From Synthesis Gas Production To Methanol Synthesi 2015 Journal of Natural G

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of Natural Gas Science and Engineering 25 (2015) 303e316

Contents lists available at ScienceDirect

Journal of Natural Gas Science and Engineering


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

Review article

From synthesis gas production to methanol synthesis and potential


upgrade to gasoline range hydrocarbons: A review
Ahmad Galadima a, Oki Muraza a, b, *
a
Center of Research Excellence in Nanotechnology, King Fahd University of Petroleum & Minerals, Dhahran 30261, Saudi Arabia
b
Chemical Engineering Department, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia

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

Article history: Methanol to gasoline (MTG) process is among the methanol to hydrocarbon (MTH) technologies with
Received 19 March 2015 industrial interest. The key sequential routes involved syngas (H2, CO) production, methanol synthesis
Received in revised form and the subsequent upgrade to gasoline range paraffins via the MTG process. The paper reviewed recent
27 April 2015
literature on prospective catalysts for the MTG technology. Issues related to the role of catalyst-topology,
Accepted 8 May 2015
Available online 17 May 2015
acidity properties, reaction parameters and their effects on catalytic activity, selectivity and catalyst
lifetime have been critically covered. The review also captured background details on the updates related
to the syngas production and subsequent upgrade to methanol prior to the MTG reaction.
Keywords:
Shale gas
© 2015 Elsevier B.V. All rights reserved.
Syngas
Methanol
Gasoline
Catalysts

1. Introduction alternative given consideration today is the production from non-


fossil sources. High octane gasoline can be produced from the
Gasoline is an important liquid hydrocarbon-based fuel derived hydrotreatment and subsequent hydroisomerization of vegetable
primarily from fractional distillation of petroleum fractions. It can oils. This technology is already under commercial consideration by
be produced in a variety of grades depending on the demand and the global refineries (Milne et al., 1990; Saxena and Viswanadham,
applications. The commodity with major application as fuel for 2014; Malleswara Rao et al., 2012). Several researches have also
internal combustion engines, comprised mainly of light to medium been published and are underway for this technology. Recently,
alkanes (straight chains and isomers), with certain concentrations García-Da vila et al. (2014) demonstrated the potential of jatropha
of aromatics as octane enhancers (Erofeev et al., 2014; Song et al., oil for upgrade to linear and isomerized alkanes, with composition
2015; Galadima et al., 2012), although their usage have been ban- in the gasoline range, using supported nickel catalysts. Their
ned by environmental agencies in the recent times, due to associ- hydrodeoxygenation-hydrocracking approach involved the initial
ated environmental and health consequences (Ou et al., 2015; conversion of the oil into high molecular weight alkanes that were
Agarwal et al., 2015). Globally, gasoline is popularly employed as subsequently cracked into gasoline range compounds. Several
a major commodity for transportation and fuel and other authors have also employed different catalysts, vegetable oils
petrochemicals-based industrial applications (Chang et al., 1976; and reaction conditions for this process (Maher and Bressler, 2007;
McGillivray, 1976). While its global demand was projected to rise Furimsky, 2000; Kubi cka et al., 2009; Huber et al., 2007; Liu et al.,
for many world regions, particularly due to increase in the number 2011; Charusiri et al., 2006). Another modern technology for gas-
of automobiles and industrial-based internal combustion engines, oline production is the catalytic upgrading of methanol, a process
the available crude oil reserves are on the decline (Owen et al., otherwise called “methanol to gasoline” (MTG). Methanol has been
2010; Campbell and Laherre re, 1998; Edwards, 1997). One major successfully converted into a range of olefinic and aromatic hy-
drocarbons using different solid acid catalysts like zeolites and
phosphate based catalysts (Aghamohammadi and Haghighi, 2015;
* Corresponding author. Center of Research Excellence in Nanotechnology and Yaripour et al., 2015; Aghaei and Haghighi, 2015). The technology
Chemical Engineering Department, King Fahd University of Petroleum & Minerals, is therefore being modified towards limiting the reaction selectivity
Dhahran 31261, Saudi Arabia. to these compounds with enhanced selectivity to gasoline range
E-mail address: [email protected] (O. Muraza).

http://dx.doi.org/10.1016/j.jngse.2015.05.012
1875-5100/© 2015 Elsevier B.V. All rights reserved.
304 A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316

alkanes. One important issue of interest is the possibility of by upgrading the synthesis gas (i.e. syngas) into methanol. The
obtaining the feedstock at economical scale from known sources. principal raw materials for synthesis gas production are natural gas,
Methanol can be derived from synthesis gas (H2, CO), which in methane gas from associated petroleum, shale gas, coal and
principle can be produced from the reforming of abundant natural biomass. Synthesis from fossil sources can be successfully achieved
gas reserves or biomass-based materials. by the reforming technologies (i.e. dry and steam reforming)
The reforming process, which involved the reaction of methane whereas pyrolysis and gasification processes for the biomass-based
gas derived from biomass or natural gas with carbon dioxide or production option. The increasing global interest on the production
water, is achieved catalytically at normally high reaction temper- and valorization of shale gas (Bruijnincx and Weckhuysen, 2013;
atures, under controlled pressure conditions to produce hydrogen- Gaboriaud et al., 1991; Bowker, 2007; King, 2010), therefore imply
rich synthesis gas (Edwards and Maitra, 1995; Johnsen et al., 2006). that the syngas production will attract future shifts towards the
The process was initially viewed as a future alternative for the shale gas utilization as the main feedstock. Currently, the produc-
production of clean hydrogen fuel. But recent studies, indicated a tion of shale gas takes place mainly in the countries like the United
reasonable shift to its linkage as a background route for producing States, Canada and countries in the North America. Although large
synthetic raw materials required by the energy and petrochemical recoverable deposits have been technically discovered in the Asia
industries, because the gaseous products could be upgraded to Pacific, the commercialization is still at the low scale. There are
fuels and petrochemicals, especially via the FishereTrospch reac- indications that, the shale gas is gradually replacing the regular
tion (Surisetty et al., 2012; Hoek et al., 1985; Dry, 2002). Therefore, natural gas in the former regions, with production expected to rise
methanol is one of the key industrially-derivable petrochemicals in the future years. Among the current market options (Fig. 2),
from the reforming technology. Although methanol had been transportation industry had been identified as the fastest growing
employed previously for a variety of applications, covering the market for the future years (Rearch, March 2014). The overall global
production of pigments, plastics and paints, as solvent, wastewater market had been projected to rise by 5.3%, worthy of at least 9.19
denitrification, biodiesel production and in electricity generation billion US dollars by the year 2020.
by driving turbines (Kaeding and Butter, 1980; Pretzer and Dry reforming proceeds via the interaction of methane with
Kobylinski, 1980; Hamnett, 1997; McNicol et al., 1999), the cur- carbon dioxide (Fig. 3), at usually high temperatures that could be
rent status of the energy industry gave a considerable emphasis to up to 800  C, over reforming catalysts like the Ni based systems
hydrocarbon production. The methanol to gasoline (MTG) tech- (Pakhare and Spivey, 2014). Numerous studies have been con-
nology is a forefront catalytic route under exploration. For these ducted on this reforming technology, therefore the paper will
reasons, the global methanol production is reasonably high (Fig. 1) present details on key recent reports. Yu et al. (2015) showed Ni/
(The University of York, 2014), and the demand had been predicted CeO2 catalyst as an active catalyst for the reforming reaction at
to be on the increase (Herder and Stikkelman, 2004; Olah, 2005). 760  C and 0.1 MPa pressure, but the catalyst deactivates with time
The paper therefore presents a review on the MTG process with due to massive coke deposition. The authors further demonstrated
emphasis to updates on the catalyst systems under consideration. that, modification with Ag as a promoter favors both the yield of
Important parameters such as the reaction conditions and reaction synthesis gas and the catalyst lifetime by improving resistance to
mechanisms have also been adequately reviewed. Recent updates coke production. The Ag particles hindered the nucleation and
on syngas production and methanol preparation from syngas were growth of carbonaceous materials at the susceptible Ni sites and
also initially discussed. enhanced the gasification of already deposited coke. In a related
development, Wang et al. (2015) found the incorporation of CaO
into Ni/ZrO2 catalyst as another important option for limiting
2. Methanol from syngas (CO, H2) catalyst deactivation by coking. Similarly, parameters such as con-
centration of methane in the reactants feed, catalyst morphology
2.1. Syngas production and reaction temperature must be appropriately considered. At
high temperatures, feed richer in methane are associated with
The modern methanol production technologies involved two deposition of coated-form carbon deposits, with strong potentials
major catalytic processes; the production of synthesis gas followed to cover the active Ni particles and reduce catalyst lifetime (i.e.
stability). Although conversion could reach nearly 100% with closer
values of syngas selectivity, catalyst deactivation by coking is a
great challenge with dry reforming. Therefore majority of the

Fig. 1. Global production of methanol by the year 2014. Data source: The University of
York (2014). Fig. 2. Current global markets for shale gas, 2014. Data source: Rearch (March 2014).
A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316 305

Fig. 4. Equations for the steam reforming of methane.

Fig. 3. Chemical equations for catalytic dry reforming of methane with carbon dioxide. natural gas feedstock at lower reaction temperatures like 500  C by
utilizing pure feedstock and appropriate catalyst system. They
employed Ni and Rh catalysts supported with La modified
documented literature targeted affordable ways of handling the
CeO2eZrO2 system. Both the two catalysts were very stable to
problem (Courson et al., 2000; Guo et al., 2004; Laosiripojana and
deactivation by carbonaceous deposition, especially with the Rh
Assabumrungrat, 2005; Tsyganok et al., 2003; Djaidja et al.,
catalyst for 10 h period. However, the authors showed the presence
2006). Recently, Albarazi et al. (2014) reported the nature of cata-
of ethane and propane in natural feedstock as unfavorable. At low
lyst preparation method as another key influential factor. They
reaction temperatures, these species promoted the accumulation of
prepared a Ni/SBA-15 catalyst modified with CeO2eZrO2 composite
carbon deposits, especially with the Ni catalyst. The net conse-
using two key methods (i.e. co-precipitation and impregnation
quence was lost of the catalyst lifetime and syngas selectivity. Ma
methods). The co-precipitated catalyst was less-active with short
et al. (2014) investigated the role support material for the Ni cat-
lifetime. Preparation by impregnation produced more active cata-
alysts. Under constant conditions, catalyst supported with g-Al2O3,
lyst, due to enhanced interaction between the Ni/SBA-15 and
was active but deactivates more rapidly compared to the catalyst
CeO2eZrO2 dopant. The catalyst which allowed direct reaction of
supported with either La2Zr2O7 or La2Sn2O7 material. The latter
methane at the active surfaces, was richer in NiO species, and
catalysts does not produce any coke deposits, but the catalyst
showed enhanced catalyst stability with time. On their own part,
containing La2Zr2O7 displayed extremely higher performance to-
Jabbour et al. (2015) showed the preparation of Ni supported
wards syngas production, thus indicating its industrial potentials.
dolomite catalyst by deposition method to produced highly active
In a similar trend, Gubareni et al. (2014) showed the addition of
and stable catalyst. The catalyst containing 5 wt.% of Ni was more
ceria or lanthanum oxide to Ni catalyst supported with Al2O3/
promising than a comparable Ni/SiO2 system under constant con-
cordierite, to significantly suppressed the wateregas shift reaction
ditions. At a temperature of 800  C, it yields 90% methane con-
and consequently reduced the amount of CO2 in the net syngas.
version with nearly 100% selectivity to syngas (i.e. H2:CO ¼ 1). This
This behavior in turn enhanced the syngas selectivity. The modifi-
behavior was not noticed with the Ni/SiO2 catalyst due to rapid
cation was very positive in decreasing the water to methane feed
catalyst deactivation. Interestingly, the former catalyst was stable
ratio. Generally, good steam reforming catalyst stability, activity
for 12 h when the reaction temperature was lowered to 650  C.
and selectivity to syngas could be achieved by appropriate catalyst
Similarly, the minor carbonaceous deposits were easily removed by
modification and selection of reaction conditions (Palma et al.,
regeneration and full activity restored.
2014; Zhang et al., 2014; Vita et al., 2015; Zyryanova et al., 2014).
An alternative to the dry reforming technology is the steam
Pyrolysis and gasification processes are the catalytic options
reforming process. Unlike the dry reforming process which employ
normally employed for the production of syngas from biomass-
CO2 as the co-reactant, steam reforming proceeds by an endo-
based feedstock. The phase-change associated processes involved
thermic reaction of purified methane feed with steam at compa-
the decomposition of biomass-based feedstock thermos-
rable temperatures. The technology is widely utilized for hydrogen
chemically to produce syngas and sometimes liquid reaction
gas production, especially in the United States, where the produc-
products, depending on the reaction parameters employed. The
tion was believed to hit 95% in the recent years (Hoffmann, 2012).
reactions are carried out also at high temperatures, especially the
Some authors described it as the most cost-effective technology for
gasification process that is mainly known for the production of
syngas production (Crabtree et al., 2004; Clarke et al., 1997;
syngas and carbon dioxide as the reaction products. Numerous
Achenbach and Riensche, 1994). The reaction chemistry involved
studies have also been conducted on these commercial processes
the interaction of methane and steam in a catalytic reactor at
(Bridgwater, 1995; Sutton et al., 2001; Tijmensen et al., 2002;
usually 850  C for achieving desired conversion. It comprised of a
Raveendran et al., 1996, 1995). Recently, Xie et al. (2012)
wateregas shift reaction in the subsequent step, by which CO reacts
employed Ni-based catalysts for simultaneous pyrolysis and gasi-
with steam to produce more H2 gas (Fig. 4). The most active cata-
fication of sawdust, that was rich in cellulose with low ash content.
lysts are those considered to be stable under these conditions, with
The feasibility of the process depended on the working tempera-
potentials to generate large quantities of reaction products and
ture. While the pyrolysis produced optimal syngas production at
cost-effectively regenerable.
750  C, the temperature was further raised to 850  C to derive
Like the dry reforming, numerous studies have also been carried
sufficient syngas selectivity via gasification. The authors demon-
out on the steam reforming, with the view of achieving enhanced
strated the combined approach as more beneficial for syngas pro-
syngas production and catalyst stability properties. Recently, Angeli
duction than when either pyrolysis or gasification was employed
et al. (2014) demonstrated that, the reaction can be achieved with
alone. However, Ni catalyst deactivation was also predominant as
306 A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316

normally observed with the reforming reactions. From their nu- CeO2 catalyst produced the best syngas conversion and methanol
merical modeling studies, Ravaghi-Ardebili et al. (2015) showed selectivity that was very stable for 100 h. They attributed the ac-
that, the gasification temperature can significantly be reduced, tivity to the oxidizable behavior of hydrogen sulfide over this
achieving high syngas selectivity, by employing integrated steam catalyst. Cu supported with ZnO catalyst yields the worst stability
supply. They demonstrated that, an equilibrium must appropriately properties, because its activity rapidly decayed after 7 h, due to high
be established for parameters such as residence time of feedstock in susceptibility to trapping hydrogen sulfide. On the other hand, Pd/
the reaction chamber, particle sizes of the biomass-based feed- Al2O3 catalyst showed only a reduction of 5.3% activity after the
stock, moisture content and the quantity of steam supplied. 100 h of reaction. Catalysts containing Pd supported with oxides or
Recently, Kim et al. (2015) found the choice of biomass-based other materials have been identified as good candidates for the
feedstock as a favorable alternative for lowering the pyrolysis SGTM and related processes (Shen et al., 2001; Matsumura et al.,
temperature. They showed the residues an algal specie, to 2001; Sudhakar and Vannice, 1985). Chu et al. (2013) demon-
decomposed at temperatures as low as 200e500  C, producing strated that, the activity and stability of Cu-based catalysts can be
reasonable quantities of syngas. However, production of bio-oils enhanced by the appropriate choice of precipitating agent. They
may also be dominant at certain low temperatures. employed tetraethyl ammonium hydroxide as precipitating agent
for the synthesis of Cu/ZnO/Al2O3 catalyst, and compared its per-
2.2. Methanol synthesis formance with equivalent material prepared using polyethylene
glycol as the precipitating agent. Syngas conversion was carried out
Following its successful production, the syngas is employed as at 230  C and 10, 000 h1 using CO to H2 ratio of 15:70. Their results
an affordable feedstock for the production of high grade methanol. indicated the former precipitating agent to yield the best catalyst
At the industrial scale (Fig. 5), purified syngas, usually containing system. Smaller crystallites of Cu, with associated large BET surface
2e2.5 M ratio of hydrogen to carbon monoxide, is employed for the area, enhanced methanol conversion and stability were produced.
reaction at 3000 to 5000 psi pressure. This extreme pressure per- The bigger Cu crystallites formed using polyethylene glycol pro-
mits the compression of the reactant species. Under appropriate duced 5.2% activity decay for the 20 h. Due to the growing interest
space velocity, the mixture can be fed into the reactor at temper- in the production of methanol from syngas, the Cu-based catalysts
atures in the range of 300e400  C. The exothermic nature of the are considered the most cost-effective compared to the Pd or other
reaction implies that, sufficiently high pressure is desired to shift noble metal-based catalysts. Modification or selection of appro-
the equilibrium towards methanol production. Under the synthesis priate preparation route is very critical to designing effective
conditions, the exit products composed of gaseous methanol in a catalyst system. Some authors (Spencer, 1998; Wang et al., 2010),
mixture with carbonyl impurities, thus must be condensed at indicated the incorporation of ZnO to enhance the activities of Cu-
pressures ranging between 3000 and 4000 psi. The liquid methanol based catalysts due to the improved synergic effect. Promotion with
must then be carefully depressurized and adequately purified to alumina can produce two unique properties; it retards and protect
eliminate unwanted impurities such as aldehydes, ketones and the active catalyst particles from agglomeration and also enhances
dimethyl ether, whose concentration is usually in the range of carbon monoxide adsorption and activation (Liu et al., 2003). In
2e5%. Various catalyst systems are employed by the industries for addition to alumina, oxides of Mn and Zr have been reported as
this reaction (Wender, 1996; Rostrup-Nielsen, 2000; Klier, 1982). good promoters for the Cu-based catalysts (Meshkini et al., 2010;
Among them, mixed oxide systems of alumina, magnesium oxide Samei et al., 2012). Another important option is the preparation
and other transition metal oxides are commonly utilized. of the catalysts to the nano-scale (Hu et al., 1996a, 1996b). This
In a trend similar to the syngas production, different in- permits their adsorption and activation properties towards CO to be
vestigations have been carried out on the suitable catalysts and enhanced and also promotes the selectivity to methanol than other
reaction conditions for the syngas to methanol (SGTM) process. Ma side reaction products. In most scenarios, the modified catalysts are
et al. (2008) studied the sulfur tolerance potentials of transition more resistance to deactivation than the unmodified analogs.
metals supported with oxides as catalysts for the SGTM reaction. Due to the current global potentials of shale gas, it is increas-
Syngas conversion was performed at 240  C, using a feed compo- ingly being considered as replacement to the natural gas, for the
sition ratio of ~2 for hydrogen to carbon monoxide at a space ve- production of syngas and consequently methanol. Ehlinger et al.
locity of 1000 h1 and 3.0 MPa pressure. The authors employed (2013) reported a possible commercial project for the Barnett
30 ppm of hydrogen sulfide gas for the deactivation studies. Pd/ shale gas fields in the United States. However, direct processing of

Fig. 5. Typical industrial unit for the production of methanol from synthesis gas.
A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316 307

shale gas in the existing conventional natural gas equipment must dioxide conversion was found to increase when the reaction tem-
be carried out with substantial care. Due to its compositions, the perature was increased. The initial conversions of 3e5% increased
shale gas must be cleaned-off impurities to remove all the un- to 25e40% when the temperature was raised from 460 to 600 K.
wanted acidic and non-acidic gases, with potentials to create dif- The rate of methanol production increased also linearly with Ga
ficulties, before the upgrading processes. One interesting thing is loading. For example, the catalyst containing 3.0 wt.% Ga produced
that, methanol can be produced from other important alternative a rate that was 30% higher than the rate found with the Ga-free Cu/
ways, closer to the syngas technology. Reduction of carbon dioxide ZnO/ZrO2 catalyst. However, beyond the optimal temperature of
is among these options (Olah et al., 2008; Arab et al., 2014). The 250  C, the effect of Ga modification slightly declined. Generally,
process involved direct reaction of carbon dioxide with hydrogen to addition of a modifier, appropriate choice of support or synthesis
produce methanol and water. The reaction is normally carried out procedure permits the control of catalyst (Toyir et al., 2009) basicity
at 260e300  C, to obtain excellent methanol selectivity (up to 99%) by increasing the density of equilibrated strong basic sites, pro-
(Joo et al., 1999; Goehna and Koenig, 1994). One key important motes synergism, retards the rate of catalyst deactivation, prevents
benefit is the limitation in the production of large quantities of side reaction interference and therefore promotes the overall
carbon dioxide that may be associated with the syngas option (Tao catalyst selectivity to methanol (Surisetty et al., 2011; Guo et al.,
et al., 2013). Similarly, the environmental challenges associated 2011; Collins et al., 2005; Saito et al., 1996; Fujitani et al., 1995).
with carbon dioxide implied the technology as a good mitigation
option. There are some literature evidences that, the technology 3. Methanol to gasoline (MTG)
had been used by methanol production companies in the United
States during the 1920's and 1930's (Cheng, 1994). Catalysts iden- Modern processes for the conversion of methanol to gasoline
tified as promising candidates for the reaction are similar to those gave emphasis to the methanol production from syngas before
developed for the syngas method (Saito and Murata, 2004; Saito, subsequent upgrade to gasoline in the MTG operation unit. Unlike
1998). They are primarily catalysts based on copper, zinc or other other synthetic methods for gasoline production, the methanol to
transition metals and their oxides (An et al., 2007; Gao et al., gasoline route produce gasoline with compatible octane properties
2013a). that is also free from impurities. At the industrial scale, the MTG
Recently, Cai et al. (2015) demonstrated the role of catalyst process, which can be achieved at complete conversion, is
synthesis procedure and support nature on the catalyst activity exothermic with heat of reaction of 1.74 MJ/kg methanol (Yurchak,
during this reaction. They employed a Cu catalyst supported with 1988). To successfully handle the reaction, two splitting steps are
ZnO and Ga2O3, prepared from microwave-assisted process. normally adopted. Initially, methanol gets transformed into water
Through this method, Cu dispersion and associated interaction with and dimethyl ether equilibrated with unreacted methanol over a
support materials were enhanced. Similarly, the catalyst basicity catalyst that is usually non-zeolitic in nature. This stage eliminates
was adequately improved. The catalyst produced 50% selectivity to 15% of the heat. In the final/second stage, a zeolite catalyst like H-
methanol production for a reaction conducted at 250  C, employing ZSM-5 is used to convert the mixture with a recycled gas to gasoline
1:3 ratio of carbon dioxide to hydrogen and 3000 h1. The catalyst range hydrocarbons and water. Several MTG commercialization
yields constant stability under these optimal conditions. Li et al. units are considered in different parts of the world. The petroleum
(2015) reported a similar study. Incorporation of surfactant was giant Mobil, commercialized a plant in New Zealand in 1987. The
very beneficial in designing CuO/ZnO/ZrO2 catalyst than the operational unit comprised of a syngas production facility, which
normal/conventional co-precipitation process. The surfactant- converts natural gas from field sources to methanol. The 14,500
assisted method ensured the production of homogenized Cu- barrels per day MTG unit uses fixed-bed configuration to generate
support crystallites, with adequate basicity and strong resistance clean gasoline with properties similar to those of the conventional
to deactivation. While the conventional catalysts produced meth- fuel.
anol selectivity of 39% as the maximum, the former catalysts pro- TOPSOE technology (i.e. TIGAS process) was also one of those
duced up to 73% stable selectivity. For these catalysts the selectivity early commercial MTG processes (Topp-Jørgensen, 1988). Recently,
increased from 54 to 73%, when the calcination temperature was the Exxon Mobil and Uhde Corporation escalated the New Zealand
raised from 300 to 600  C. Oyola-Rivera et al. (2015) reported technology to planned new plants in the United States (i.e.
deactivation properties with Pd catalysts. The authors used poly- Wyoming and West Virginia) (NETL, 2014). The Wyoming facility
morphic forms of gallium oxides as support materials during the was designed to a capacity of 20,000 to 22,000 barrels of gasoline
methanol production. Deactivation was mainly caused by loss of per day. It will use coal as the source of primary gases for the MTG
catalyst basicity. Support that allowed high basicity retention were process. The project planned to commence operation in 2016 or
less susceptible to deactivation. For instance, Pd/a-G2O3 was more beyond due to technical arrangements, is also expected to comprise
stable than Pd/a, b-G2O3, due to the sudden basicity lost associated of carbon capture facility for commercial CO2 sequestration process.
with the latter catalyst. Gao et al. (2013b) showed the incorporation Another commercial proposal by Transgas Development Systems
of fluorine to affect the catalyst performance. Addition of fluorine to (TGDS) aimed at building a similar facility worthy of 3 billion US
a Cu/Al/Zr catalyst improved the densities of strong basic sites with dollars, with possibility of annual gasoline production of 6.5 million
capabilities for methanol production. Without fluorine, the strong barrels. The TGDS Company licensed to Uhde Corporation, expected
basic sites were significantly low in concentrations, leading to low the plant to uses similar syngas from coal gasification technology to
methanol selectivity. When fluorinated, weak basic sites were generate methanol and consequently gasoline using the Exxon-
significantly converted to strong basic sites, shifting the catalyst Mobil MTG process. The operation period had also been slated late
selectivity towards methanol production. Addition of fluorine had 2016. Several of these facilities are being planned across the globe.
been found very positive for this reaction by some other authors Therefore, triggering research on the most suitable catalyst systems
(Gao et al., 2013b, 2014). Ladera et al. (2013) studied the modifi- that can provide economic yields of gasoline under affordable and
cation of Cu/ZnO/ZrO2 with Ga. Influenced of Ga loadings from 0.2 cost-effective conditions.
to 3.0 wt. % was investigated. Addition of Ga showed a slight The actual reaction mechanism for the production of paraffins
decreased in BET surface area of the parent catalyst. Without and other hydrocarbon products from methanol is still a debatable
modification, the surface area was 129 m2/g but reduced to 102 m2/ issue in the literature. Among the proposed mechanisms, direct
g when the Ga loading reached 3.0 wt.%. For all the catalysts, carbon coupling of C1 species, methylation of alkenes and their subsequent
308 A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316

cracking (i.e. olefins methylation and cracking pathway) and the irrespective of the metal or other parent catalyst, the reaction
hydrocarbon pool pathways involving methylbenzene in- products composed of dimethyl ether and hydrocarbon com-
termediates have been critically studied (Marcus et al., 2006; pounds. Side reaction products such as CO and CO2 were produced
Dessau, 1986; Dessau and LaPierre, 1982; Dahl and Kolboe, 1996; to only very negligible concentrations. Similarly, the composition of
Mole et al., 1983). In the last two routes, chain growth and subse- aromatic species was very low. In all cases, the metal salts of TSA
quent cracking are usually involved. These mechanisms are iden- and TPA were more active than the parent systems, among which
tified as more favorable for the production of mainly olefins and the Ag containing catalysts produced the best activities. (98% for Ag
aromatics. Fig. 6 represents a general scheme for the production of containing TPA and 79% for Ag containing TSA). Although propane
all hydrocarbon types (Wang et al., 2013). It could be seen that, was produced to significant selectivity, butanes and higher paraf-
gasoline range paraffins and alkenes with possibility of hydroge- fins dominated the reaction products. The authors argued that, the
nation to gasoline compounds can be produced through mecha- reactions of pre-generated olefins with methanol during the in-
nistic pathways involving deprotonation of carbenium ions, termediate stages, is slower with these compounds compared to
isomerization, cracking and hydride transfer. The carbenium ions when zeolite systems are used. Recently, Dagle et al. (2014) studied
are fundamentally derived from the interaction of methanol with the activity of a bifunctional catalyst (Pd/ZnO/Al2O3-ZSM-5) as
Brønsted acid sites of a catalyst. Therefore the densities and catalyst for this reaction. Syngas was used directly for conversion to
strengths of these sites are very critical factors to be considered in gasoline hydrocarbons. The mechanism of action of the catalyst
designing any catalyst for the MTG reactions. proceeds by sequential methanol production from syngas, its
dehydration over the acidic sites of ZSM-5 to form dimethyl ether
3.1. Non-zeolite catalysts and the subsequent ether conversion to gasoline range hydrocar-
bon species. Role of parameters such as temperature and pressure
There are literature evidences that non-zeolitic catalysts have were investigated. At a temperature of 375  C, increasing the re-
been previously evaluated as candidates for the MTG reaction. As action temperature had a positive effect on syngas conversion. For
early as 1978, Kim et al. (1978) showed the bulk zinc iodide as good example, raising the pressure from 300 to 500 psi produced an
catalyst for the production of butane and higher paraffins from increased in conversion from 75 to 85%. The higher pressures were
methanol. They conducted reactions at 200  C and 200 psi for a therefore very critical for initial methanol production with this
period of 2 h. Methanol conversion was nearly complete (>99%) and catalyst. However, the distribution of hydrocarbon products had
the catalyst actively produced mainly branched gasoline com- been only slightly affected by change in pressure. Lighter hydro-
pounds of high octane number. Triptane (i.e. 2,2,3- carbons (C1 to C4) dominated the reaction products at all pressures,
trimethylbutane) was produced to selectivity closer to 50%. Simi- with selectivity close to 90%. Gasoline range hydrocarbons were
larly, lighter cracking species (i.e. with carbon number below only detected to low concentration of 6e10%, with significant
butane) were found only in trace quantities. The authors also amount of aromatics. On the other hand, increasing the reaction
demonstrated zinc bromide to produced comparable activities temperature at 300 psi, shifted the conversion until an equilibrium
when the temperature was raised to the range of 220e245  C. was established, beyond which the conversion declined. Unfortu-
However, with zinc chloride triptane selectivity decays due to nately, the selectivity to lighter species was more pronounced at
higher reaction temperature required. Compounds of heteropoly higher temperatures, limiting any selectivity to gasoline range
acids like the tiungstosilicic acid (TSA) and tungstophosphoric acid products. The overall reaction results indicated the composite
(TPA) are another category of catalysts fully studied their behaviors catalyst as more selective to the production of lighter hydrocarbons
have been attributed to their Keggin structure and Brønsted acidity (C1 to C4) and polymethylbenzenes than gasoline range paraffins.
(Baba et al., 1982a; Misono, 1987; Niiyama et al., 1982). Baba et al. Other non-zeolite catalysts employed include those based on
(1982b) employed metal salts of TSA and TPA, containing variable supported transition metals like cobalt and rhenium and ammo-
metals including La, Cd, Ag and other non-transition metal species nium 12-tungstophosphates (Hayashi and Moffat, 1983; Mauldin,
like Cs and K. Methanol conversion studies were performed using 1986). They are known for the production of C10 and linear al-
3 g of catalyst at 300  C and a methanol flowrate of 0.0212 mol/h. kanes in addition to olefin compounds (Mauldin, 1986). Majority of

Fig. 6. A representative scheme for the production of different hydrocarbon products from methanol upgrading.
A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316 309

the non-zeolite catalysts employed for gasoline synthesis from and H-ZSM-48, all the catalysts produced significantly high meth-
methanol or directly from the syngas have so far received limited anol conversions. The deactivation was attributed to be due to coke
industrial consideration due to some vital reasons. Difficulty in formation, blocking mainly the pores of H-EU-1 and H-ZSM-48.
achieving excellent yield of gasoline range paraffins due to asso- With the exception of H-EU-1, all the other catalyst systems were
ciated side reactions is a forefront issue. Acidity control and catalyst highly selective to C5 and other gasoline range hydrocarbons.
deactivation with time due to coke deposition or active sites However, H-ZSM-48 and H-EU-1 catalysts yield reasonable selec-
poisoning is another great challenge. Compounds based on halides tivity to aromatic compounds. These results therefore indicated
(like zinc iodide) are corrosive and difficult to be regenerated. Other that, formation of aromatic species is possible even with one
issues include sintering and environmental disposal problems. dimensional zeolites. Thus, appropriate adjustment of reaction
These challenges accounted for a shift to the zeolitic and analog parameters or modification is necessary with either the single or
catalysts. multiple pore systems to avoid aromatics formation or high selec-
tivity to lighter (C1 to C3) products. Zaidi and Pant (2004) con-
3.2. MTG with zeolite catalysts ducted a study with modified H-ZSM-5 catalysts at 400  C and
0.1 MPa. Catalysts were modified by the incorporation of 7 wt.% of
Zeolite catalysts are the primary catalyst systems employed by ZnO, CuO or the composite CuOeZnO to the parent H-ZSM-5 ma-
the companies for the MTG reactions. Their behaviors are based on terial with a Si/Al ratio of 45. While the catalysts pore volume
shape-selectivity, dimensional structure, stability and acidity remain closely the same, the BET surface area of the parent material
properties with possibility of modifications under controlled con- decreased from 290 to 244, 255 and 241 m2/g for the ZnO/H-ZSM-5,
ditions. Catalysts based on H-ZSM-5 or the modified analogs were CuO/H-ZSM-5 and CuO/ZnO/H-ZSM-5, respectively. Modification
the main materials given preference by the Companies, with with the oxides enhanced catalyst resistance to deactivation with
numerous literature studies documented (Bjørgen et al., 2007; time and influenced both conversion and products distribution.
Svelle et al., 2006; Anderson and Klinowski, 1990; Choudhary Without modification, the conversion was 38%, but increased to up
et al., 2005; Rownaghi et al., 2011; Saxena et al., 2014). to 97% for the modified zeolite catalysts. Although paraffins and Cþ 5
alkenes were detected, the production of aromatic species
3.2.1. Effect of zeolite topology and modification increased with modification. For the parent H-ZSM-5, the selec-
The structural and compositional properties or the modification tivity to aromatics was 21% but increased to 67, 69 and 77% for the
strategies adopted are very critical factors for achieving satisfactory ZnO/H-ZSM-5, CuO/ZnO/H-ZSM-5 and CuO/H-ZSM-5 catalysts,
yields of gasoline range hydrocarbons. Teketel et al. (2009) con- respectively (see Fig. 7). The reaction mechanism was believed to
ducted a study with H-ZSM-22 catalyst (Si/Al ¼ 30). The catalyst consist of two key steps consistent with those explained in Section
with BET surface area of 173 m2/g produced methanol conversion 3 above. Primarily, methanol undergo dehydration into dimethyl
up to 100% at 500  C and 2.05 h1. Interestingly, replacing this with ether. Whereas in the secondary stages, the equilibrium mixture
a similar material of higher BET surface area (173 m2/g) yields produces light olefins that are susceptible for conversion to higher
comparable conversions at lower temperatures. Although deacti- hydrocarbon products (paraffins and aromatics or even the higher
vation was observed with time, the catalyst remarkably produced olefins). It could be established that, the oxides doped catalysts
normal and branched alkenes, with possibility of hydrogenation to favor coupling and cyclization reactions. The mechanism of coke
gasoline range paraffins. The content of aromatics had been deposition during the MTG reaction is similarly dependent on the
attributed to non-intersecting 10 MR pore system of the H-ZSM-22 zeolite topology.
catalyst. Therefore, alkene methylation and subsequent cracking is Dejaifve et al. (1981) showed that, with catalyst like H-ZSM-5,
the main reaction mechanism with this catalyst. Several authors carbonaceous materials are mainly deposited on the outer surface.
have attributed the zeolite structure to influence the products This lead to slight alteration to the shape-selective behaviors, with
distribution (Olsbye et al., 2012). Zeolites with 2 or 3-dimensional net resistance to deactivation. On the other hand, zeolites such as
pore systems are good candidates for the production of aromatic mordenite and offrerite showed different depositional mode due to
species (i.e. both mono- and polycyclic aromatics). A work by Sassi their larger channels. They accommodated carbonaceous deposits
et al. (2002) demonstrated that, over an H-BEA zeolite, bicyclic inside the pores, and the zeolites are therefore more susceptible to
aromatic compounds can be produced to significant selectivity activity decay than the H-ZSM-5 catalyst. Similar results have also
from an initially formed benzene ring by simple coupling of C3 been reported Svelle et al. (2007). In a related development, Haw
hydrocarbon substituents pre-generated. Thus, under this situa-
tion, the polyaromatics can dominate the reaction products. In a
related development, Bjørgen et al. (2003) reported the existence of
cyclization intermediates, comparable to the heptamethyl benze-
nium ion over the H-BEA zeolite. The authors introduced hexam-
ethylbenzene into the reaction feed and found intermediates such
as the dihydromethylnaphthalene. In another study, Bjørgen et al.
(2010) documented an evidence of the formation of polyaromatic
species, including trimethylnaphthalenes over H-BEA and H-MCM-
22 zeolites. Thus, the structural features of these types of catalysts
may be negative toward the production of gasoline range paraffinic
products or the corresponding olefins with potentials to be hy-
drogenated compared to the one-dimensional zeolite systems.
The effect of one dimensional zeolite was recently reported by
Teketel et al. (2011). They used 10 MR zeolites (i.e. H-EU-1, H-ZSM-
22, H-ZSM-23 and H-ZSM-48) as model catalysts. Catalytic re-
actions were conducted under variable conditions of temperature
(350e500  C) and space velocities (2e6 h1). Although catalyst Fig. 7. Effect of H-ZSM-5 catalyst modification on selectivity to aromatic hydrocarbons.
deactivation was encountered with time, especially with H-EU-1 Reaction conditions; 400  C and 0.1 MPa. Data source: Zaidi and Pant (2004).
310 A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316

et al. (2003) showed the topological difference between H-Fer- to deactivation, than the conventional catalyst under constant re-
rierite and H-ZSM-5, of comparable acidity, to yield a remarkable action conditions, conversion and closer selectivity.
difference in terms of products distribution. The former catalyst
produced mainly butenes and pentenes. It does not favor cycliza- 3.2.2. Effect of zeolite acidity properties
tion or the consequent production of methylbenzenes. On the other The density and strength of either Brønsted or Lewis acid sites
hand, the channel structure of H-ZSM-5 was susceptible to permit play an important role during acidity-dependent reactions like the
the reaction to proceeds by cyclization and possibly hydride MTG process. Bjørgen et al. (2008) reported increasing the H-ZSM-
transfer. Under this condition, olefins can easily be transformed 5 zeolite acidity, be desilication with 0.05e0.20 M NaOH, to
into alkanes and aromatic products. enhanced catalytic activity. It was observed that, the most severely
Table 1 presents the activities of some zeolites during the MTG desilicated catalyst produced optimal conversion that was 3.3 times
reaction. According to He et al. (2013), treatment of a nano- more stable than for the untreated catalyst. On the other hand, the
crystalline H-ZSM-5 with tetrapropyl ammonium hydroxide overall selectivity to gasoline hydrocarbons multiplied by 1.7, due
(TPAH) can modify the catalytic properties of the parent catalyst, to increase in the speed of hydrogen transfer reactions. Benito et al.
with the effect being dependent on the treatment period. Irre- (1996) conducted a related study with H-ZSM-5 catalysts of Si/Al
spective of the treatment period, the selectivity to gasoline range ratio in the range of 24e154 and comparable crystallinity. Their
hydrocarbons was nearly constant at 45%. However, catalyst sta- results indicated increasing the Si/Al ratio to drastically decrease
bility was favored by the TPAH treatment. For a 24 h treatment the total zeolite acidity. When the Si/Al was 24, the total acidity was
period, the stability of conversion increased by 100h. A further 0.5 mmol/g but decreased to 0.10 mmol/g when the Si/Al ratio was
increased in the treatment time to 48 h further increased the sta- raised to 154. However, the Brønsted to Lewis acidity ratio (B/L)
bility by 170 h, but the conversion under this condition was slightly increased from 2.86 to 4.42. For the purpose of comparison, the
below 98%. On the other hand, Di et al. (2013) showed reaction with MTG reactions were performed using 0.01 g of catalyst at variable
H-ZSM-5/H-MCM-48 composite catalyst to produced higher temperatures (383e448 K). The results indicated Brønsted acid
selectivity (34%) to gasoline range products than when the reaction sites as the most important acid sites for the reaction. The selec-
was carried out with H-ZSM-5 catalyst alone (28% selectivity). tivity to gasoline hydrocarbon increased with increase in the B/L
Under constant conditions, the composite mixture limits the ratio. The authors also demonstrated ethene to butene as the pri-
chance of producing aromatic coke precursors and C2 to C3 hy- mary reaction products, and therefore reported decomposition and
drocarbon species. Bjørgen et al. (2009) compared the effect of propagation of oxonium ions as the main reaction mechanism. The
zeolite nature. While the methanol conversion doubled for H-ZSM- role of Brønsted acid sites was also demonstrated by Wang et al.
5 (Si/Al ¼ 50) compared to H-BEA catalyst, the selectivity was also (2003) They employed H-Y and H-ZSM-5 zeolites and showed
higher by 7%. Thus, the H-ZSM-5 catalyst formed a better system that, methanol interacts with these acid sites at the initiation stages
under the employed reaction conditions. Treatment of ZnO/CuO/H- to produce methoxy groups, which in turn can undergo further
ZSM-5 with oxalic acid shifted the selectivity of the parent catalyst reactions to yield hydrocarbon. There are various reaction possi-
by 4% at comparable conversion (Zaidi and Pant, 2008). Therefore, bilities under this circumstance. The methoxy groups have the ca-
careful catalyst modification can be very beneficial for the gasoline pacity to react with water to methanol at low reaction
hydrocarbons production. Lee et al. (2010) demonstrated different temperatures. They can also interact with alkanes or aromatic
behaviors for the H-UZM-12 and H-SSZ-13 zeolites. The former compounds in the reaction feed by the process of methylation. The
catalyst yields 60% methanol conversion that was stable for only reaction results further showed that, at higher reaction tempera-
4 h. But the H-SSZ-13 produced 100% conversion to higher lifetime tures beyond 250  C, methoxy groups can undergo direct conver-
(8 h). However, the H-UZM-12 catalyst was more selective to the sion to the gasoline hydrocarbons. In view of these observations,
desired hydrocarbon products. A further study on the H-SSZ-13 by the hydrocarbon pool pathway remain the main reaction mecha-
Wu et al. (2013) showed a hierarchical catalyst as the most resistant nism. The key aspect of this mechanism is the reaction of methanol

Table 1
Effect of zeolite topology or modification on MTG reaction properties.

Catalyst Reaction conditions Methanol Selectivity/yield Ref.


conversion, % of gasoline range
hydrocarbons (Cþ 5 ), %

Nano-crystalline 1.0 g catalyst mixed with 1.0 g alumina, 380  C, 2.0 h1 98, stable for 70 h 45 He et al., 2013
H-ZSM-5
Nano-crystalline 1.0 g catalyst mixed with 1.0 g alumina, 380  C, 2.0 h1 98, stable for 170 h 46 He et al., 2013
H-ZSM-5, treated with tetrapropyl
ammonium hydroxide for 24 h
Nano-crystalline 1.0 g catalyst mixed with 1.0 g alumina, 380  C, 2.0 h1 90, stable for 240 h 45 He et al., 2013
H-ZSM-5, treated with tetrapropyl
ammonium hydroxide for 48 h
H-ZSM-5 1.40 g catalyst, 380  C, 1.0 MPa, 2.0 h1 98 28 Di et al., 2013
H-ZSM-5/H-MCM-48 mixture 1.40 g catalyst, 380  C, 1.0 MPa, 2.0 h1 97 34 Di et al., 2013
H-ZSM-5 (Si/Al ¼ 50) 0.06 g catalyst, 350  C, 7 h1, 15 min reaction 95 47 Bjørgen et al., 2009
H-BEA 0.06 g catalyst, 350  C, 7 h1, 15 min reaction 55 40 Bjørgen et al., 2009
ZnO/CuO/HZSM-5 400  C, 0.1 MPa, 0.129 g of catalyst$h/g of methanol fed, 18 h 95 08 Zaidi and Pant, 2008
ZnO/CuO/HZSM-5, modified 400  C, 0.1 MPa, 0.129 g of catalyst$h/g of methanol fed, 18 h 94 12 Zaidi and Pant, 2008
with oxalic acid
H-UZM-12 0.1 g of catalyst, 0.67 h1, 350  C 60, stable for 4 h 50 Lee et al., 2010
H-SSZ-13 0.1 g of catalyst, 0.67 h1, 350  C 100, stable for 8 h 40 Lee et al., 2010
Conventional 0.05 g of catalyst, 0.8 h1, 350  C, 10 h 100, stable for 3 h 2 Wu et al., 2013
H-SSZ-13
Hierarchical 0.05 g of catalyst, 0.8 h1, 350  C, 10 h 100, stable for 8 h 3 Wu et al., 2013
H-SSZ-13
A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316 311

or the methoxy species with hydrocarbon species in the reaction nearly complete but different stability features were observed. H-
feed, with consequent production of primary olefins (ethene to MFI exhibited the highest stability properties followed by H-BEA,
butene) and later the desired hydrocarbons. This catalytic cycle also and H-MOR formed the least stable catalyst under comparable re-
generates the original hydrocarbon specie (Haw et al., 2003). The action conditions. H-MOR was more selective to the production of
existence of carbenium ions based on methylbenzenes had been light olefinic compounds throughout the reaction period. On the
demonstrated by a number of authors (Sullivan et al., 1961; Goguen other hand, in addition to the production of C5 to C8 paraffins, the
et al., 1998; Xu et al., 1998; Haw et al., 2000). This factor further H-MFI and H-BEA catalysts were highly selective to aromatic
established the participation of the Brønsted acid sites as the key compounds.
active catalyst sites.
Recently, Lee et al. (2014) indicated the control of acid strength 3.2.3. Effect of reaction parameters
by successful incorporation of Al or Fe species as beneficial for the While the structural and acidity properties of the zeolites play
H-ZSM-5 zeolite. Catalyst with co-presence of these species pro- the vital role during the MTG reactions, reaction parameters such as
duced remarkable Brønsted acid site strengths than those con- temperature, space velocity, pressure and possibly reactor config-
taining either Al or Fe only. Its selectivity to gasoline range uration must also be appropriately selected. Under uncontrolled
hydrocarbons was much higher than could be obtained with other conditions, the reaction products may composed mainly of lighter
catalysts at 500  C, when the comparable conversion was ~100%. olefin compounds or the aromatics. Similarly, the catalyst system
The results indicated that, strong Brønsted acid sites are favorable can be prone to quick deactivation with time. According to Chang
for the MTG reaction (Forselv, 2011). Lacarriere et al. (2011) con- et al. (1984), high reaction temperature and low catalyst acidity
ducted a study with two different zeolites (i.e. H-MCM-22 and H- are good parameters for the formation of olefin compounds to very
MCM-36) at 450  C and 2 h1. The H-MCM-22 catalyst that was high selectivity. The opposite parameters are however suitable for
three times denser in the concentrations of acid sites was more aromatization reactions. Ghavipour et al. (2013) conducted a study
favorable to aromatization and hydride transfer reactions. It was at 5 h1 and variable temperatures (375e475  C), using a 0.2 M
25% selective to the formation of gasoline range hydrocarbons NaOH modified H-ZSM-5 (Si/Al ¼ 38) as catalyst. The results indi-
compared to 12% for the H-MCM-36 catalyst (see Fig. 8). However, cated 425  C as the best reaction temperature. Under this condition,
its textural features enhanced catalyst deactivation due to higher the catalyst was more selective to the hydrocarbon products with
susceptibility to the formation of coke precursors. On the other improved lifetime. Even though increasing the temperature can
hand, H-MCM-36 produced mainly aliphatic hydrocarbons to 85% promote methanol conversion and selectivity to desired hydro-
selectivity, with high degree of resistance to coking. It was estab- carbons, a negative effect on the selectivity could be seriously
lished that, the combined meso- and micro-porosity characteristics apparent beyond the optimal temperature. Jiao et al. (2014)
of this catalyst enhanced non-restricted diffusion of bulkier reac- demonstrated the effect of temperature using H-ZSM-5 catalyst
tion products, with potentials as coke precursors. Therefore, modified with Fe (i.e. 0.3 wt.% Fe/H-ZSM-5). The results indicated
appropriate balance between acidity and textural properties is that, at complete conversion when the temperature was raised
necessary to achieve desired selectivity to the targeted hydrocar- from 400 to 670  C, the production of light gaseous olefins
bon products. A similar study was reported by Park and Seo (2009), increased, reaching maximum value at 540  C or the lower tem-
who employed zeolites such as H-BEA, H-MOR and the H-MFI of peratures. The production of gasoline range compounds decreased
varying acidity properties. Reactions were performed using 0.1 g of linearly with increased in reaction temperature.
catalyst at 350  C and 0.7 h1. Their characterization data showed According to Hajimirzaee et al. (2015), at a temperature of
H-MOR and H-BEA as richer in strong and weak acid sites, 400  C, raising the reaction pressure from 1 to 10 bar produced an
respectively. On the other hand, the H-MFI zeolite was stronger in increased in the conversion of methanol from 93 to 98%. However,
acid sites than H-BEA but weaker than the H-MOR catalyst. Irre- the conversion declined slightly to 96% when the pressure was
spective of the zeolite acidity or topology, the initial conversion was further raised to 20 bar. While the selectivity to gasoline range
compounds increased with pressure, the case was different for the
light olefins. The production of propylene reduced by 50% (i.e. from
36 to 18%) whereas the production of ethene risen from 24% at low
pressures to 37% at high pressures. Butene yields similar trend with
propylene because the overall selectivity reduced from 10 to only
3%. A noticeable observation with the H-ZSM-5 catalyst employed
was the increased in the amount of deposited coke with pressure.
At 0.1 MPa, the percentage deposited coke was 3.4 wt.% but
increased to 4.5 and 6.8 wt.% at 1 and 2 MPa, respectively. The ef-
fects of feed ratio (methanol/water ratio) and space velocity were
also monitored by these authors (Hajimirzaee et al., 2015).
Reducing the methanol composition in the feed from 75 to 25%
shifted the methanol conversion from 87 to 97%. Similarly, this
enhanced the selectivity to light olefins with a linear decreased in
the concentration of gasoline range hydrocarbons. On the other
hand, increasing the space velocity from 7 to 53 h1 reduced
methanol conversion (from 99 to 86%) and ethene selectivity
(27e25%) with increased in selectivity to propylene (28e36%). The
production of gasoline range hydrocarbons also increased from 3 to
about 9%. The results were correlated to some previous studies
(Chang and Silvestri, 1977; Al-Jarallah et al., 1997).
Studies by Echevskii et al. (1988) involving H-ZSM-5 catalyst
Fig. 8. Role of zeolite acidity on its selectivity to gasoline range aromatic hydrocarbons. showed tubular and shelf reactors as better configurations for the
Reaction conditions; 450  C and 2 h1. Data source: Lacarriere et al. (2011). methanol to hydrocarbon reactions than an adiabatic reactor
312 A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316

system. While the production of liquid hydrocarbon compounds 1999). Interestingly, the selectivity was dependent on the reactor
was generally high with the former reactors, the selectivity to configuration, Fluidized-bed reactor formed the best configuration,
lighter gaseous products was more prominent with the latter with 85% selectivity. The various results presented (Table 2)
reactor. Similarly, the tubular and shelf reactors were favorable in therefore indicated that, parameters such as reaction conditions,
reducing the chance of coke formation, given rise to low deacti- catalyst type and reactor configuration must be carefully and
vation and less catalyst regeneration difficulties. Zeolite-based appropriately selected for the zeolite systems.
membrane reactors are another category of reactors with good
prospects for the MTG reactions (Tago et al., 2005; Masuda et al., 3.3. Prospect of hierarchical zeolite catalysts
2003; Coronas and Santamaria, 2004). They are usually designed
to composed of meso- and microporous zeolitic materials on sup- The hierarchical zeolites otherwise referred to as “mesoporous
port of typical macroporosity (common examples include alumina zeolites” have attracted great attention in the recent times as cat-
and carbon or the stainless steel). They are characterized by multi- alysts for catalyzing many reactions including aromatization,
functional properties. The commonest of which are as catalyst and isomerization, alkylation, cracking and dehydration (Li et al., 2008;
for selective diffusion of specific reaction products (McLeary et al., Wu et al., 2012; Al-Yassir et al., 2012). They can therefore be pro-
2006). spective materials because these reactions also occur at the inter-
The details presented in Table 2 represent an informative mediate stages of the MTG process, leading to the formation of
summary on the role of reaction conditions/parameters for some gasoline range paraffins and aromatics. These zeolites are charac-
selected zeolite catalysts during the MTG reaction. According to terized by at least one additional level of porosity besides the
Bjørgen et al. (2007), reaction temperature plays a vital role in normal microporosity of zeolite systems. They have significant
defining methanol conversion and the corresponding selectivity to mesopore volume (i.e. for the pores with diameters between 2 and
gasoline hydrocarbons. Increasing the reaction temperature from 50 nm) (Holm et al., 2011). The hierarchical zeolites have unique
310 to 390  C shifted the conversion from 5 to 91%. Therefore, the properties of allowing reactions to be catalyzed both at the meso-
higher the temperature the more favorable the methanol conver- pore surfaces and within the pores due to elimination of diffusion
sion. The hydrocarbon selectivity on the other hand increased to problems (Zhou et al., 2013). Therefore, large number of reactants
39% at 330  C and remained reasonably constant at 390  C. Thus, in can be converted into the desired products to very high selectivity.
terms of hydrocarbon yield, the 390  C remained the best working Another important feature of these catalysts is their resistance to
temperature. Studies on the effect of catalyst nature on corre- deactivation with time by carbonaceous deposits (Milina et al.,
sponding lifetime was demonstrated by Shareh et al. (2014). Under 2014).
constant reaction conditions, H-MOR and Ca/H-MOR produced Recent studies revealed the hierarchical zeolites to show good
similar conversion and selectivity. However, the Ca modified cata- activity, selectivity and stability properties during the MTG reaction
lyst was more stable as it conversion decreased by 20% compared to than their conventional counterparts. Wan et al. (2014) employed a
40% with the unmodified H-MOR zeolite, for the 10 h reaction hierarchical H-ZSM-5 catalyst of 0.18 mL/g mesopore volume and
period employed. Therefore, the modification was in this case very compared its MTG performance with that of a conventional H-ZSM-
favorable. Unfortunately, both conventional and mesoporous H- 5 catalyst with 0.03 mL/g mesopore volume. The MTG reaction was
SSZ-13 (Si/Al ¼ 50) yield very low selectivity to the desired gasoline conducted at 350  C, 1.2 h1 and 1.1 MPa for a period of 24 h. The
hydrocarbons under similar reaction conditions and comparable hierarchical catalyst produced 99.6% methanol conversion
conversion (Wu and Hensen, 2014). These catalysts are therefore compared to 89.7% for the conventional catalyst. Similarly, the
less-attractive for the MTG reaction than for example the mor- selectivity to gasoline range hydrocarbons (i.e. paraffins and aro-
denite (H-MOR) based zeolites. According results from Zhang et al. matics) was 58.9% for the hierarchical catalyst compared to 28.4%
(2011), reaction conditions (0.7 g catalyst, 370  C and 2 h1) pro- for the conventional catalyst. The larger mesopore volume of the
duced mainly olefin compounds (C2 to C4) with negligible con- hierarchical H-ZSM-5 catalyst improved the diffusion of reactants
centration of gasoline hydrocarbons for the H-ZSM-5 (Si/Al ¼ 80) and products and consequently increased the selectivity to desired
zeolite. This combination of parameters is therefore unfavorable for reaction products (Gilson and Derouane, 1984; Serrano et al., 2013).
the MTG reaction with this catalyst. However, when a similar Coke analysis showed the conventional H-ZSM-5 to contain
commercial catalyst was employed at varied conditions, complete 7.7 wt.% of carbonaceous deposits compared to 1.7 wt.% for the
conversion and excellent selectivity (>80%) were observed (Keil, hierarchical catalyst after 24 h. Therefore, the hierarchical catalyst

Table 2
Role of reaction parameters during MTG reaction with zeolite catalysts.

Catalyst Reaction conditions Methanol conversion, % Selectivity/yield Ref.


of gasoline range
hydrocarbons (Cþ 5 ), %

H-ZSM-5 0.06 g of catalyst, 310  C, 7.0 h1, 0.13 MPa, 20 min 05 23 Bjørgen et al., 2007
H-ZSM-5 0.06 g of catalyst, 330  C, 7.0 h1, 0.13 MPa, 20 min 52 39 Bjørgen et al., 2007
H-ZSM-5 0.06 g of catalyst, 390  C, 7.0 h1, 0.13 MPa, 20 min 91 38 Bjørgen et al., 2007
H-MOR 2.0 g of catalyst, 1 h1, 460  C, 50% methanol in water, 12 h 100, decayed to 60% in 10 h 20 Shareh et al., 2014
Ca/HMOR 2.0 g of catalyst, 1 h1, 460  C, 50% methanol in water, 12 h 100, decayed to 80% in 10 h 20 Shareh et al., 2014
H-SSZ-13 (Si/Al ¼ 50) 0.05 g of catalyst, 350  C, 0.8 h1, 20 h 100, decayed after 20 h 02 Wu and Hensen, 2014
Mesoporous 0.05 g of catalyst, 350  C, 0.8 h1, 20 h 100, decayed after 20 h 01 Wu and Hensen, 2014
H-SSZ-13 (Si/Al ¼ 50)
H-ZSM-5 0.7 g catalyst, 370  C, 2 h1 91 Mainly selective to Zhang et al., 2011
(Si/Al ¼ 80) C2 to C4 olefins, no
higher hydrocarbons
detected
H-ZSM-5, commercial Methanol/water (83:17), 415  C, 2.0 h1 100 85 Keil, 1999
H-ZSM-5, commercial Methanol/water (83:17), 415  C, 2.0 h1 100 88 Keil, 1999
A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316 313

was more resistant to deactivation by coking. Fathi et al. (2014) production of aromatic coke precursors with potential for catalyst
prepared a hierarchical H-ZSM-5 catalyst by the treatment of a deactivation. So far, high reaction temperatures and/low methanol/
conventional H-ZSM-5 catalyst with 0.1 M NaOH at 75  C for 3 h. water ratios are associated with high methanol conversion but very
Both the conventional and hierarchical H-ZSM-5 catalysts were low selectivity to gasoline hydrocarbons. On the other hand,
evaluated for MTG reaction at 400  C and 28 h1. They possessed moderate pressures and/or space velocities are very favorable fac-
0.126 and 0.101 mL/g of mesopore volumes for the hierarchical and tors for enhanced selectivity with most zeolites. Tubular, shelf and
conventional catalysts, respectively. Although both catalysts pro- membrane reactors have good prospects to replace the fixed-bed
duced similar methanol conversion of 95% in the first 1 h of the reactors due to enhanced selectivity to liquid hydrocarbons and
reaction, the stability was catalyst dependent. The conversion the reduced chance of coke formation.
dropped to 50% within 10 h for the conventional catalyst compared Hierarchical zeolites have demonstrated good activity and sta-
to 70% for the hierarchical catalyst. Therefore, the latter was more bility properties as catalysts for the MTG process. Therefore, further
stable with time. Similarly, the hierarchical catalyst produced 42% studies with the view of developing strategies for improving the
selectivity to gasoline range hydrocarbons compared to 21% for the selectivity to gasoline range hydrocarbons are essential. Similarly,
conventional catalyst. details on their mechanism of action could be very important for
The work of Ni et al. (2011) employed a hierarchical Zn/HZSM-5 designing industrial scale catalysts.
catalyst developed by the treatment of a conventional Zn/HZSM-5
catalyst with 0.3 M NaOH at 80  C for 2 h. The hierarchical cata- Acknowledgments
lyst have a mesopore volume of 0.17 mL/g compared to 0.04 mL/g
for the conventional catalyst. Both catalysts were evaluated for The authors would like to acknowledge the funding provided by
MTG reaction at 437  C and 3.2 h1. Although both catalysts pro- King Abdulaziz City for Science and Technology (KACST) through
duced an initial methanol conversion of 100%, the stability with the Science & Technology Unit in Center of Research Excellence in
time was catalyst dependent. The conversion for convention Zn/H- Nanotechnology at King Fahd University of Petroleum & Minerals
ZSM-5 catalyst dropped to 55%v after 10 h, whereas that of the (KFUPM) for supporting this work through project No. 13-
hierarchical Zn/H-ZSM-5 catalyst was completely stable at 100% for NAN1702-04 as part of the National Science, Technology and
30 h. After a period of 10 h, the selectivity to gasoline range hy- Innovation Plan.
drocarbons was 54% for the hierarchical catalyst compared to 25%
for the conventional catalyst. Therefore, the hierarchical catalyst
References
was more active, selective and stable. According to García-Martínez
and Li (2015), the enhanced mesoporosity in hierarchical zeolites Achenbach, E., Riensche, E., 1994. Methane/steam reforming kinetics for solid oxide
promote catalyst lifetime (i.e. stability) by hindering the chance of fuel cells. J. Power Sources 52, 283e288.
coke formation. It similarly ensures the non-restricted diffusion of Agarwal, A.K., Gupta, T., Bothra, P., Shukla, P.C., February 2015. Emission profiling of
diesel and gasoline cars at a city traffic junction. Particuology 18, 186e193.
Cþ5 hydrocarbons within the pores and therefore improve their Aghaei, E., Haghighi, M., 2015. Effect of crystallization time on properties and cat-
selectivity. On the other hand, the pores of conventional zeolites are alytic performance of nanostructured SAPO-34 molecular sieve synthesized at
easily blocked by coke species. This lower their selectivity to Cþ 5
high temperatures for conversion of methanol to light olefins. Powder Technol.
269, 358e370.
hydrocarbons and trigger rapid catalyst deactivation. Aghamohammadi, S., Haghighi, M., 2015. Dual-template synthesis of nano-
structured CoAPSO-34 used in methanol to olefins: effect of template combi-
4. Conclusions nations on catalytic performance and coke formation. Chem. Eng. J. 264,
359e375.
Al-Jarallah, A.M., El-Nafaty, U.A., Abdillahi, M.M., 1997. Effects of metal impregna-
There are indications that shale gas have good potential to tion on the activity, selectivity and deactivation of a high silica MFI zeolite when
compete with natural gas or biomass, as raw material for syngas converting methanol to light alkenes. Appl. Catal. A Gen. 154, 117e127.
Al-Yassir, N., Akhtar, M.N., Ogunronbi, K., Al-Khattaf, S., 2012. Synthesis of stable H-
production. However, cost-effective and efficient shale gas purifi-
galloaluminosilicate MFI with hierarchical pore architecture by surfactant-
cation processes must be developed, considering its chemical mediated base hydrolysis, and their application in propane aromatization.
compositions in comparison to other feedstocks. The reforming (i.e. J. Mol. Catal. A Chem. 360, 1e15.
Albarazi, A., Ga lvez, M.E., Da Costa, P., 2014. Synthesis strategies of ceria-zirconia
dry and steam) technologies are considered to be very effective
doped Ni/SBA-15 catalysts for methane dry reforming. Catal. Commun. 59,
syngas production processes, with supported and/or promoted Ni- 108e112.
based systems as the main catalysts. However, the recent literature An, X., Li, J., Zuo, Y., Zhang, Q., Wang, D., Wang, J., 2007. A Cu/Zn/Al/Zr fibrous
indicated that, effort to reduce the reaction temperature required to catalyst that is an improved CO2 hydrogenation to methanol catalyst. Catal. Lett.
118, 264e269.
achieve optimal syngas yield would be very beneficial. A combi- Anderson, M.W., Klinowski, J., 1990. Solid-state NMR studies of the shape-selective
nation of good catalyst design and appropriate selection of reaction catalytic conversion of methanol into gasoline on zeolite ZSM-5. J. Am. Chem.
parameters is necessary to ensure longer catalyst lifetime and Soc. 112, 10e16.
Angeli, S.D., Pilitsis, F.G., Lemonidou, A.A., 2014. Methane steam reforming at low
prevent deactivation by carbonaceous depositions. temperature: effect of light alkanes' presence on coke formation. Catal. Today
Catalysts based on noble metals are very active for methanol 242, 119e128.
production from syngas. However, sintering, poisoning and asso- Arab, S., Commenge, J.-M., Portha, J.-F., Falk, L., 2014. Methanol synthesis from CO2
and H2 in multi-tubular fixed-bed reactor and multi-tubular reactor filled with
ciated costs accounted for a shift to other catalysts based on Zn and monoliths. Chem. Eng. Res. Des. 92, 2598e2608.
Cu. Further studies are necessary to fully explore the most appro- Baba, T., Sakai, J., Watanabe, H., Ono, Y., 1982. The conversion of methanol into
priate supports, reaction conditions and catalyst modifications that hydrocarbons over Dodecatungstophosphoric acid. Bull. Chem. Soc. Jpn. 55,
2555e2559.
could give the best activity, stability and selectivity properties. Baba, T., Sakai, J., Ono, Y., 1982. The conversion of methanol into hydrocarbons over
Zeolite catalysts are the main systems employed for the MTG metal salts of Heteropolyacids. Bull. Chem. Soc. Jpn. 55, 2657e2658.
reaction due to the possibility of structural and acidity control. Benito, P.L., Gayubo, A.G., Aguayo, A.T., Olazar, M., Bilbao, J., 1996. Effect of Si/Al ratio
and of acidity of H-ZSM5 zeolites on the primary products of methanol to
Attaining optimal methanol conversion is easy with most zeolite
gasoline conversion. J. Chem. Technol. Biotechnol. 66, 183e191.
catalysts but achieving high yield of gasoline range paraffins is Bjørgen, M., Olsbye, U., Kolboe, S., 2003. Coke precursor formation and zeolite
difficult. Therefore, careful selection of zeolite structures with deactivation: mechanistic insights from hexamethylbenzene conversion.
sufficient Brønsted acidity that can preferentially and selectively J. Catal. 215, 30e44.
Bjørgen, M., Svelle, S., Joensen, F., Nerlov, J., Kolboe, S., Bonino, F., Palumbo, L.,
produce both n- and i-gasoline range paraffins to very high selec- Bordiga, S., Olsbye, U., 2007. Conversion of methanol to hydrocarbons over
tivity is necessary. The catalysts must also be resistance to zeolite H-ZSM-5: on the origin of the olefinic species. J. Catal. 249, 195e207.
314 A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316

Bjørgen, M., Joensen, F., Spangsberg Holm, M., Olsbye, U., Lillerud, K.-P., Svelle, S., Appl. Catal. A Gen. 338, 100e113.
2008. Methanol to gasoline over zeolite H-ZSM-5: Improved catalyst perfor- Di, Z., Yang, C., Jiao, X., Li, J., Wu, J., Zhang, D., 2013. A ZSM-5/MCM-48 based catalyst
mance by treatment with NaOH. Appl. Catal. A Gen. 345, 43e50. for methanol to gasoline conversion. Fuel 104, 878e881.
Bjørgen, M., Joensen, F., Lillerud, K.-P., Olsbye, U., Svelle, S., 2009. The mechanisms Li, L., Mao, D., Yu, J., Guo, X., 2015. Highly selective hydrogenation of CO2 to
of ethene and propene formation from methanol over high silica H-ZSM-5 and methanol over CuOeZnOeZrO2 catalysts prepared by a surfactant-assisted co-
H-beta. Catal. Today 142, 90e97. precipitation method. J. Power Sources 279, 394e404.
Bjørgen, M., Akyalcin, S., Olsbye, U., Benard, S., Kolboe, S., Svelle, S., 2010. Methanol Fathi, S., Sohrabi, M., Falamaki, C., 2014. Improvement of HZSM-5 performance by
to hydrocarbons over large cavity zeolites: toward a unified description of alkaline treatments: comparative catalytic study in the MTG reactions. Fuel 116,
catalyst deactivation and the reaction mechanism. J. Catal. 275, 170e180. 529e537.
Bowker, K.A., 2007. Barnett shale gas production, Fort Worth Basin: issues and Forselv, S., 2011. An in-situ FTIR Study of the Methanol to Hydrocarbons Reaction
discussion. AAPG Bull. 91, 523e533. over Zeolite Catalysts.
Bridgwater, A., 1995. The technical and economic feasibility of biomass gasification Fujitani, T., Saito, M., Kanai, Y., Watanabe, T., Nakamura, J., Uchijima, T., 1995.
for power generation. Fuel 74, 631e653. Development of an active Ga2O3 supported palladium catalyst for the synthesis
Bruijnincx, P.C., Weckhuysen, B.M., 2013. Shale gas revolution: an opportunity for of methanol from carbon dioxide and hydrogen. Appl. Catal. A Gen. 125,
the production of biobased chemicals? Angew. Chem. Int. Ed. 52, 11980e11987. L199eL202.
Cai, W., De La Piscina, P.R., Toyir, J., Homs, N., 2015. CO2 hydrogenation to methanol Furimsky, E., 2000. Catalytic hydrodeoxygenation. Appl. Catal. A Gen. 199, 147e190.
over CuZnGa catalysts prepared using microwave-assisted methods. Catal. Gaboriaud, F., Vantelon, J.-P., Guelzim, A., Julien, L., 1991. Gas evolution during
Today 242, 193e199. isothermal pyrolysis of Timahdit oil shale. J. Anal. Appl. Pyrolysis 21, 119e131.
Campbell, C.J., Laherre re, J.H., 1998. The end of cheap oil. Sci. Am. 278, 60e65. Galadima, A., Wells, R.P.K., Anderson, J.A., 2012. n-Alkane hydroconversion over
Chang, C.D., Silvestri, A.J., 1977. The conversion of methanol and other O-com- carbided molybdena supported on sulfated zirconia. Appl. Petrochem. Res. 1,
pounds to hydrocarbons over zeolite catalysts. J. Catal. 47, 249e259. 35e43.
Chang, M.-F., Evans, L., Herman, R., Wasielewski, P., 1976. Gasoline Consumption in Gao, P., Li, F., Zhan, H., Zhao, N., Xiao, F., Wei, W., Zhong, L., Wang, H., Sun, Y., 2013.
Urban Traffic. Transportation Research Record. Influence of Zr on the performance of Cu/Zn/Al/Zr catalysts via hydrotalcite-like
Chang, C.D., Chu, C.T.W., Socha, R.F., 1984. Methanol conversion to olefins over ZSM- precursors for CO2 hydrogenation to methanol. J. Catal. 298, 51e60.
5: I. Effect of temperature and zeolite SiO2Al2O3. J. Catal. 86, 289e296. Gao, P., Li, F., Zhang, L., Zhao, N., Xiao, F., Wei, W., Zhong, L., Sun, Y., 2013. Influence
Charusiri, W., Yongchareon, W., Vitidsant, T., 2006. Conversion of used vegetable of fluorine on the performance of fluorine-modified Cu/Zn/Al catalysts for CO2
oils to liquid fuels and chemicals over HZSM-5, sulfated zirconia and hybrid hydrogenation to methanol. J. CO2 Util. 2, 16e23.
catalysts. Korean J. Chem. Eng. 23, 349e355. Gao, P., Li, F., Zhan, H., Zhao, N., Xiao, F., Wei, W., Zhong, L., Sun, Y., 2014. Fluorine-
Cheng, W.-H., 1994. Methanol Production and Use. CRC Press. modified Cu/Zn/Al/Zr catalysts via hydrotalcite-like precursors for CO2 hydro-
Choudhary, V.R., Mondal, K.C., Mulla, S.A., 2005. Simultaneous conversion of genation to methanol. Catal. Commun. 50, 78e82.
methane and methanol into gasoline over bifunctional Ga-, Zn-, In-, and/or Mo- García-Da vila, J., Ocaranza-Sa nchez, E., Rojas-Lo  pez, M., Mun
~ oz-Arroyo, J.A.,
modified ZSM-5 zeolites. Angew. Chem. 117, 4455e4459. Ramírez, J., Martínez-Ayala, A.L., 2014. Jatropha curcas L. oil hydroconversion
Chu, Z., Chen, H., Yu, Y., Wang, Q., Fang, D., 2013. Surfactant-assisted preparation of over hydrodesulfurization catalysts for biofuel production. Fuel 135, 380e386.
Cu/ZnO/Al2O3 catalyst for methanol synthesis from syngas. J. Mol. Catal. A García-Martínez, J., Li, K., 2015. Mesoporous Zeolites: Preparation, Characterization
Chem. 366, 48e53. and Applications. John Wiley & Sons.
Clarke, S.H., Dicks, A.L., Pointon, K., Smith, T.A., Swann, A., 1997. Catalytic aspects of Ghavipour, M., Behbahani, R.M., Moradi, G.R., Soleimanimehr, A., 2013. Methanol
the steam reforming of hydrocarbons in internal reforming fuel cells. Catal. dehydration over alkali-modified H-ZSM-5; effect of temperature and water
Today 38, 411e423. dilution on products distribution. Fuel 113, 310e317.
Collins, S.E., Chiavassa, D.L., Bonivardi, A.L., Baltana s, M.A., 2005. Hydrogen spillover Gilson, J.P., Derouane, E.G., 1984. On the external and intracrystalline surface cat-
in Ga2O3-Pd/SiO2 catalysts for methanol synthesis from CO2/H2. Catal. Lett. 103, alytic activity of pentasil zeolites. J. Catal. 88, 538e541.
83e88. Goehna, H., Koenig, P., 1994. Producing methanol from CO2. Chemtech 24, 36e39.
Coronas, J., Santamaria, J., 2004. State-of-the-art in zeolite membrane reactors. Top. Goguen, P.W., Xu, T., Barich, D.H., Skloss, T.W., Song, W., Wang, Z., Nicholas, J.B.,
Catal. 29, 29e44. Haw, J.F., 1998. Pulse-quench catalytic reactor studies reveal a carbon-pool
Courson, C., Makaga, E., Petit, C., Kiennemann, A., 2000. Development of Ni catalysts mechanism in methanol-to-gasoline chemistry on zeolite HZSM-5. J. Am.
for gas production from biomass gasification. Reactivity in steam-and dry- Chem. Soc. 120, 2650e2651.
reforming. Catal. Today 63, 427e437. Gubareni, I.V., Kurilets, Y.P., Soloviev, S.O., November 2014. Effect of additives La2O3
Crabtree, G.W., Dresselhaus, M.S., Buchanan, M.V., 2004. The hydrogen economy. and CeO2 on the activity and selectivity of Ni-Al2O3/cordierite catalysts in steam
Phys. Today 57, 39e44. reforming of methane. Theor. Exp. Chem. 50 (5), 311e317.
Dagle, R.A., Lizarazo-Adarme, J.A., Lebarbier Dagle, V., Gray, M.J., White, J.F., Guo, J., Lou, H., Zhao, H., Chai, D., Zheng, X., 2004. Dry reforming of methane over
King, D.L., Palo, D.R., 2014. Syngas conversion to gasoline-range hydrocarbons nickel catalysts supported on magnesium aluminate spinels. Appl. Catal. A Gen.
over Pd/ZnO/Al2O3 and ZSM-5 composite catalyst system. Fuel Process. Technol. 273, 75e82.
123, 65e74. Guo, X., Mao, D., Lu, G., Wang, S., Wu, G., 2011. The influence of la doping on the
Dahl, I.M., Kolboe, S., 1996. On the reaction mechanism for hydrocarbon formation catalytic behavior of Cu/ZrO2 for methanol synthesis from CO2 hydrogenation.
from methanol over SAPO-34: 2. Isotopic labeling studies of the Co-reaction of J. Mol. Catal. A Chem. 345, 60e68.
propene and methanol. J. Catal. 161, 304e309. Hajimirzaee, S., Ainte, M., Soltani, B., Behbahani, R.M., Leeke, G.A., Wood, J., 2015.
Dejaifve, P., Auroux, A., Gravelle, P.C., Ve drine, J.C., Gabelica, Z., Derouane, E.G., 1981. Dehydration of methanol to light olefins upon zeolite/alumina catalysts: effect
Methanol conversion on acidic ZSM-5, offretite, and mordenite zeolites: a of reaction conditions, catalyst support and zeolite modification. Chem. Eng.
comparative study of the formation and stability of coke deposits. J. Catal. 70, Res. Des. 93, 541e553.
123e136. Hamnett, A., 1997. Mechanism and electrocatalysis in the direct methanol fuel cell.
Dessau, R.M., 1986. On the H-ZSM-5 catalyzed formation of ethylene from methanol Catal. Today 38, 445e457.
or higher olefins. J. Catal. 99, 111e116. Haw, J.F., Nicholas, J.B., Song, W., Deng, F., Wang, Z., Xu, T., Heneghan, C.S., 2000.
Dessau, R.M., LaPierre, R.B., 1982. On the mechanism of methanol conversion to Roles for cyclopentenyl cations in the synthesis of hydrocarbons from methanol
hydrocarbons over HZSM-5. J. Catal. 78, 136e141. on zeolite catalyst HZSM-5. J. Am. Chem. Soc. 122, 4763e4775.
Djaidja, A., Libs, S., Kiennemann, A., Barama, A., 2006. Characterization and activity Haw, J.F., Song, W., Marcus, D.M., Nicholas, J.B., 2003. The mechanism of methanol
in dry reforming of methane on NiMg/Al and Ni/MgO catalysts. Catal. Today 113, to hydrocarbon catalysis. Acc. Chem. Res. 36, 317e326.
194e200. Hayashi, H., Moffat, J., 1983. Conversion of methanol into hydrocarbons over
Dry, M.E., 2002. The FischereTropsch process: 1950e2000. Catal. Today 71, ammonium 12-tungstophosphate. J. Catal. 83, 192e204.
227e241. He, Y., Liu, M., Dai, C., Xu, S., Wei, Y., Liu, Z., Guo, X., 2013. Modification of nano-
Echevskii, G.V., Ione, K.G., Nosyreva, G.N., Litvak, G.S., 1988. Effect of the tempera- crystalline HZSM-5 zeolite with tetrapropylammonium hydroxide and its cat-
ture regime of methanol conversion to hydrocarbons on coking of zeolite cat- alytic performance in methanol to gasoline reaction. Chin. J. Catal. 34,
alysts and their regeneration. Appl. Catal. 43, 85e89. 1148e1158.
Edwards, J.D., 1997. Crude oil and alternate energy production forecasts for the Herder, P.M., Stikkelman, R.M., 2004. Methanol-based industrial cluster design: a
twenty-first century: the end of the hydrocarbon era. AAPG Bull. 81, study of design options and the design process. Indust. Eng. Chem. Res. 43,
1292e1305. 3879e3885.
Edwards, J.H., Maitra, A.M., 1995. The chemistry of methane reforming with carbon Hoek, A., Lednor, P.W., Minderhoud, J.K., Post, M.F., 1985. Process for the preparation
dioxide and its current and potential applications. Fuel Process. Technol. 42, of a Fischer-Tropsch catalyst and preparation of hydrocarbons from syngas in,
269e289. Google Patents.
Ehlinger, V.M., Gabriel, K.J., Noureldin, M.M.B., El-Halwagi, M.M., 2013. Process Hoffmann, P., 2012. Tomorrow's Energy: Hydrogen, Fuel Cells, and the Prospects for
design and integration of shale gas to methanol. ACS Sustain. Chem. Eng. 2, a Cleaner Planet. MIT Press.
30e37. Holm, M.S., Taarning, E., Egeblad, K., Christensen, C.H., 2011. Catalysis with hierar-
Erofeev, V.I., Khomyakov, I.S., Egorova, L.A., 2014. Production of high-octane gaso- chical zeolites. Catal. Today 168, 3e16.
line from straight-run gasoline on ZSM-5 modified zeolites. Theor. Found. Hu, Z., Chen, S., Peng, S., 1996. Preparation of nanosize Cu-ZnO/Al2O3 catalyst for
Chem. Eng. 48, 71e76. methanol synthesis by phase transfer with metal surfactant. 2. Effect of addi-
Li, Y., Liu, S., Zhang, Z., Xie, S., Zhu, X., Xu, L., 2008. Aromatization and isomerization tives and chain length of surfactant. J. Colloid Interface Sci. 182, 461e464.
of 1-hexene over alkali-treated HZSM-5 zeolites: improved reaction stability. Hu, Z., Chen, S., Peng, S., 1996. Preparation of nanosize Cu-ZnO/Al2O3 catalyst for
A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316 315

methanol synthesis by phase transfer with metal surfactant. 1. A study of synthesis from syngas using factorial experimental design. Fuel 89, 170e175.
preparation conditions. J. Colloid Interface Sci. 182, 457e460. Milina, M., Mitchell, S., Crivelli, P., Cooke, D., Pe rez-Ramírez, J., 2014. Mesopore
Huber, G.W., O'Connor, P., Corma, A., 2007. Processing biomass in conventional oil quality determines the lifetime of hierarchically structured zeolite catalysts.
refineries: production of high quality diesel by hydrotreating vegetable oils in Nat. Commun. http://dx.doi.org/10.1038/ncomms4922.
heavy vacuum oil mixtures. Appl. Catal. A Gen. 329, 120e129. Milne, T.A., Evans, R.J., Nagle, N., 1990. Catalytic conversion of microalgae and
Jabbour, K., El Hassan, N., Davidson, A., Massiani, P., Casale, S., 2015. Characteriza- vegetable oils to premium gasoline, with shape-selective zeolites. Biomass 21,
tions and performances of Ni/diatomite catalysts for dry reforming of methane. 219e232.
Chem. Eng. J. 264, 351e358. Misono, M., 1987. Heterogeneous catalysis by heteropoly compounds of molybde-
Jiao, M., Fan, S., Zhang, J., Su, X., Zhao, T.-S., 2014. Methanol-to-olefins over FeHZSM- num and tungsten. Catal. Rev. Sci. Eng. 29, 269e321.
5: further transformation of products. Catal. Commun. 56, 153e156. Mole, T., Whiteside, J.A., Seddon, D., 1983. Aromatic co-catalysis of methanol con-
Johnsen, K., Ryu, H.J., Grace, J.R., Lim, C.J., 2006. Sorption-enhanced steam reforming version over zeolite catalysts. J. Catal. 82, 261e266.
of methane in a fluidized bed reactor with dolomite as -acceptor. Chem. Eng. NETL, 2014. U.S. Department of Energy; Liquid Fuels, Conversion of Methanol to
Sci. 61, 1195e1202. Gasoline. National Energy Technology Laboratory, United State of America.
Joo, O.-S., Jung, K.-D., Moon, I., Rozovskii, A.Y., Lin, G.I., Han, S.-H., Uhm, S.-J., 1999. Ni, Y., Sun, A., Wu, X., Hai, G., Hu, J., Li, T., Li, G., 2011. Preparation of hierarchical
Carbon dioxide hydrogenation to form methanol via a reverse-water-gas-shift mesoporous Zn/HZSM-5 catalyst and its application in MTG reaction. J. Nat. Gas
reaction (the CAMERE process). Indust. Eng. Chem. Res. 38, 1808e1812. Chem. 20, 237e242.
Kaeding, W.W., Butter, S.A., 1980. Production of chemicals from methanol: I. Low Niiyama, H., Saito, Y., Yoshida, S., Echigoya, E., 1982. Physical and acid-catalytic
molecular weight olefins. J. Catal. 61, 155e164. properties of 12-molybdophosphates. Nippon. Kagaku Kaishi 1982, 569e573.
Keil, F.J., 1999. Methanol-to-hydrocarbons: process technology. Microporous Mes- Olah, G.A., 2005. Beyond oil and gas: the methanol economy. Angew. Chem. Int. Ed.
oporous Mater. 29, 49e66. 44, 2636e2639.
Kim, L., Wald, M.M., Brandenberger, S.G., 1978. One-step catalytic synthesis of 2,2,3- Olah, G.A., Goeppert, A., Prakash, G.K.S., 2008. Chemical recycling of carbon dioxide
trimethylbutane from methanol. J. Org. Chem. 43, 3432e3433. to methanol and dimethyl ether: from greenhouse gas to renewable, environ-
Kim, S.S., Ly, H.V., Kim, J., Lee, E.Y., Woo, H.C., 2015. Pyrolysis of microalgae residual mentally carbon neutral fuels and synthetic hydrocarbons. J. Org. Chem. 74,
biomass derived from Dunaliella tertiolecta after lipid extraction and carbo- 487e498.
hydrate saccharification. Chem. Eng. J. 263, 194e199. Olsbye, U., Svelle, S., Bjørgen, M., Beato, P., Janssens, T.V.W., Joensen, F., Bordiga, S.,
King, G.E., 2010. Thirty Years of Gas Shale Fracturing: what Have We Learned?. Lillerud, K.P., 2012. Conversion of methanol to hydrocarbons: how zeolite cavity
Paper presented at the SPE Annual Technical Conference and Exhibition. and pore size controls product selectivity. Angew. Chem. Int. Ed. 51, 5810e5831.
Klier, K., 1982. Methanol synthesis. Adv. Catal. 31, 243e313. Ou, J., Guo, H., Zheng, J., Cheung, K., Louie, P.K.K., Ling, Z., Wang, D., 2015. Con-
Kubi  a
cka, D., Sim  
cek, P., Zilkova, N., 2009. Transformation of vegetable oils into centrations and sources of non-methane hydrocarbons (NMHCs) from 2005 to
hydrocarbons over mesoporous-alumina-supported CoMo catalysts. Top. Catal. 2013 in Hong Kong: a multi-year real-time data analysis. Atmos. Environ. 103,
52, 161e168. 196e206.

Lacarriere, A., Luck, F., Swierczy n ski, D., Fajula, F., Hulea, V., 2011. Methanol to hy- Owen, N.A., Inderwildi, O.R., King, D.A., 2010. The status of conventional world oil
drocarbons over zeolites with MWW topology: effect of zeolite texture and reservesdhype or cause for concern? Energy Policy 38, 4743e4749.
acidity. Appl. Catal. A Gen. 402, 208e217. Oyola-Rivera, O., Baltan as, M.A., Cardona-Martínez, N., 2015. CO2 hydrogenation to
Ladera, R., Pe rez-Alonso, F.J., Gonz alez-Carballo, J.M., Ojeda, M., Rojas, S., methanol and dimethyl ether by Pd-Pd2Ga catalysts supported over Ga2O3
Fierro, J.L.G., 2013. Catalytic valorization of CO2 via methanol synthesis with Ga- polymorphs. J. CO2 Util. 9, 8e15.
promoted CueZnOeZrO2 catalysts. Appl. Catal. B Environ. 142e143, 241e248. Pakhare, D., Spivey, J., 2014. A review of dry (CO2) reforming of methane over noble
Laosiripojana, N., Assabumrungrat, S., 2005. Catalytic dry reforming of methane metal catalysts. Chem. Soc. Rev. 43, 7813e7837.
over high surface area ceria. Appl. Catal. B Environ. 60, 107e116. Palma, V., Miccio, M., Ricca, A., Meloni, E., Ciambelli, P., 2014. Monolithic catalysts
Lee, J.H., Park, M.B., Lee, J.K., Min, H.-K., Song, M.K., Hong, S.B., 2010. Synthesis and for methane steam reforming intensification: experimental and numerical in-
characterization of ERI-type UZM-12 zeolites and their methanol-to-olefin vestigations. Fuel 138, 80e90.
performance. J. Am. Chem. Soc. 132, 12971e12982. Park, J.W., Seo, G., 2009. IR study on methanol-to-olefin reaction over zeolites with
Lee, K.-Y., Lee, S.-W., Ihm, S.-K., 2014. Acid strength control in MFI zeolite for the different pore structures and acidities. Appl. Catal. A Gen. 356, 180e188.
methanol-to-hydrocarbons (MTH) reaction. Indust. Eng. Chem. Res. 53, Pretzer, W.R., Kobylinski, T.P., 1980. Methanol carbonylation as an alternate route to
10072e10079. chemicals. Ann. N. Y. Acad. Sci. 333, 58e66.
Liu, X.-M., Lu, G.Q., Yan, Z.-F., Beltramini, J., 2003. Recent advances in catalysts for Ravaghi-Ardebili, Z., Manenti, F., Corbetta, M., Pirola, C., Ranzi, E., 2015. Biomass
methanol synthesis via hydrogenation of CO and CO2. Indust. Eng. Chem. Res. gasification using low-temperature solar-driven steam supply. Renew. Energy
42, 6518e6530. 74, 671e680.
Liu, Y., Sotelo-Boya s, R., Murata, K., Minowa, T., Sakanishi, K., 2011. Hydrotreatment Raveendran, K., Ganesh, A., Khilar, K.C., 1995. Influence of mineral matter on
of vegetable oils to produce bio-hydrogenated diesel and liquefied petroleum biomass pyrolysis characteristics. Fuel 74, 1812e1822.
gas fuel over catalysts containing sulfided Ni-Mo and solid acids. Energy Fuels Raveendran, K., Ganesh, A., Khilar, K.C., 1996. Pyrolysis characteristics of biomass
25, 4675e4685. and biomass components. Fuel 75, 987e998.
Ma, Y., Ge, Q., Li, W., Xu, H., 2008. Study on the sulfur tolerance of catalysts for Rearch, G.V. Grand View Research, Shale Gas Market Trends, Analysis and Segment
syngas to methanol. Catal. Commun. 10, 6e10. Forecasts to 2020, Published: March 2014 ISBN Code: 978-1-68038-025-5,
Ma, Y., Wang, X., You, X., Liu, J., Tian, J., Xu, X., Peng, H., Liu, W., Li, C., Zhou, W., Grand View Research, Market Research and Consulting, San Francisco, United
Yuan, P., Chen, X., 2014. Nickel-supported on La2Sn2O7 and La2Zr2O7 pyro- States.
chlores for methane steam reforming: insight into the difference between tin Rostrup-Nielsen, J.R., 2000. New aspects of syngas production and use. Catal. Today
and zirconium in the b site of the compound. ChemCatChem 6, 3366e3376. 63, 159e164.
Maher, K.D., Bressler, D.C., 2007. Pyrolysis of triglyceride materials for the produc- Rownaghi, A.A., Rezaei, F., Hedlund, J., 2011. Yield of gasoline-range hydrocarbons as
tion of renewable fuels and chemicals. Bioresour. Technol. 98, 2351e2368. a function of uniform ZSM-5 crystal size. Catal. Commun. 14, 37e41.
Malleswara Rao, T.V., Dupain, X., Makkee, M., 2012. Fluid catalytic cracking: pro- Saito, M., 1998. R&D activities in Japan on methanol synthesis from CO2 and H2.
cessing opportunities for FischereTropsch waxes and vegetable oils to produce Catal. Surv. Asia 2, 175e184.
transportation fuels and light olefins. Microporous Mesoporous Mater. 164, Saito, M., Murata, K., 2004. Development of high performance Cu/ZnO-based cat-
148e163. alysts for methanol synthesis and the water-gas shift reaction. Catal. Surv. Asia
Marcus, D.M., McLachlan, K.A., Wildman, M.A., Ehresmann, J.O., Kletnieks, P.W., 8, 285e294.
Haw, J.F., 2006. Experimental evidence from H/D exchange studies for the Saito, M., Fujitani, T., Takeuchi, M., Watanabe, T., 1996. Development of copper/zinc
failure of direct C-C coupling mechanisms in the methanol-to-olefin process oxide-based multicomponent catalysts for methanol synthesis from carbon
catalyzed by HSAPO-34. Angew. Chem. Int. Ed. 45, 3133e3136. dioxide and hydrogen. Appl. Catal. A Gen. 138, 311e318.
Masuda, T., Asanuma, T., Shouji, M., Mukai, S.R., Kawase, M., Hashimoto, K., 2003. Samei, E., Taghizadeh, M., Bahmani, M., 2012. Enhancement of stability and activity
Methanol to olefins using ZSM-5 zeolite catalyst membrane reactor. Chem. Eng. of Cu/ZnO/Al2O3 catalysts by colloidal silica and metal oxides additives for
Sci. 58, 649e656. methanol synthesis from a CO2-rich feed. Fuel Process. Technol. 96, 128e133.
Matsumura, Y., Shen, W.J., Ichihashi, Y., Okumura, M., 2001. Low-temperature Sassi, A., Wildman, M.A., Ahn, H.J., Prasad, P., Nicholas, J.B., Haw, J.F., 2002. Meth-
methanol synthesis catalyzed over ultrafine palladium particles supported on ylbenzene chemistry on zeolite HBeta: multiple insights into methanol-to-
cerium oxide. J. Catal. 197, 267e272. olefin catalysis. J. Phys. Chem. B 106, 2294e2303.
Mauldin, C.H., 1986. Cobalt catalysts for the conversion of methanol to hydrocar- Saxena, S.K., Viswanadham, N., 2014. Selective production of green gasoline by
bons and for Fischer-Tropsch synthesis in, Google Patents. catalytic conversion of Jatropha oil. Fuel Process. Technol. 119, 158e165.
McGillivray, R.G., 1976. Gasoline Use by Automobiles. Transportation Research Saxena, S.K., Viswanadham, N., Al-Muhtaseb, A.H., 2014. Enhanced production of
Record. high octane gasoline blending stock from methanol with improved catalyst life
McLeary, E.E., Jansen, J.C., Kapteijn, F., 2006. Zeolite based films, membranes and on nano-crystalline ZSM-5 catalyst. J. Indust. Eng. Chem. 20, 2876e2882.
membrane reactors: progress and prospects. Microporous Mesoporous Mater. Serrano, D.P., Escola, J.M., Pizarro, P., 2013. Synthesis strategies in the search for
90, 198e220. hierarchical zeolites. Chem. Soc. Rev. 42, 4004e4035.
McNicol, B., Rand, D., Williams, K., 1999. Direct methanoleair fuel cells for road Shareh, F.B., Kazemeini, M., Asadi, M., Fattahi, M., 2014. Metal promoted mordenite
transportation. J. Power Sources 83, 15e31. catalyst for methanol conversion into light olefins. Pet. Sci. Technol. 32,
Meshkini, F., Taghizadeh, M., Bahmani, M., 2010. Investigating the effect of metal 1349e1356.
oxide additives on the properties of Cu/ZnO/Al2O3 catalysts in methanol Shen, W.J., Ichihashi, Y., Ando, H., Matsumura, Y., Okumura, M., Haruta, M., 2001.
316 A. Galadima, O. Muraza / Journal of Natural Gas Science and Engineering 25 (2015) 303e316

Effect of reduction temperature on structural properties and CO/CO2 hydroge- ZSM-5 zeolite and its performance in catalyzing methanol to gasoline conver-
nation characteristics of a Pd-CeO2 catalyst. Appl. Catal. A Gen. 217, 231e239. sion. Indust. Eng. Chem. Res. 53, 19471e19478.
Song, H., Wang, N., Song, H.-L., Li, F., 2015. LaeNi modified S2O82/ZrO2-Al2O3 Wang, W., Buchholz, A., Seiler, M., Hunger, M., 2003. Evidence for an initiation of the
catalyst in n-pentane hydroisomerization. Catal. Commun. 59, 61e64. methanol-to-olefin process by reactive surface methoxy groups on acidic
Spencer, M.S., 1998. Role of ZnO in methanol synthesis on copper catalysts. Catal. zeolite catalysts. J. Am. Chem. Soc. 125, 15260e15267.
Lett. 50, 37e40. Wang, L., Yang, L., Zhang, Y., Ding, W., Chen, S., Fang, W., Yang, Y., 2010. Promoting
Sudhakar, C., Vannice, M.A., 1985. Methanol and methane formation over palla- effect of an aluminum emulsion on catalytic performance of Cu-based catalysts
dium/rare earth oxide catalysts. J. Catal. 95, 227e243. for methanol synthesis from syngas. Fuel Process. Technol. 91, 723e728.
Sullivan, R.F., Egan, C.J., Langlois, G.E., Sieg, R.P., 1961. A new reaction that occurs in Wang, C.-M., Wang, Y.-D., Xie, Z.-K., 2013. Insights into the reaction mechanism of
the hydrocracking of certain aromatic hydrocarbons. J. Am. Chem. Soc. 83, methanol-to-olefins conversion in HSAPO-34 from first principles: are olefins
1156e1160. themselves the dominating hydrocarbon pool species? J. Catal. 301, 8e19.
Surisetty, V.R., Dalai, A.K., Kozinski, J., 2011. Alcohols as alternative fuels: an over- Wang, C., Sun, N., Zhao, N., Wei, W., Sun, Y., Sun, C., Liu, H., Snape, C.E., 2015. Coking
view. Appl. Catal. A Gen. 404, 1e11. and deactivation of a mesoporous Ni-CaO-ZrO2 catalyst in dry reforming of
Surisetty, V.R., Epelde, E., Tre panier, M., Kozinski, J., Dalai, A.K., 2012. The role of methane: a study under different feeding compositions. Fuel 143, 527e535.
catalytic site deposition on cobalt catalysts supported on carbon nanotubes for Wender, I., 1996. Reactions of synthesis gas. Fuel Process. Technol. 48, 189e297.
Fisher-Trospch synthesis. Int. J. Chem. React. Eng. 10. Wu, L., Hensen, E.J.M., 2014. Comparison of mesoporous SSZ-13 and SAPO-34
Sutton, D., Kelleher, B., Ross, J.R., 2001. Review of literature on catalysts for biomass zeolite catalysts for the methanol-to-olefins reaction. Catal. Today 235,
gasification. Fuel Process. Technol. 73, 155e173. 160e168.
Svelle, S., Joensen, F., Nerlov, J., Olsbye, U., Lillerud, K.-P., Kolboe, S., Bjørgen, M., Wu, Y., Tian, F., Liu, J., Song, D., Jia, C., Chen, Y., 2012. Enhanced catalytic isomeri-
2006. Conversion of methanol into hydrocarbons over zeolite H-ZSM-5: ethene zation of a-pinene over mesoporous zeolite beta of low Si/Al ratio by NaOH
formation is mechanistically separated from the formation of higher alkenes. treatment. Microporous Mesoporous Mater. 162, 168e174.
J. Am. Chem. Soc. 128, 14770e14771. Wu, L., Degirmenci, V., Magusin, P.C.M.M., Lousberg, N.J.H.G.M., Hensen, E.J.M., 2013.
Svelle, S., Olsbye, U., Joensen, F., Bjørgen, M., 2007. Conversion of methanol to al- Mesoporous SSZ-13 zeolite prepared by a dual-template method with improved
kenes over medium- and large-pore acidic zeolites: steric manipulation of the performance in the methanol-to-olefins reaction. J. Catal. 298, 27e40.
reaction intermediates governs the ethene/propene product selectivity. J. Phys. Xie, Q., Kong, S., Liu, Y., Zeng, H., 2012. Syngas production by two-stage method of
Chem. C 111, 17981e17984. biomass catalytic pyrolysis and gasification. Bioresour. Technol. 110, 603e609.
Tago, T., Iwakai, K., Morita, K., Tanaka, K., Masuda, T., 2005. Control of acid-site Xu, T., Barich, D.H., Goguen, P.W., Song, W., Wang, Z., Nicholas, J.B., Haw, J.F., 1998.
location of ZSM-5 zeolite membrane and its application to the MTO reaction. Synthesis of a benzenium ion in a zeolite with use of a catalytic flow reactor.
Catal. Today 105, 662e666. J. Am. Chem. Soc. 120, 4025e4026.
Tao, X., Wang, J., Li, Z., Ye, Q., 2013. Theoretical study on the reaction mechanism of Yaripour, F., Shariatinia, Z., Sahebdelfar, S., Irandoukht, A., 2015. Effect of boron
CO2 hydrogenation to methanol. Comput. Theor. Chem. 1023, 59e64. incorporation on the structure, products selectivities and lifetime of H-ZSM-5
Teketel, S., Svelle, S., Lillerud, K.-P., Olsbye, U., 2009. Shape-selective conversion of nanocatalyst designed for application in methanol-to-olefins (MTO) reaction.
methanol to hydrocarbons over 10-Ring unidirectional-channel acidic H-ZSM- Microporous Mesoporous Mater. 203, 41e53.
22. ChemCatChem 1, 78e81. Yu, M., Zhu, Y.A., Lu, Y., Tong, G., Zhu, K., Zhou, X., 2015. The promoting role of Ag in
Teketel, S., Skistad, W., Benard, S., Olsbye, U., Lillerud, K.P., Beato, P., Svelle, S., 2011. Ni-CeO2 catalyzed CH4eCO2 dry reforming reaction. Appl. Catal. B Environ. 165,
Shape selectivity in the conversion of methanol to hydrocarbons: the catalytic 43e56.
performance of one-dimensional 10-ring zeolites: ZSM-22, ZSM-23, ZSM-48, Yurchak, S., 1988. Development of Mobil's fixed-bed methanul-to-gasoline (MTG)
and EU-1. ACS Catal. 2, 26e37. process. Stud. Surf. Sci. Catal. 36, 251e272.
(2015) Methanol, Methanol Uses. The Essential Chemical Industry Online. Article, Zaidi, H.A., Pant, K.K., 2004. Catalytic conversion of methanol to gasoline range
The University of York. Data adopted from Methanol Market Services Asia hydrocarbons. Catal. Today 96, 155e160.
(MMSA), 2014. Zaidi, H.A., Pant, K.K., 2008. Activity of oxalic acid treated ZnO/CuO/HZSM-5 catalyst
Tijmensen, M.J., Faaij, A.P., Hamelinck, C.N., van Hardeveld, M.R., 2002. Exploration for the transformation of methanol to gasoline range hydrocarbons. Indust. Eng.
of the possibilities for production of Fischer Tropsch liquids and power via Chem. Res. 47, 2970e2975.
biomass gasification. Biomass Bioenergy 23, 129e152. Zhang, J., Zhang, H., Yang, X., Huang, Z., Cao, W., 2011. Study on the deactivation and
Topp-Jørgensen, J., 1988. Topsøe integrated gasoline synthesis e the TIGAS process. regeneration of the ZSM-5 catalyst used in methanol to olefins. J. Nat. Gas
Stud. Surf. Sci. Catal. 36, 293e305. Chem. 20, 266e270.
Toyir, J., Miloua, R., Elkadri, N.E., Nawdali, M., Toufik, H., Miloua, F., Saito, M., 2009. Zhang, J., Zhang, T., Zhang, X., Liu, W., Liu, H., Qiu, J., Yeung, K.L., 2014. New synthesis
Sustainable process for the production of methanol from CO2 and H2 using Cu/ strategies for Ni/Al2O3-Sil-1 core-shell catalysts for steam reforming of
ZnO-based multicomponent catalyst. Phys. Procedia 2, 1075e1079. methane. Catal. Today 236, 34e40.
Tsyganok, A.I., Tsunoda, T., Hamakawa, S., Suzuki, K., Takehira, K., Hayakawa, T., Zhou, J., Liu, Z., Li, L., Wang, Y., Gao, H., Yang, W., Xie, Z., Tang, Y., 2013. Hierarchical
2003. Dry reforming of methane over catalysts derived from nickel-containing mesoporous ZSM-5 zeolite with increased external surface acid sites and high
MgeAl layered double hydroxides. J. Catal. 213, 191e203. catalytic performance in o-xylene isomerization. Chin. J. Catal. 34, 1429e1433.
Vita, A., Cristiano, G., Italiano, C., Pino, L., Specchia, S., 2015. Syngas production by Zyryanova, M.M., Snytnikov, P.V., Shigarov, A.B., Belyaev, V.D., Kirillov, V.A.,
methane oxy-steam reforming on Me/CeO2 (Me¼Rh, Pt, Ni) catalyst lined on Sobyanin, V.A., 2014. Low temperature catalytic steam reforming of propane-
cordierite monoliths. Appl. Catal. B Environ. 162, 551e563. methane mixture into methane-rich gas: experiment and macrokinetic
Wan, Z., Wu, W., Chen, W., Yang, H., Zhang, D., 2014. Direct synthesis of hierarchical modeling. Fuel 135, 76e82.

You might also like