effect on the number of Bronsted acid sites confirming
ethane- and ethylene formation rate for promoted Mo/HZSM-5 catalysts at 700 °C
that Fe is mostly deposited on the catalyst surface, as the XRD and under WHSV of 3000 scc/g/hr and under atmospheric pressure.
BET results suggested. A larger temperature peak shift was
observed in Bronsted acid sites for MoFe at 440 °C compared to
Mo at 403 °C similar to what has been reported by Masiero et al. toluene formation rates for all catalysts tested. After 24 min the
[32], where Fe increased the strength of the strong acid sites in benzene formation rate was in the order MoFe > MoZn > MoZ-
MoFe catalyst. nFe > Mo suggesting an initial promoting effect of Zn and Fe. As
the reaction proceeded the MoZnFe catalyst started to decline
3.2. Catalyst evaluation quickly to about 56% of its initial yield, the lowest benzene forma-
tion rate of all catalysts. The MoZn catalyst produced up to 10%
Fig. 4a shows the methane conversion for all studied catalysts. higher benzene formation rate within the first 3 h, after which both
Initially conversion was between 4.6% and 5.3%, but it declined dif- Mo and MoZn started to produce a similar benzene yield, which
ferently for each catalyst with time on stream (TOS) due to coke decreased at the same rate through the end of the test. Similar
formation, which deactivated the metal and Bronsted acid sites effects were reported in the literature [41,42,44] on Zn promoted
[61–64]. The methane conversion was found in the order Mo and W catalysts. Xiong et al. [42,43] reported that Zn promoted
MoFe > MoZn  Mo > MoZnFe. Mo and MoZn catalysts showed W on ZSM-5 had increased conversion by 3%; however, the
almost the same deactivation pattern. The conversion for MoZnFe benzene selectivity was the same as the unpromoted catalyst.
catalyst decreased rapidly; however, the conversion for the MoFe Similarly, previous results on 5 wt% Mo promoted with 0.5 wt%
catalyst was significantly higher. Fig. 4b shows the benzene and Zn on ZSM-5 showed an initial increase in benzene yield over the
406 V. Abdelsayed et al. / Fuel 139 (2015) 401–410

unpromoted catalyst [34]. These results could indicate that the Zn 3.3. Electron microscopy and X-ray microanalysis
species remaining on the catalyst after 3 h were not active and
were probably strongly bound to the zeolite framework. We pro- Fig. 5 displays the SEM images of MoZnFe fresh catalyst, which
pose the presence of two Zn species on the HZSM-5 surface: loosely is a representative for all other fresh catalysts studied. The SEM
bound ZnO and strongly bound Zn species via Bronsted acid sites, images showed aggregates of particles with a primary particle size
Al exchanged mechanism in the ZSM-5 structure [60], or even less than 200 nm and a secondary particle size about 10 lm. Fig. 6
reacted with extraframework Al [65]. Under reaction conditions displays the SEM images for the spent samples after dehydroaro-
of 700 °C, a reductive gaseous atmosphere of hydrocarbons and matization of methane for 12 h. The particles are still aggregated
H2, and solid deposits of carbon formed during catalyst deactiva- (especially in Mo and MoZn) with faceted hexagonal shapes about
tion, loose ZnO could reduce to Zn metal, vaporize (vapor pressure 40 to 60 nm in length. No major morphological differences were
of Zn metal at 700 °C is about 0.1 bar [66]), and eventually leave observed between fresh and spent Mo and MoZn catalysts. Carbon
the reaction zone under the influence of the gas velocity [67]. nanostructures were observed only on spent Fe-containing cata-
Fig. 4b shows the aromatic products (benzene and toluene) lysts (MoFe and MoZnFe) displayed in Fig. 6c and d, respectively.
detected during the methane aromatization. An improvement in Fe possibly induces the formation of these carbon nanostructures
benzene formation rate, about 35%, was observed on MoFe com- under reaction conditions by converting methane or polyaromatic
pared to Mo catalyst after 12 h of reaction. In order to confirm molecules into these 1-D carbon structures as previously reported
the promotional effect of Fe on Mo, the benzene formation rate [29,34,35]. No carbon nanostructures were observed on spent Mo
was measured on 0.3 wt% Fe/HZSM-5 (Fe catalyst). The results or MoZn samples (Fig. 6a and b), which confirm the role of Fe
are shown in Fig. 4b and it showed that Fe had a negligible activity and suggests that the presence of Zn only promote the formation
towards benzene formation for 12 h. Thus, the presence of Fe pro- of amorphous carbon deposits (3-D structures) on the catalyst.
moted the activity of Mo and it is not simply the summation of the Two types of carbon were observed on spent MoZnFe catalyst;
yields produced separately by Fe and Mo on HZSM-5. The higher amorphous and nanostructured due to the presence of both Zn
conversion observed with MoFe catalysts could be due to the for- and Fe on the catalyst as shown in Fig. 6d. The SEM also showed
mation of carbon nano structures as reported previously a heavy carbon formation with web-like structures.
[29,34,35], rather than amorphous microscale carbon deposites, The surface concentrations for Mo, Zn and Fe are listed in Table 4
which deactivate the catalyst more rapidly. These carbon nano- for MoZn and MoZnFe catalysts before and after reaction. The Mo
structures create channels within the working MoFe catalyst and surface concentration did not change significantly in these cata-
help aromatic products diffuse out of the ZSM-5 pore structure lysts during reaction. The same is true for Fe in MoZnFe catalyst.
before further converting into coke. In the presence of Zn these The lower concentration of 2 wt% Mo in the fresh samples suggests
carbon nanostructures were either not present (MoZn) or signifi- that some Mo diffused into the bulk of ZSM-5. The higher Zn con-
cantly reduced (MoZnFe) and therefore were unable to perform centration on fresh catalysts (1.6 wt%) suggests that Zn is mainly
this function since the carbon formed in the presence of Zn was deposited on the external surface of the ZSM-5 particles. A signif-
amorphous. The formation rates of toluene were insignificantly icant Zn loss is observed on spent catalysts. Only about 46% and
different for the catalysts. Compared to the benzene formed, the 37% of Zn is retained in the spent catalyst for MoZn and MoZnFe,
toluene formation rate was about 10%, as shown in Fig. 4b. The respectively. This result suggests the presence of two types of Zn
MoZnFe catalyst had the lowest toluene formed among all on the catalyst: a loosely bound unstable type that is easily
catalyst. reduced ZnO under reaction conditions, and a strongly bounded,
Fig. 4c shows the formation rates of ethane and ethylene. The stable Zn species possibly exchanged with Al or anchored via sur-
amount of ethylene in the product stream, believed to be an inter- face hydroxyl groups in ZSM-5 framework [60,65].
mediate for benzene formation [12,68], was found to be inversely
proportional to that of benzene formation rate. For example the 3.4. Temperature programmed oxidation (TPO) analysis
MoFe had the lowest ethylene and the highest benzene, which is
indicative of its high activity towards aromatizing intermediate The TPO profiles of the spent catalysts are displayed in Fig. 7a.
ethylene into benzene. The Zn promoted catalysts (MoZn and MoZ- Two main carbon types were observed on the spent catalysts: a
nFe) produced a higher ethylene, but lower benzene. The tendency low temperature carbon with an oxidation temperature between
to produce more ethylene and propylene were interpreted as a 431 and 460 °C, and a high temperature carbon with an oxidization
stronger interaction between the promoter and Al acid sites [69], temperature between 610 and 643 °C. The broadness of each peak
which could be applied to Zn as a promoter in this work. The eth- is related to the significant role played by metal catalysts on the
ane production did not show significant difference over these shape of TPO profiles as well as the presence of more than one car-
catalysts. bon structure within each peak. Table 5 shows the oxidation

Fig. 5. SEM images for fresh MoZnFe catalyst measured at two different magnifications (a and b) scale bar of 10 and 50 lm.
V. Abdelsayed et al. / Fuel 139 (2015) 401–410 407

Fig. 6. SEM images of spent catalysts for (a) Mo, (b) MoZn, (c) MoFe and (d) MoZnFe after 12 h of reaction.

Table 4 polyaromatic products [11,71]. Lunsford et al. [62] reported the


Elemental surface composition for fresh and spent MoZn and MoZnFe samples presence of three carbon types on deactivated Mo/HZSM-5 catalyst
obtained by EDX analysis. run at 700 °C during MDHA reaction (a) graphitic-like carbon in the
Catalyst/M (wt%)a MoZn MoZnFe pore structure of ZSM-5, (b) carbidic-like carbon mainly on the
Mo2C predominantly located on the outer surface of ZSM-5 and
Fresh Spent Fresh Spent
(c) hydrogen-poor-sp-type or pregraphitic-type carbon located on
Mo 2.2 ± 0.2 2.1 ± 0.4 2.1 ± 0.3 1.9 ± 0.2
the outer surface of ZSM-5.
Zn 1.6 ± 0.2 0.7 ± 0.2 1.6 ± 0.2 0.6 ± 0.2
Fe 0 0 0.3 ± 0.1 0.4 ± 0.1 The detection of water signal during the TPO experiment of
a
MoFe catalyst accompanying the low temperature oxidation peak
Average metal wt% concentrations and errors were calculated from at least 10
suggests that the burned carbon contained hydrogen (Fig. 7b)
EDX point measurements.
while high temperature carbon was not accompanied by water for-
mation. Tan et al. [72] showed that the hydrogenated carbon com-
temperature peaks (T1 and T2) and their corresponding integrated busts at a higher temperature than non-hydrogenated carbon in
CO2 peak areas (A1 and A2) for each catalyst. contrast to what Hassan and Sayari [73] reported. Our results are
A negligible amount of carbon was observed on spent HZSM-5, in agreement with Hassan and Sayari [73] where the low temper-
which indicates that the presence of metal catalyst is necessary for ature carbon peaks corresponded to hydrogenated coke. These TPO
methane activation [7,12,70]. After Zn metal loading, both spent results are also in agreement with previous work [62,73,74] show-
MoZn and MoZnFe catalysts showed a similarly high temperature ing that the low temperature hydrogenated coke usually does not
carbon peak at 611 and 616 °C, corresponding to an integrated lead to catalyst deactivation, but it is the non-hydrogenated high
CO2 peak area of 46.3 and 52.3, respectively. At the same time a temperature coke that gradually blocks zeolite pores and metal
slight increase in the low temperature carbon peak location was surfaces.
observed from 430 °C for Mo to 441 °C for MoZn to 461 °C for MoZ- Table 5 shows that the high temperature carbon (A2) is a more
nFe, possibly due to higher metal coverage introduced by Zn and stable graphitic carbon, and requires a higher temperature to be
Fe, which could result in more coke formation. oxidized compared to hydrogenated carbon and/or carbidic carbon
The presence of Fe with Zn in MoZnFe catalyst did not affect the (A1) [73–78]. Higher A2 carbon corresponded to more deactivation
amount of high temperature carbon compared to Mo or MoZn cat- of the catalyst. In order to qualitatively determine the state of cat-
alysts, while the amount of low temperature carbon decreased in alyst deactivation, the ratio between these carbon types (A2/A1)
the presence of Fe as shown in Table 5. The MoFe catalyst showed was calculated and listed in Table 5. These ratios correlate well
a lower carbon formation compared to the rest of catalysts. The to the observed reactivity where MoFe catalyst had the highest
drop in the high temperature carbon peak, mostly sp-type carbon activity and stability (A2/A1 = 0.77) while MoZnFe had the lowest
[62] indicates its relative stability and tendency to resist coking activity and stability (A2/A1 = 2.02). These results indicate that
during reaction compared to other catalysts. These results agree the type and amount of high temperature coke is crucial in the cat-
well with its catalytic stability shown in Fig. 4. alyst stability [74]. An increase in ethylene and decrease in aro-
The amount and type of carbon deposited on the catalyst sur- matic formation rates were observed in deactivating catalysts
face plays a significant role in catalyst activity and stability. It is (Mo, MoZn and MoZnFe) with higher A2 compared to A1 agreeing
generally accepted that there are two sources of carbon that lead with the reported [75,76] inertness and irreversible nature of high
to coke formation. One is directly from methane decomposition temperature coke that covers the acid sites responsible for the aro-
and the other is indirectly from the formation of heavy matization step [7,12,70] on these catalysts.
408 V. Abdelsayed et al. / Fuel 139 (2015) 401–410

catalyst showed the formation of carbon nanostructures in the


10 presence of Fe suggesting better reactant and product diffusion
5
(a) MoFe
and coke resistance, also supported by the TPO results.
0 The addition of Zn on Mo (MoZn catalyst) only slightly
increased the benzene formation rate during the first 3 h of reac-
10 MoZnFe tion, and then performed similarly to unpromoted Mo catalyst.
5 The EDX and NH3-TPD showed that the Zn is mostly on the exter-
0 nal catalyst surface. Two types of Zn are possibly on the catalyst
% CO2 (a.u.)

surface: a loosely bound ZnO, which reduced to Zn metal and


10
vaporized under reaction conditions, and a strongly bound Zn spe-
5 MoZn cies that exchanged with the Mo-free anchoring Al sites in the
0 ZSM-5 framework. The EDX results for the spent Zn catalysts
showed more than 50% of the Zn was lost. Additionally, the NH3-
10
TPD showed that the Zn catalysts had significantly reduced Bron-
5 sted acid sites, further supporting the above conclusion. Finally,
Mo
0 the addition of both Fe and Zn on Mo catalyst resulted in a drastic
decrease in benzene formation rate and methane conversion com-
10
pared to unpromoted catalyst due to coke formation. Overall, the
5
HZSM-5 Fe promoted Mo/HZSM-5 produced the greatest amount of ben-
0 zene and toluene, and displayed the most stability of the catalysts
200 300 400 500 600 700 in this study in a single pass 12-h test. Additional studies are
Temperature ( C)
o needed, but are beyond the scope of this study, to determine the
stabilizing effect of these promoters for catalysts during
regeneration.

(b)
H 2O
Intensity (a.u.)

Disclaimer
O2
CO2 This project was funded by the Department of Energy, National
Energy Technology Laboratory, an agency of the United States Gov-
ernment, through a support contract with URS Energy &Construc-
tion, Inc. Neither the United States Government nor any agency
thereof, nor any of their employees, nor URS Energy & Construc-
tion, Inc., nor any of their employees, makes any warranty,
200 400 600 expressed or implied, or assumes any legal liability or responsibil-
o
Temperature ( C) ity for the accuracy, completeness, or usefulness of any informa-
tion, apparatus, product, or process disclosed, or represents that
Fig. 7. (a) In-situ TPO profiles for spent HZSM-5 support, Mo, MoZn, MoZnFe and
its use would not infringe privately owned rights. Reference herein
MoFe catalysts. (b) Formation profiles of O2, H2O and CO2 during the TPO of spent
MoFe catalyst. to any specific commercial product, process, or service by trade
name, trademark, manufacturer, or otherwise, does not necessarily
constitute or imply its endorsement, recommendation, or favoring
Table 5 by the United States Government or any agency thereof. The views
TPO Peaks and integrated CO2 peak area of different Mo-promoted HZSM-5 catalysts. and opinions of authors expressed herein do not necessarily state
Catalyst Peak temperature (°C) Integrated CO2 peak area or reflect those of the United States Government or any agency
T1 T2 A1 A2 A2/A1
thereof.

Mo 431 628 36.3 51.1 1.41


MoZn 441 611 38.7 46.3 1.20
Acknowledgements
MoZnFe 460 616 25.9 52.3 2.02
MoFe 443 643 30.7 23.7 0.77
This work was performed in support of the National Energy
Technology Laboratory’s ongoing research under the RES contract
4. Conclusion DE-FE0004000. We gratefully acknowledge Donald Floyd and
James Poston for reactor setup and electron microscopy,
The promotional effect of Fe and Zn on Mo/HZSM-5 catalyst was respectively. We also acknowledge Dr. Dirk Link and Dr. Bryan
tested for methane dehydroaromatization reaction. Characteriza- D. Morreale for their technical guidance throughout this project.
tion of Fe-promoted catalyst (MoFe) by XRD and micropore analy-
sis suggested that the concentration of Mo inside the pore
References
structure of HZSM-5 remained unchanged with Fe addition and
that Fe is mostly on the external surface of the catalyst blocking [1] Skutil K, Taniewski M. Indirect methane aromatization via oxidative coupling,
some of the pore structure of the ZSM-5. Similarly, the NH3-TPD products separation and aromatization steps. Fuel Process Technol
2007;88:877–82.
results showed a significant increase in the number of Lewis acid
[2] Majhi S, Mohanty P, Wang H, Pant KK. Direct conversion of natural gas to
sites with Fe. The benzene formation rate increased by 35% for higher hydrocarbons: a review. J Energy Chem 2013;22:543–54.
MoFe catalyst compared to unpromoted Mo catalyst after 12 h of [3] Gharibi M, Zangeneh FT, Yaripour F, Sahebdelfar S. Nanocatalysts for
reaction. MoFe catalyst showed the lowest ethylene formation, conversion of natural gas to liquid fuels and petrochemical feedstocks. Appl
Catal A 2012;443–444:8–26.
which is indicative of its high activity towards aromatizing inter- [4] Choudhary TV, Aksoylu E, Wayne Goodman D. Nonoxidative activation of
mediate ethylene into benzene. The SEM images of spent MoFe methane. Catal Rev 2003;45:151–203.
V. Abdelsayed et al. / Fuel 139 (2015) 401–410 409

[5] Alvarez-Galvan MC, Mota N, Ojeda M, Rojas S, Navarro RM, Fierro JLG. Direct CH4 dehydroaromatization under periodic CH4–H2 switching operation at
methane conversion routes to chemicals and fuels. Catal Today 1073 K. Appl Catal A 2013;452:105–16.
2011;171:15–23. [36] Kubota T, Oshima N, Nakahara Y, Yanagimoto M, Okamoto Y. XAFS
[6] Lunsford JH. Catalytic conversion of methane to more useful chemicals and characterization of Mo/ZSM-5 catalysts for methane conversion to benzene:
fuels: a challenge for the 21st century. Catal Today 2000;63:165–74. effect of additives. J Japan Petrol Inst 2006;49:127–33.
[7] Holmen A. Direct conversion of methane to fuels and chemicals. Catal Today [37] Aboul-Gheit AK, El-Masry MS, Awadallah AE. Oxygen free conversion of
2009;142:2–8. natural gas to useful hydrocarbons and hydrogen over monometallic Mo and
[8] Skutil K, Taniewski M. Some technological aspects of methane aromatization bimetallic Mo–Fe, Mo–Co or Mo–Ni/HZSM-5 catalysts prepared by mechanical
(direct and via oxidative coupling). Fuel Process Technol 2006;87:511–21. mixing. Fuel Process Technol 2012;102:24–9.
[9] Abdelsayed V, Shekhawat D, Poston Jr JA, Spivey JJ. Synthesis, characterization, [38] Liu JF, Liu Y, Peng LF. Aromatization of methane by using propane as co-
and catalytic activity of Rh-based lanthanum zirconate pyrochlores for higher reactant over cobalt and zinc-impregnated HZSM-5 catalysts. J Mol Catal A:
alcohol synthesis. Catal Today 2013;207:65–73. Chem 2008;280:7–15.
[10] Wang L, Tao L, Xie M, Xu G, Huang J, Xu Y. Dehydrogenation and aromatization [39] Luzgin MV, Rogov VA, Arzumanov SS, Toktarev AV, Stepanov AG, Parmon VN.
of methane under non-oxidizing conditions. Catal Lett 1993;21:35–41. Methane aromatization on Zn-modified zeolite in the presence of a co-reactant
[11] Ding W, Li S, Meitzner GD, Iglesia E. Methane conversion to aromatics on Mo/ higher alkane: how does it occur? Catal Today 2009;144:265–72.
H-ZSM5: structure of molybdenum species in working catalysts. J Phys Chem B [40] Lubango LM, Scurrell MS. Light alkanes aromatization to BTX over Zn-ZSM-5
2001;105:506–13. catalysts: enhancements in BTX selectivity by means of a second transition
[12] Xu Y, Lin L. Recent advances in methane dehydro-aromatization over metal ion. Appl Catal A 2002;235:265–72.
transition metal ion-modified zeolite catalysts under non-oxidative [41] Zeng JL, Xiong ZT, Zhang HB, Lin GD, Tsai KR. Nonoxidative dehydrogenation
conditions. Appl Catal A 1999;188:53–67. and aromatization of methane over W/HZSM-5-based catalysts. Catal Lett
[13] Ma S, Guo X, Zhao L, Scott S, Bao X. Recent progress in methane 1998;53:119–24.
dehydroaromatization: from laboratory curiosities to promising technology. [42] Xiong Z-T, Zhang H-B, Lin G-D, Zeng J-L. Study of W/HZSM-5-based catalysts
J Energy Chem 2013;22:1–20. for dehydro-aromatization of CH4 in absence of O2. II. Action of promoters Zn
[14] Zhang C-L, Li S, Yuan Y, Zhang W-X, Wu T-H, Lin L-W. Aromatization of and Li. Catal Lett 2001;74:233–9.
methane in the absence of oxygen over Mo-based catalysts supported on [43] Xiong Z-T, Chen L-L, Zhang H-B, Zeng J-L, Lin G-D. Study of W/HZSM-5-based
different types of zeolites. Catal Lett 1998;56:207–13. catalysts for dehydro-aromatization of CH4 in absence of O2. I. Performance of
[15] Ding W, Meitzner GD, Iglesia E. The effects of silanation of external acid sites catalysts. Catal Lett 2001;74:227–32.
on the structure and catalytic behavior of Mo/H–ZSM5. J Catal [44] Aboul-Gheit AK, Awadallah AE, Aboul-Enein AA, Mahmoud A-LH.
2002;206:14–22. Molybdenum substitution by copper or zinc in H-ZSM-5 zeolite for
[16] Kikuchi S, Kojima R, Ma H, Bai J, Ichikawa M. Study on Mo/HZSM-5 catalysts catalyzing the direct conversion of natural gas to petrochemicals under non-
modified by bulky aminoalkyl-substituted silyl compounds for the selective oxidative conditions. Fuel 2011;90:3040–6.
methane-to-benzene (MTB) reaction. J Catal 2006;242:349–56. [45] Sily PD, Noronha FB, Passos FB. Methane direct conversion on Mo/ZSM-5
[17] Li Y, Liu L, Huang X, Liu X, Shen W, Xu Y, et al. Enhanced performance of catalysts modified by Pd and Ru. J Nat Gas Chem 2006;15:82–6.
methane dehydro-aromatization on Mo-based HZSM-5 zeolite pretreated by [46] Brunauer S, Emmett PH, Teller E. Adsorption of gases in multimolecular layers.
NH4F. Catal Commun 2007;8:1567–72. J Am Chem Soc 1938;60:309–19.
[18] Burns S, Hargreaves JSJ, Pal P, Parida KM, Parija S. The effect of dopants on the [47] Lippens BC, Linsen BG, Boer JHD. Studies on pore systems in catalysts I. The
activity of MoO3/ZSM-5 catalysts for the dehydroaromatisation of methane. adsorption of nitrogen; apparatus and calculation. J Catal 1964;3:32–7.
Catal Today 2006;114:383–7. [48] El-Shall MS, Abdelsayed V, Khder AERS, Hassan HMA, El-Kaderi HM, Reich TE.
[19] Li S, Zhang C, Kan Q, Wang D, Wu T, Lin L. The function of Cu(II) ions in the Mo/ Metallic and bimetallic nanocatalysts incorporated into highly porous
CuH-ZSM-5 catalyst for methane conversion under non-oxidative condition. coordination polymer MIL-101. J Mater Chem 2009;19:7625–31.
Appl Catal A 1999;187:199–206. [49] Liu H, Shen W, Bao X, Xu Y. Methane dehydroaromatization over Mo/HZSM-5
[20] Chul Hwang I, Heui Kim D, Ihl Woo S. Role of oxygen on NOx SCR catalyzed catalysts: the reactivity of MoCx species formed from MoOx associated and
over Cu/ZSM-5 studied by FTIR TPD XPS and micropulse reaction. Catal Today non-associated with Brönsted acid sites. Appl Catal A 2005;295:79–88.
1998;44:47–55. [50] Li B, Li S, Li N, Chen H, Zhang W, Bao X, et al. Structure and acidity of Mo/ZSM-5
[21] Liu BS, Jiang L, Sun H, Au CT. XPS, XAES, and TG/DTA characterization of synthesized by solid state reaction for methane dehydrogenation and
deposited carbon in methane dehydroaromatization over Ga–Mo/ZSM-5 aromatization. Microporous Mesoporous Mater 2006;88:244–53.
catalyst. Appl Surf Sci 2007;253:5092–100. [51] Borry RW, Kim YH, Huffsmith A, Reimer JA, Iglesia E. Structure and density of
[22] Ma H, Kojima R, Kikuchi S, Ichikawa M. Effective coke removal in methane to Mo and acid sites in Mo-exchanged H-ZSM5 catalysts for nonoxidative
benzene (MTB) reaction on Mo/HZSM-5 catalyst by H2 and H2O Co-addition to methane conversion. J Phys Chem B 1999;103:5787–96.
methane. Catal Lett 2005;104:63–6. [52] Mhamdi M, Ghorbel A, Delahay G. Influence of the V+Mo/Al ratio on vanadium
[23] Bai J, Liu S, Xie S, Xu L, Lin L. Shape selectivity in methane and molybdenum speciation and catalytic properties of V-Mo–ZSM-5
dehydroaromatization over Mo/MCM-22 catalysts during a lifetime prepared by solid-state reaction. Catal Today 2009;142:239–44.
experiment. Catal Lett 2003;90:123–30. [53] Liu H, Yang S, Hu J, Shang F, Li Z, Xu C, et al. A comparison study of mesoporous
[24] Bradford MCJ, Te M, Konduru M, Fuentes DX. CH4–C2H6–CO2 conversion to Mo/H-ZSM-5 and conventional Mo/H-ZSM-5 catalysts in methane non-
aromatics over Mo/SiO2/H-ZSM-5. Appl Catal A 2004;266:55–66. oxidative aromatization. Fuel Process Technol 2012;96:195–202.
[25] Rodrigues ACC, Monteiro JLF. The use of CH4/H2 cycles on [54] Cui Y, Xu Y, Lu J, Suzuki Y, Zhang Z-G. The effect of zeolite particle size on the
dehydroaromatization of methane over MoMCM-22. Catal Commun activity of Mo/HZSM-5 in non-oxidative methane dehydroaromatization. Appl
2008;9:1060–5. Catal A 2011;393:348–58.
[26] Sarıoğlan A, Savasßçı ÖT, Erdem-S
ß enatalar A, Tuel A, Sapaly G, Ben Taârit Y. The [55] Tessonnier J-P, Louis B, Rigolet S, Ledoux MJ, Pham-Huu C. Methane dehydro-
effect of support morphology on the activity of HZSM-5-supported aromatization on Mo/ZSM-5: about the hidden role of Brønsted acid sites. Appl
molybdenum catalysts for the aromatization of methane. J Catal Catal A 2008;336:79–88.
2007;246:35–9. [56] Aboul-Gheit AK, Awadallah AE, El-Kossy SM, Mahmoud A-LH. Effect of Pd or Ir
[27] Iliuta MC, Larachi F, Grandjean BPA, Iliuta I, Sayari A. Methane nonoxidative on the catalytic performance of Mo/H-ZSM-5 during the non-oxidative
aromatization over Ru–Mo/HZSM-5 in a membrane catalytic reactor. Ind Eng conversion of natural gas to petrochemicals. J Nat Gas Chem 2008;17:337–43.
Chem Res 2002;41:2371–8. [57] Gao J, Zheng Y, Fitzgerald GB, de Joannis J, Tang Y, Wachs IE, et al. Structure of
[28] Liu Z, Li L, Iglesia E. Catalytic pyrolysis of methane on Mo/H-ZSM5 with Mo2Cx and Mo4Cx molybdenum carbide nanoparticles and their anchoring
continuous hydrogen removal by permeation through dense oxide films. Catal sites on ZSM-5 zeolites. J Phys Chem C 2014;118:4670–9.
Lett 2002;82:175–80. [58] Bedard J, Hong D-Y, Bhan A. CH4 dehydroaromatization on Mo/H–ZSM-5: 1.
[29] Xu Y, Wang J, Suzuki Y, Zhang Z-G. Effect of transition metal additives on the effects of co-processing H2 and CH3COOH. J Catal 2013;306:58–67.
catalytic stability of Mo/HZSM-5 in the methane dehydroaromatization under [59] Zheng H, Ma D, Bao X, Hu JZ, Kwak JH, Wang Y, et al. Direct observation of the
periodic CH4–H2 switch operation at 1073 K. Appl Catal A 2011;409– active center for methane dehydroaromatization using an ultrahigh field 95Mo
410:181–93. NMR spectroscopy. J Am Chem Soc 2008;130:3722–3.
[30] Rofer-DePoorter CK. A comprehensive mechanism for the Fischer–Tropsch [60] Ni Y, Sun A, Wu X, Hai G, Hu J, Li T, et al. The preparation of nano-sized H[Zn,
synthesis. Chem Rev 1981;81:447–74. Al]ZSM-5 zeolite and its application in the aromatization of methanol.
[31] Gaube J, Klein HF. Studies on the reaction mechanism of the Fischer–Tropsch Microporous Mesoporous Mater 2011;143:435–42.
synthesis on iron and cobalt. J Mol Catal A: Chem 2008;283:60–8. [61] Solymosi F, Szöke A, Cserényi J. Conversion of methane to benzene over Mo2C
[32] Masiero S, Marcilio N, Perez-Lopez O. Aromatization of methane over Mo–Fe/ and Mo2C/ZSM-5 catalysts. Catal Lett 1996;39:157–61.
ZSM-5 catalysts. Catal Lett 2009;131:194–202. [62] Weckhuysen B, Rosynek M, Lunsford J. Characterization of surface carbon
[33] Liu S, Dong Q, Ohnishi R, Ichikawa M. Remarkable non-oxidative conversion of formed during the conversion of methane to benzene over Mo/H-ZSM-5
methane to naphthalene and benzene on Co and Fe modified Mo/HZSM-5 catalysts. Catal Lett 1998;52:31–6.
catalysts. Chem Commun 1997:1455–6. [63] Honda K, Yoshida T, Zhang Z-G. Methane dehydroaromatization over Mo/
[34] Xu Y, Wang J, Suzuki Y, Zhang Z-G. Improving effect of Fe additive on the HZSM-5 in periodic CH4–H2 switching operation mode. Catal Commun
catalytic stability of Mo/HZSM-5 in the methane dehydroaromatization. Catal 2003;4:21–6.
Today 2012;185:41–6. [64] Honda K, Chen X, Zhang Z-G. Identification of the coke accumulation and
[35] Xu Y, Suzuki Y, Zhang Z-G. Comparison of the activity stabilities of nanosized deactivation sites of Mo2C/HZSM-5 catalyst in CH4 dehydroaromatization.
and microsized zeolites based Fe–Mo/HZSM-5 catalysts in the non-oxidative Catal Commun 2004;5:557–61.
410 V. Abdelsayed et al. / Fuel 139 (2015) 401–410

[65] Cho S, Lee K-H. Formation of zinc aluminum mixed metal oxide [72] Tan PL, Wong KW, Au CT, Lai SY. Effects of Co-fed O2 and CO2 on the
nanostructures. J Alloys Compd 2011;509:8770–8. deactivation of Mo/HZSM-5 for methane aromatization. Appl Catal A
[66] Greenbank JC, Argent BB. Vapour pressure of magnesium, zinc and cadmium. 2003;253:305–16.
Trans Faraday Soc 1965;61:655–64. [73] Hassan A, Sayari A. Highly active, selective and stable Mo/Ru-HZSM-5 catalysts
[67] Xue MS, Li W, Wang FJ. Effect of surface ZnO coatings on oxidation and thermal for oxygen-free methane aromatization. Appl Catal A 2006;297:159–64.
stability of zinc films. Superlattices Microstruct 2010;48:213–20. [74] Ma D, Shu Y, Cheng M, Xu Y, Bao X. On the Induction period of methane
[68] Spivey JJ, Hutchings G. Catalytic aromatization of methane. Chem Soc Rev aromatization over Mo-based catalysts. J Catal 2000;194:105–14.
2014;43:792–803. [75] Rodrigues A, Monteiro J. CO2 addition on the non-oxidative dehydro-
[69] Aboul-Gheit AK, Awadallah AE. Effect of combining the metals of group VI aromatization of methane over MoMCM-22. Catal Lett 2007;117:166–70.
supported on H-ZSM-5 zeolite as catalysts for non-oxidative conversion of [76] Liu H, Li T, Tian B, Xu Y. Study of the carbonaceous deposits formed on a Mo/
natural gas to petrochemicals. J Nat Gas Chem 2009;18:71–7. HZSM-5 catalyst in methane dehydro-aromatization by using TG and
[70] Mamonov NA, Fadeeva EV, Grigoriev DA, Mikhailov MN, Leonid MK, Alkhimov temperature-programmed techniques. Appl Catal A 2001;213:103–12.
SA. Metal/zeolite catalysts of methane dehydroaromatization. Russ Chem Rev [77] Kojima R, Kikuchi S, Ma H, Bai J, Ichikawa M. Promotion effects of Pt and Rh on
2013;82:567. catalytic performances of Mo/HZSM-5 and Mo/HMCM-22 in selective
[71] Yuan S, Li J, Hao Z, Feng Z, Xin Q, Ying P, et al. The effect of oxygen on the methane-to-benzene reaction. Catal Lett 2006;110:15–21.
aromatization of methane over the Mo/HZSM-5 catalyst. Catal Lett [78] Jiang H, Wang L, Cui W, Xu Y. Study on the induction period of methane
1999;63:73–7. aromatization over Mo/HZSM-5: partial reduction of Mo species and
formation of carbonaceous deposit. Catal Lett 1999;57:95–102.

You might also like