Vous êtes sur la page 1sur 20

Fuel Processing Technology 145 (2016) 42–61

Contents lists available at ScienceDirect

Fuel Processing Technology

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

Review

Synthesis of methanol from methane: Challenges and advances on the


multi-step (syngas) and one-step routes (DMTM)
Marcio Jose da Silva
Chemistry Department, Federal University of Viçosa, Viçosa, Minas Gerais 36570-000, Brazil

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

Article history: To transform the methane oxidation to methanol in a selective, straight, economically attractive, and less energy-
Received 26 October 2015 intense process is a goal pursued by the industry since its discovery. Methane is the main constituent of shale and
Received in revised form 13 January 2016 natural gas while methanol is either a fuel as feedstock in the chemical industry. Thus, to develop a technology
Accepted 19 January 2016
that combines an affordable raw material with a strategic product became pivotally important for the chemical
Available online xxxx
industry. Currently, the industrial route for methanol production from methane is accomplished via a syngas pro-
Keywords:
cess, passing by stream catalytic reform of products (i.e., CO and H2). This is a costly route due to its high-energy
Methane consumption. Alternatively, methane partial oxidation to methanol (i.e. DMTM route) in a single step can consti-
Methanol tute a more economically viable strategy. Toward achieving this goal, different approaches were proposed; none-
Syngas theless, until now, any was industrially feasible. In this review, we paid particular attention to the methane direct
Natural gas oxidation processes to methanol carried out in a gas phase under homogeneous or heterogeneous conditions. In
Solid catalysts general, heterogeneous processes are solid-catalyzed in the gas phase, while homogeneous processes occur with-
out a catalyst. We too assessed the advances achieved in the traditional route to producing methanol from syngas,
as well as recent developments of syngas production from methane.
© 2016 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
1.1. Natural gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2. Syngas-based multistep process for the production of methanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.1. A brief overview of methane conversion to syngas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3. Oxidation of methane to methanol in a multistep processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.1. Brief background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.2. Conventional production route of methanol from syngas: from BASF to Synetix process . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.3. Cu–Zn/Al2O3 catalyst: Synetix process for methanol production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.4. Main challenges of the conventional production process of methanol from syngas . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4. Oxidation of methane to methanol in a one-step process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.1. Main challenges of direct methane partial direct oxidation to methanol routes (DMTM routes) . . . . . . . . . . . . . . . . . . . . . . 49
4.2. Activation of C–H bonds by enzymatic complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.3. DMTM route (i): free-catalyst process of methane direct partial oxidation to methanol by oxygen (i.e. homogeneous process) . . . . . . . . 50
4.3.1. The use of other oxidants in catalyst-free process of methane direct partial oxidation to methanol . . . . . . . . . . . . . . . . . 50
4.4. DMTM route (ii): solid-catalyzed processes of direct oxidation of methane to methanol . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.4.1. Copper–zinc/alumina catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.4.2. Zeolite catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.4.3. Molybdenum catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.4.4. Iron sodalite catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.4.5. Other solid catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5. Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

E-mail address: silvamj2003@ufv.br.

http://dx.doi.org/10.1016/j.fuproc.2016.01.023
0378-3820/© 2016 Elsevier B.V. All rights reserved.
M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61 43

Fig. 1. Conversion route of natural gas to chemicals and fuels [3].

1. Introduction highly accessible consumed it as fuel for industrial and domestic


heating, or electric power generation. Consequently, almost all methane
1.1. Natural gas produced has become an energy source [4].
Rather than be burned as fuel, natural gas is potentially useful as
Natural gas is a clean and effective energy source because its com- an industrial feedstock. Although there are different approaches to
bustion generates fewer greenhouse gases than coal or petroleum liquid using the methane present in the natural gas as a raw material, the
fuels. However, to store and transport natural gas from remote sites up- most explored and economically viable is its transformation to syn-
holds a great challenge, making it difficult to make the prices of its de- gas (i.e., CO and H 2 ). Industrially, methane has been upgraded to
rivatives (i.e., chemicals or liquid fuels) competitive in relation to syngas throughout multistep processes that use steam and/or
those of fossil oil [1]. Nevertheless, the inevitable depletion of petroleum oxygen (i.e., steam reforming, autothermic reforming, or partial oxi-
reserves and the recent discovery of natural gas sites may make it more dation) [5].
economically attractive, and fulfill society's demand for alternative en- As shown in Fig. 1, syngas is a starting material to valuable chemicals
ergy sources [2]. (i.e., dimethyl ether, formaldehyde, acetic acid) or liquid fuels (i.e., via
Different from shale gas, which contains other hydrocarbons, the Fischer–Tropsch catalysis). Nevertheless, the most attractive routes
natural gas comprises only methane as a major component. Currently, are those that involve a single step, such as oxidative coupling of meth-
the most of the methane is consumed as fuel, resulting in greenhouse ane (i.e., OCM), homologation, aromatization, and methane partial oxi-
gas (i.e., CO and CO2). Indeed, methane itself is too more deleterious dation to methanol (i.e., DMTM). This latter constitutes the focus of this
molecule to the atmosphere, and also contributes to the greenhouse ef- review.
fect. Nonetheless, it can also provide clean fuels, such as hydrogen or In this work, we paid special attention to two processes of methanol
methanol. production from methane: the first, the syngas-based multistep pro-
Methanol itself can be converted to gasoline and valuable fine cess; and the second, the single-step process, which directly transforms
chemicals throughout catalytic processes (Fig. 1) [3]. Natural gas con- methane into methanol (i.e. DMTM route).
sumption in this century has exponentially increased compared to
other energy sources (Fig. 2) [4]. Indeed, countries, where methane is 2. Syngas-based multistep process for the production of methanol

2.1. A brief overview of methane conversion to syngas

After exhaustive research, Green et al. grouped three main types of


catalysts as the more effective for the conversion of methane to syngas:
(1) supported nickel, cobalt, or iron catalysts, (2) supported noble metal
and (3) transition metal carbide catalysts [6,7]. However, herein we will
focus only on supported Ni catalysts.
The reforming of methane over Ni catalysts with steam, carbon diox-
ide, or oxygen, separately, generates syngas at different stoichiometries
[8]. Table 1 displays these reactions and their enthalpy obtained at
1173 K [8].
Currently, the syngas route has been the only economically viable
process for the conversion of methane to chemicals or high value-

Table 1
Process for syngas production from natural or shale gasa.

Process Reaction H2:CO molar ΔH1173 K


ratio (kJ mol−1)

Steam reforming CH4 + H2O → 3 H2 + CO 3:1 225.7


CO2 reforming CH4 + CO2 →2 H2 + 2 CO 1:1 258.8
Fig. 2. Projection of world energy consumption by fuel type 2000 to 2020.
Oxygen reforming CH4 + 0.5 O2 → 2 H2 + CO 2:1 −23.1
Adapted from U.S. Energy Information Administration/International Energy
a
Outlook, 2005 [4]. Adapted from Ref. [8].
44 M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61

Fig. 3. Production of syngas from traditional methane steam reforming.


Adapted from Ref. [10].

added liquid fuels; nevertheless, syngas production is unquestionably Significant advances in the use of promoters (i.e., B, La, and Rh) have
the most expensive step of this production chain (Fig. 3) [8]. outstandingly improved the activity of Ni catalysts on the syngas gener-
The reaction of methane present in the natural or shale gas with ation process [16]. On the other hand, the replacing of alumina supports
steam, oxygen, or a mixture of them, provides syngas at industrial by ceria, zirconia, and silica, all doped with rhodium nanoparticles, sig-
scale (Fig. 3). Lunsford et al. studied the activity of alumina-supported nificantly enhanced the process of syngas production through steam
nickel on conversion of methane to syngas in the temperature ranging methane reforming [17].
of 720 to 1173 K [9]. Several alternative routes have tried to circumvent the high costs as-
Industrially, in this method, steam reforming of methane over sociated with large-scale methanol plants. The Topsøe Group developed
Ni/− Al2O3 catalyst heated to 1173 K occurs in a primary reactor. The a technology that employs the concept known as “two-step reforming”
unreacted methane is then reformed with oxygen and steam in a sec- for syngas production (Fig. 5) [1]. This conventional process makes
ondary reactor, producing a mixture of CO and H2 in equilibrium methanol from syngas since the last century.
(Fig. 3) [10]. As depicted in the layout of Fig. 5, this traditional route includes adi-
During the next two steps, steam is then added to the syngas gener- abatic pre-reforming, tubular reforming, and oxygen-blown secondary
ated under milder thermal treatment than earlier (ca. 673 K), over iron reforming. In this process, oxygen is the source for internal combustion
oxide or copper catalysts. Throughout these water gas shift stages, the of hydrocarbons. The two-step reforming process combines a fired tu-
addition of steam may then adjust the H2:CO molar ratio to ones bular reforming and oxygen-fired adiabatic reforming allows obtaining
required for the further use of syngas [10,11]. In general, the syngas syngas at most suitable composition. On methanol synthesis process,
stoichiometry is handled toward its use as feedstock for diesel synthesis the quantity of hydrogen and carbon oxides in the syngas should be
via Fischer–Tropsch process (Fe2O3 catalyst) or methanol synthesis properly balanced, to reduce the consumption per unit methanol pro-
(Cu–ZnO/Al2O3 catalyst) [11]. Although not shown, these reaction con- duction rate.
ditions produce coke, resulting in catalyst deactivation. The use of Ni In two-step reforming process, the primary reforming produces ex-
supported over rare earth oxides or alkaline metals minimizes this un- cess of hydrogen, while secondary reforming has use this excess hydro-
desirable reaction, and improves the lifetime of the catalyst restricting gen on step of hydrogenation of CO to produce methanol. The
the carbon deposit [12–15]. combination of the two types of reforming creates an energy efficient
Alternatively, methane can be reformed with a steam and oxygen creation of synthesis gas.
mixture heated to a high temperature (ca. 2273 K) without catalyst.
These conditions favor radical reaction commonly called as “homoge- CO + 2 H2 → CH3OH (−ΔH298 K, 50 bar = 90.7 kJ mol−1)
neous oxidation” (Fig. 4).
The following step occurs in another reactor, where the mixture ob- On the other hand, the use of stand-alone autothermal reforming
tained is then reformed over Ni catalysts resulting in syngas and water, (ATR technology) together with a suitable selection of reaction condi-
which can be processed under water gas shift reaction conditions. tions and suitable catalysts can produce methanol with high availability

Fig. 4. Production of syngas from methane autothermal reforming.


Adapted from Ref. [18].
M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61 45

Fig. 5. Methanol process layout with Topsøe's two-step reforming.


Adapted from Refs. [18, 19].

conditions and viability of operation. The ATR technology operates at ATR Technology has great advantageous such as avoidance of the supply
low steam to carbon molar ratio, thus decreasing the flow through the or dissipation of thermal energy to or from the reaction [18]. Arguably,
plant and diminishing the investment. this characteristic makes ATR suitable to industrially provide syngas
Stand-alone autothermal reforming (ATR) technology merges par- with high quality in large-scale plants.
tial combustion and catalytic steam reforming in one compact refracto- Stand-alone autothermal reforming (ATR) technology comprises a
ry lined reactor to yield syngas for methanol production in large scale. stand-alone reformer and oxygen-fired reformer. The autothermal

Fig. 6. Syngas production equipped with Topsøe's ATR stand-alone reforming.


Adapted from Refs.18, 19].
46 M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61

Scheme 1. Equations of reactions to produce methanol and syngas.


Adapted from Ref. [31].

reformer comprises a burner, a combustion zone, and a catalyst bed IV. The energy releasing should be well employed;
in a refractory lined pressure vessel. The burner mixes the feed and V. It is required that reactants and residual aftermaths can be ade-
the oxidant. In the combustion zone, the feed and oxygen are quately disposed.
burned.
The catalyst bed brings the steam reforming and shift conversion of
reactions to equilibrium. By adjusting of low molar ratio of steam to The main positive figures of the stand-alone ATR reforming process
carbon, ATR plants can run similar to a two-step reforming plant but are the high rates of methanol reaction (i.e. due to syngas with high
with a single incoming stream. CO/CO2 proportion), the low steam throughput and low necessity of
Nowadays, ATR technology replaces “two-step reforming process” steam that reduce investment cost and operation of large scale plants.
by a Topsøe's optimized solution, aiming large-scale methanol pro-
duction. The route based on “stand-alone ATR technology” avoids
the use of a tubular reformer, required in the old two-step process 3. Oxidation of methane to methanol in a multistep processes
(Fig. 6).
Topsøe's ATR consists of a fixed bed reactor (i.e. Ni-based on ceramic 3.1. Brief background
rings and pellets), where the reforming process takes place. Therefore,
unlike the two step process, in a stand-alone ATR reforming process Giulio Natta, who won the Nobel Prize in Chemistry in 1963, stated,
the natural gas is pre-reformed and straightly sent to an ATR reformer, “among the main industrial organic reactions, the synthesis of methanol
where a mixer/burner burning all the hydrocarbons with oxygen and is an outstanding example of practical importance of catalytic process-
steam. In the combustion chamber, partial combustion reactions es” [22]. Undeniably, until the half of the twenty century, most metha-
occur, followed by methane steam reforming reaction and shifting of nol had a natural origin (i.e., waste woods distillation). Conversely, the
equilibrium conversion over the catalyst bed. Because tubular reformer petrochemical industry today is responsible by the entire methanol
is not required, in a stand-alone reforming process the addition of consumed.
steam to the feed stream is drastically decreased [20]. Therefore, Although its high toxicity or lower energy content than other liquid
the stand-alone ATR technology operates at low ratio molar steam hydrocarbon fuels, methanol has employ as a pure liquid fuel or gaso-
to carbon. line blend [3,23]. Moreover, it is also a feedstock to synthesize commod-
However, though the syngas generated from stand-alone ATR ities, such as methyl terc-butyl ether (ca. 28%), formaldehyde (ca. 34%),
reforming process is more reactive than that obtained from two-step acetic acid (ca. 7%), and other chemicals, such as methyl acrylate or sol-
reforming, it is less reactive than syngas produced from gasification vents (ca. 31%) [24–27]. Moreover, MTO (i.e., methane to olefins) and
route. For this reason, currently the most of industrial units produce MTG (i.e., methane to gasoline) processes that produce light olefins or
methanol from syngas obtained through gasification process. The low gasoline, respectively, start from methanol [28,29].
reactivity is not the main challenge of manufacturing methanol from Industrially, methanol has been produced throughout the syngas
syngas obtained via stand-alone ATR. Indeed, the integration between route, which involves two consecutive steps: the first, in which meth-
the reforming units and the methanol synthesis is the greatest hamper. ane is steam reforming over Ni/Al2O3 catalysts resulting in syngas
To overcome these drawbacks, we should considerer some points (Fig. 4), and the second, in which the syngas is then converted to meth-
[21]; anol over Zn–CuO/Al2O3 catalyst (see the next sections).
Currently, few industrial processes synthesize methanol (i.e., Jonhson
I. The process will operate with syngas having a high CO/CO2 molar Matthey, Lurgi, Mitsubishi, or Kellogg processes; Fig. 5). Indeed, all of
ratio; them employ Zn–CuO/Al2O3 catalyst, and operate under pressures of
II. The process is highly exothermal and should be operated at low 50 to 100 atm over a range of temperature of 473 to 573 K. Although
temperature; are highly intensive-energy routes, this catalytic process produces meth-
III. Side-product formation should be reduced; anol with 99% selectivity and energy efficiency higher than 70% [30].

Scheme 2. Methanol production from syngas route.


M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61 47

3.2. Conventional production route of methanol from syngas: from BASF to sensitive to poisoning by chloride and mainly sulfur, and thus re-
Synetix process quires pre-purification steps, mainly when the feedstock is the
shale gas.
Three equations describe the methanol production through the syn- From a mechanical perspective, the methanol synthesis over Cu–
gas route (Eqs. (1)–(3); Scheme 1). Methanol synthesis (Eqs. (1) and ZnO/Al2O3 via the syngas route has many issues that are currently the
(2)) is an exothermic process that involves a decreasing number of gas- subject of some controversy. For instance, identifying the nature of ac-
eous mol. Consequently, the reactions carried out at low temperatures tive sites or the methanol precursor (i.e., CO or CO2) and the main
and high pressures reach the maximum conversion [31]. adsorbed intermediate species, as well as determining the reaction
The BASF process (named as a “high pressure process”) convert syn- pathway or the rate-determining step, are actively discussed topics [37].
gas to methanol over ZnO–Cr2O3 catalysts, operating under high pres- Kinetic studies, besides investigations about the adsorption of
sures (ca. 250 to 350 bar) and in a temperature range of 573 to 673 K isotopic-labeled species on the surface of different catalysts could be
[32]. This catalyst is highly tolerant to sulfur, a poison present in large useful to understand the methanol formation from the CO2 and H2
amounts in shale gas, which was a feedstock widely used in the begin- adsorbed in the catalyst surface [38–40]. The combination of these
ning of the twentieth century. adsorbed molecules provides the adsorbed format (H2COOads) [38–40].
Nonetheless, the BASF process employs drastic reaction conditions The most probable rate-limiting step is the hydrogenolysis of the
for methanol production. For this reason, several researchers intensively adsorbed format (i.e., H2COOads) to give methoxy species (i.e., CH3Oads)
worked to achieve milder reaction conditions. During the first thirty that capture a proton, which then generates methanol [16].
years of the past century, copper oxide catalysts permitted a lowering On the other hand, determining how to increase the catalysts' toler-
of the temperature and pressure of the BASF process. However, copper ance to the poisons (i.e., chloride or sulfur) or thermal sintering is still a
catalysts possessed a major drawback, i.e., high sensitivity to poison great challenge. The poisoning and sintering of catalyst result in the de-
by sulfur. activation and consequent stopping of the process. In this regard, during
Imperial Chemical Industries, Ltd. (ICI) developed syngas purifica- methanol synthesis, the catalyst undergoes deactivation even in the ab-
tion systems. They discovered that Cu–ZnO catalyst was much more ac- sence of a poison; catalyst activity drops to less than 30% after the first
tive than ZnO–Cr2O3, even though the former continued to be easily thousand hours of work [41]. Studies revealed that even in the absence
poisoned by sulfur [33,34]. The development of effective purification of steam or carbon dioxide, Cu/ZnO catalyst was irreversibly deactivated
systems and active catalysts resulted in the process currently used, due to the reduction of CuO catalyst to Cu2O [42]. Therefore, increase the
which works over Cu–ZnO/Al2O3 catalyst to convert “metgas” catalyst lifetime is a key issue that determines economic feasibility of
(i.e., syngas with an adequate molar ratio between CO and H2) to meth- the process.
anol under pressure of 50 to 100 bar and temperatures in the range of Although it is easily synthesized by alkaline co-precipitation of metal
513 to 533 K (Scheme 2) [10]. nitrates (i.e., Cu, Zn, and Al), it is difficult to reproduce the industrially
used catalyst (i.e. Cu/ZnO). It complicates evaluating of catalyst stability
3.3. Cu–Zn/Al2O3 catalyst: Synetix process for methanol production at the laboratory scale. In addition, there is another experimental prob-
lem. The heating to 450 to 510 K activates catalyst oxides under a reduc-
This process was the first route commercialized for methanol ing mixture diluted with inert gas (i.e., natural gas and N2). This mixture
production from syngas at low pressure, and was labeled as the “ICI commonly contains hydrogen, generated during the manufacture of
process”; however, it is currently named as the “Synetix process” [10]. syngas. Furthermore, the presence of hydrogen reduces only CuO
Initially, it was accepted that the Cu(0) species constituted the active oxide to metallic copper, whereas to ZnO or Al2O3 that remains as
sites. Nonetheless, study showed that other phases also played impor- oxides [27].
tant role in the activity and catalyst lifetime. Indeed, Nonneman and The catalyst activity depends on the high copper surface area or on
Ponec demonstrated that pure Cu is an inactive catalyst in methanol the small crystallite size, which are responsible by the contact between
synthesis [35]. They concluded that Cu(I) ions are formed throughout copper and the promoter zinc oxide on the alumina support [43]. Al-
the process and are stabilized by promoters (i.e., ZnO, CsCO 3 ) on though the true role of zinc and copper oxides in the mechanism of
the Cu(0) surface, which is supplied by adsorbed hydrogen atoms the methanol synthesis remains unclear, Spencer et al. proposed that
(i.e., Hads). only when adsorbed hydrogen concentration on copper phase is low,
The role of zinc oxide on Cu–ZnO/Al2O3-catalyzed methanol syn- the hydrogen dissociation on zinc oxide is important [36,44].
thesis has been goal of study. Zinc oxide alkalinity may minimize the The tolerance to the poisoning and the thermal stability are features
action of the acidic sites on the alumina phase, which would pro- that affect the activity and lifetime of catalysts. Spencer et al. reported
mote methanol conversion to dimethyl ether, and under certain that when the methanol is synthesized over Cu–ZnO/Al2O3 catalysts
conditions, hydrogen supply by spillover from ZnO to the CuO [35]. containing promoters, the formation of coking or the poisoning of cata-
This could be highly desirable for methanol synthesis over Cu–ZnO lyst are minimized; therefore, only thermal sintering provokes catalyst
catalysts [36]. It is noteworthy that Cu–ZnO/Al 2 O 3 catalyst is deactivation [45].

Scheme 3. Methane steam reforming coupled to metgas synthesis.


Adapted from Ref. [46].
48 M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61

Scheme 4. Block diagram of the conventional production process of methanol.


Adapted from Ref. [47].

An adequate adjustment of the syngas stoichiometry to generate serious drawbacks, such as high consumption of energy and requiring
metgas constitutes an important stage throughout the conventional pro- expensive infrastructure. Moreover, the large amount of steam needed
duction process. While syngas is produced by steam methane reforming results in the reactors' corrosion and difficulty with handling. For
at a CO:H2 molar ratio equal to 3:1, methanol synthesis requires a these reasons, the price of liquid methanol (i.e., produced via the syngas
2:1 molar proportion between CO and H2 (see Scheme 3) [46]. route) compared to petroleum derivative liquid fuels tends not to be
The metgas has been manufactured from methane after an additional competitive [30]. Indeed, the syngas process is responsible for 60 to
step, in which the homogeneous methane partial oxidation with oxygen 70% of methanol production costs [50].
(i.e., POX) and metal-catalyzed methane steam reforming (i.e., SR) are
consecutively performed (Scheme 2, step III). This extra step implies 4. Oxidation of methane to methanol in a one-step process
adding additional costs to the process [46].
The straight conversion of methane present in natural gas into meth-
anol is a process that could avoid the highly energy-dependent conven-
3.4. Main challenges of the conventional production process of methanol tional route, reducing the number of stages, and thus avoiding the large
from syngas capital investment required to build a syngas industrial plant. This route
may constitute an economically viable technology and a more environ-
Scheme 4 presents a block diagram of the conventional production mentally benign process than methanol production via the syngas path.
process of methanol from syngas over Cu–ZnO/Al2O3 catalyst [47]. Undeniably, such development could be an achievement that could
First, it is important to highlight that although methanol synthesis notably expand the production of methanol-derivatives and influence
from syngas is exothermic, the overall enthalpy of the reactions 1 to 3 the planet's economy [48]. However, until today, these processes are
(Eqs. (1)–(3), Scheme 5) is not favorable in terms of feasibility of meth- far from being considered as established or industrially practical due
anol synthesis from syngas under standard conditions [48]. to numerous reasons.
Replacing steam by CO2 could minimize process costs because Extensive research has been devoted to the direct oxidation of meth-
syngas is provided at the equimolar amount of CO and H2 (Eq. (4), ane to methanol, which comprises the following technologies cited in
Scheme 5). However, the enthalpy variation of this reaction is higher items (i) to (v) [51].
than that of steam methane reforming (Eq. (1), Scheme 5). Really, this
process is inefficient and requires high temperatures (ca. as high as (i) Free-catalyst homogeneous processes at high temperatures
1000 K), which promote CO disproportionality and result in the unde- based on radical reactions in the gas phase;
sirable formation of coke. All these aspects become the replacement of (ii) Solid-catalyzed processes in the gas phase;
steam by CO2 economically infeasible [48,49]. (iii) Solid-catalyzed processes in the liquid phase;
Despite large capital investment to build industrial plants, the syn- (iv) Homogeneous catalytic processes in the liquid phase, in the
gas route constitutes the most important process to convert methane presence of soluble catalysts;
to clean fuels, such as methanol [49]. Nevertheless, it presents some (v) Enzymatic catalytic processes.

Scheme 5. Standard enthalpy of methanol synthesis from syngas.


Adapted from Ref. [48].
M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61 49

Throughout this review, we will attempt to pay special attention to improve our understanding about how replicate this reaction at indus-
methane partial oxidation to methanol in “free-catalyst processes” trial scale [56].
(i.e., homogeneous processes based on radical chain reactions in the Methanotrophic bacteria have two methane monooxygenase
gas phase), and “catalytic processes” (i.e., activation of C–H bonds by (MMO) enzymes that convert methane to methanol at room tempera-
bioinorganic complexes, solid metal catalyzed-methane partial oxida- ture; soluble MMO (sMMO) and particulate MMO (pMMO) during the
tion in the gas phase). first step of metabolic cycle (Eq. (5)) [57].

MMO
4.1. Main challenges of direct methane partial direct oxidation to methanol CH4 þ O2 NADH þ Hþ → CH3 OH þ H2 O NADþ ð5Þ
routes (DMTM routes)
Literature describes that pMMO is a metalloenzyme that con-
The reaction is thermodynamically spontaneous even at room tem-
tains multi-copper active sites that are found in the plasma of all
perature (Scheme 4, Eq. (3)). However, since methanol is less stable
methanotrophic bacteria, responsible by its activity. Several au-
than the over oxidation products, to develop a direct route to make
thors have proposed that active sites might contain copper at a
methanol from methane under room conditions has been referred to
wide range of stoichiometries (i.e., two or three mono- or divalent
as the “holy grail” due to its immense difficulty [52] (Table 2).
copper ions) [58–60]. Indeed, the level of copper ions is determin-
Methane reactivity is lower than methanol due to its stronger C–H
ing to decide which enzyme will be expressed; if low concentration
bonds. The dissociation energy of the C–H bond of methanol is
of copper cations is feasible, the enzyme expressed is sMMO;
393 kJ mol−1; whereas of methane, where it is 440 kJ mol−1. Therefore,
whereas, at high copper concentrations the pMMO enzyme is the
it is easier to oxidize methanol to stable products of over oxidation
active site [61,62].
(i.e., CO or CO2) than to oxidize methane itself. It becomes difficult to
Conversely, only the action mechanism of sMMO is well known. A
control the selectivity of the one-step process of methane oxidation to
large number of spectroscopy study show that the sMMO enzyme con-
methanol (Table 2). Thus, because the methanol yield is always low,
tains a carboxylate-bridged diiron unit [61–63]. The atmosphere oxygen
the direct conversion route of methane to methanol is commercially in-
is the oxidant used by both enzymes to oxidize methane to methanol.
feasible [53]. Table 2 presents the effect of temperature on the straight
Kinetic parameters suggest that pMMO is more specific for methane
oxidation of methane into methanol [48].
oxidation than sMMO enzyme.
Thermodynamically, it is possible to oxidize methane to methanol at
Nowadays, there is a growing interest on the pMMO-mediated oxi-
room temperature [48]. Indeed, the highest conversion theoretically
dation reactions which involves the reactivity of copper oxygenase en-
achieved when the equilibrium is reached at room temperature
zymes. Particularly, the developing of catalysts for oxidation of
(i.e., 298 K) is near 33%, calculated based on Gibbs free energy [43]. Con-
methane to methanol at industrial scale can be improved by under-
versely, the best selectivity was obtained after maximum conversion is
standing how pMMO enzyme works. Consequently, several biomimetic
near 5%. This constitutes a greatest challenge because the carbon yield
studies have been carried out. The protein comprises three distinct sub-
of the conventional syngas process is quite superior to this value (ca.
units (i.e., PmoA, PmoB, and PmoC).
70 to 75%) [54].
On this sense, Chan et al. proposed that the catalytic site of pMMO
Deserves highlight that utilizing of molecular oxygen as oxidant is a
enzyme might be a trinuclear copper cluster, and developed a series of
key aspect to do competitive any commercial DMTM (i.e., direct meth-
tricopper complexes that are capable of catalytic oxidize the methane
ane to methanol) process. Molecular oxygen is an inexpensive, environ-
to methanol under ambient conditions [64,65]. In according with their
mentally benign, and highly affordable reactant, which are essential
experiments, those authors reported that two tricopper complexes con-
features for large-scale methane conversion to liquid fuels or chemicals
verted methane to methanol: a tricopper-peptide complex derived
[55]. However, because the molecules of methanol and methane
from pMMO and a biomimetic model tricopper complex [66].
possess singlet ground states, their oxidation by oxygen is a spin-
Chan et al. assessed the oxidation of methane to methanol medi-
forbidden reaction, and thus requires an adequate catalyst to
ated by the tricopper complex [Cu 1 +–Cu1 +–Cu1 +(7-N-Etppz)]1 +
activate them [52].
in acetonitrile solutions. The ligand (7-N-Etppz) correspond to the
In this regard, a solid catalyst could be an attractive option. Nonethe-
3,3′-(1,4-diazepane-1,4-diyl)bis[1-(4-ethyl[piperazine-1-yl)propan-2-
less, methanol has a dipole moment higher than methane. Therefore, it
ol] is showed in Fig. 9, besides catalytic cycle proposed by those authors
is preferentially activated over solid metal catalyst surface and is totally
[64–66].
oxidized, compromising thus the reaction yield. In addition, this strong
Chan et al. reported that a turnover number equal to 0.92 was
adsorption implies in additional recovery steps, where a polar solvent to
achieved when tricopper complex was activated by oxygen molecular
extract the adsorbed methanol.
in the presence of methane excess. Nonetheless, the process becomes
catalytic by addition of hydrogen peroxide, which recovery the spent
4.2. Activation of C–H bonds by enzymatic complexes catalyst, after tricopper complex transfer oxygen atom to the methane
(Fig. 10).
Nature has a notable ability in to do that the chemists pursue since Chan et al. pointed that because tricopper-peptide complex is insol-
long time. For these reasons, an extensively work has been done aiming uble in aqueous buffer, it was encapsulated in mesoporous carbon.
XANES and EPR spectra demonstrated that copper(I) cations on the
Table 2
tricopper peptide were quickly reoxidized by air or oxygen, similarly
Temperature effects on Gibbs free energy for products oxidation of methanea. to the observed on putative tricopper cluster in pMMO. Therefore, it
was clearly established that Cu1 +–Cu1 +–Cu1 +–peptide complex is
Reaction ΔG values
able to mediate transfer of oxygen atom from molecular oxygen to
Temperatures (K) methane at room temperature [64–66]. This biomimetic approach
298 650 800 1000 allows formulating either homogeneous or heterogeneous catalysts to
CH4 + 0.5 O2 → CH3OH −111 −93 −86 −76 oxidize methane to methanol [67].
CH4 + O2 → HCHO −288 −294 −295 −298 Inspired by these findings, a number of researchers have tried to
CH4 + 1.5 O2 → CO + 2 H2O −544 −573 −582 −603 build active sites similar to those of sMMO or pMMO enzymes into
CH4 + 2 O2 → CO2 + 2 H2O −801 −800 −799 −798 cavity of zeolites. In the near brief future, Cu–ZSM-5 zeolite could poten-
a
Adapted from Ref. [48]. tially serve as a model for the building of selective active sites to
50 M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61

Scheme 6. Enthalpy variation of the reactions involved in methane and methanol oxidation at 298 K [72].

understand the working mechanisms of enzymes [68]. For instance, re- equilibrium controlling variables such as temperature and reactant
actions carried out over Cu–ZSM-5 zeolite achieved high methanol se- pressure, aiming shift selectivity toward methanol. High methane pres-
lectivity (ca. 98%) [69]. sure can prevent deep oxidation if adequate temperatures and pressure
Another highlighted instance was recently described by Lercher are employed [72–74].
et al., which demonstrated that single-site trinuclear copper oxygen Zhang et al. described that other process control parameters, includ-
clusters in mordenite was an efficient catalyst for selective conversion ing reactor type, temperature, feed oxygen concentration, and gas flow
of methane to methanol [70]. Copper-exchanged mordenite zeolite also affect the efficiency of the catalyst-free gas phase reactions [75].
mimics both nuclearity and reactivity of pMMO enzyme active sites. Those authors verified that pressure exerts a pronounced effect on
Lercher et al. verified that multiple Al framework atoms of the methanol selectivity. Under methane pressure equal to 5.0 MPa and
mordenite microporous stabilize trinuclear copper-oxo clusters (i.e. heating to a temperature of 703 or 743 K, reactions gave 30 and 40% se-
[Cu3(μ-O)3]2 +), which are able to activate carbon–hydrogen bond in lectivity at conversions of 5 and 10%, respectively. They inhibited the
the methane, and its conversion to methanol. They demonstrated that production of CO2, insulating the ringed gap between the inner quartz
reversible rearrangements of the trinuclear clusters occurring during line and the reactor tube. Indeed, they realized that both reactants and
the selective transformations of methane to methanol in both enzymat- products must be isolated from contacting the metal wall of the reactor;
ic and copper-exchanged mordenite catalysts [70]. These results corrob- for this reason, the best result was obtained when the methane partial
orate with those reported by Chan et al. [64–66]. oxidation to methanol is done in quartz and Pyrex glass-lined reactors
[47,75].
4.3. DMTM route (i): free-catalyst process of methane direct partial oxida- There is experimental and theoretical evidence that demonstrates
tion to methanol by oxygen (i.e. homogeneous process) that it is impossible to achieve high methanol yields in these systems,
because an increase on the conversion always produces a decrease on
The relevance of homogeneous processes for methane oxidation the selectivity [76–78]. Consequently, the methanol production
through the radical chain in the gas phase is undeniable. The radical re- throughout homogeneous oxidation of methane will be viable at indus-
actions can occur as a consequence only of working temperature. Hence, trial scale only if the price of raw material (natural gas) is sufficiently in-
depending on the temperature, its contribution to the reaction yield expensive to compensate the high investment cost. Arutyunov et al.
even when solid catalysts are present could be significant. assessed the technological prospects and applications of the direct con-
In the absence of catalysts, methane oxidation to methanol can also version of methane to methanol via free-catalyst reactions in the gas
occur over the reactor wall, in the gas phase, or in both. Since the ener- phase and concluded that these technologies can be useful if employed
gies involved are quite close, distinguishing between the two ways be- at least for low-scale and local applications, mainly if natural gas is
comes difficult. After an initiation step, a series of radical reactions abundant and located in remote sites, where storage and transportation
govern entirely the reaction course, which comprises of propagation challenges exist [79].
and termination steps (or removal of active species).
In general, the main topics of interest to research in the field of ho- 4.3.1. The use of other oxidants in catalyst-free process of methane direct
mogenous oxidation in the gas phase are as follows: (i) optimization partial oxidation to methanol
of high-pressure conditions (ca. higher than 10 bar), aiming an ade- Methane oxidation to methanol with nitrogen oxides has been
quate conversion; (ii) effect of additives; (iii) modeling mechanistic widely described in the literature [79,80]. While in the absence of NO,
studies; (iv) design of novel reactors and (v) use of photochemical ini- DMTM gave only 1% methane conversion at 966 K and room pressure,
tiators [69]. the addition of NO (ca. 0.5%) led to a remarkable increase in methane
The methane partial oxidation to methanol under free-catalyst con- conversion to 10%, even at 808 K. The selectivity to C1-oxygenates was
ditions involves more than 1000 elementary reaction steps with numer- high, but methanol, methyl nitrite, and formaldehyde were always in
ous reactive species [69,71]. Verma et al. developed a complete kinetic relation near 1:0.5:1. The optimum pressure for methanol production
model for these reactions in the gas phase at a temperature range of was 1.0 MPa at 4% conversion of CH4 in gas phase reactions with
308 to 973 K [72]. 0.50% NO, which gave 30% methanol selectivity and virtually no formal-
Those authors show that the rate-limiting step of the partial oxida- dehyde [79,80]. However, the presence of NO2 promoted an increase of
tion of methane is the abstraction of the first hydrogen atom to form
methyl radical. Consequently, photochemical initiators could decrease
the activation energy of this step, favoring the overall reaction [72].
The data of Scheme 6 reveal that while a great amount of energy is
needed to activate the methane, the succeeding reaction also produces
other larger quantities. Clearly, since further steps are highly exother-
mic, it is difficult to stop the reaction where methanol was the only
product. Any attempt to drive the reaction toward high conversions
leads to poor selectivity. The energy gap between reactions suggests
that it is probable to obtain methanol and formaldehyde individually, al-
though from an experimental perspective, this requires a strict reaction Scheme 7. Process scheme for the net oxidation of methane to methanol (X = OSO3H).
control. Therefore, it becomes imperative to adjust the reaction Adapted from Ref. [81].
M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61 51

Table 3
Effects of reaction pressure on product selectivity in the presence or absence of Cu–ZnO/Al2O3 catalysta.
Adapted from Ref. [82].

Exp. Pressure T Catal. Selectivity (%)


(bar) (K) (g)
CH3OH HCHO CO CO2 C2 CH3NO2
(C2H4 + C2H6)

1 4 848 0 20.0 15.5 58.9 4.8 0.5 0.3


2 4 848 0.5 32.1 0 51.8 15.2 0.5 0.4
3 6 798 0 22.0 7.5 59.9 5.6 1 3.2
4 6 798 0.5 26.1 0 58.7 14 0.8 0.4
5 10 798 0 24.7 1.8 65.5 5.2 4.2 0.4
6 10 798 0.5 24.7 0 59.7 8 4.6 1.2
7 30 723 0 20.9 0 60.4 12.6 4 2.1
8 30 723 0.5 24.5 0 55.9 15.1 1.6 2.9
a
Catalyst, Cu:Zn:Al = 39.7:40.8:19.5; flow rate 120 cm3; reaction time = 180 min; catalyst bed temperature 523 K; feed gas composition = CH4 (77.5%); O2 (5.8%); NO (0.5%); Ar
(16.2%).

CO selectivity, caused by over oxidation of methane, methanol, and competes with other DMTM routes that may also performed in liquid
mainly formaldehyde. However, the applicability of this process com- phase (i.e., solid-catalyzed DMTM reactions in the liquid phase or even
pared to the process that employs only oxygen is considerably limited. with soluble catalysts). We summarized the main findings by grouping
Periana et al. described a high-yield system for the low-temperature the solid catalysts in accordance with the metal employed.
conversion of methane to methanol based on Pt(II)-bipym catalyst
(i.e., bpym = bipyrimidine ligand) that operate in liquid sulfuric acid,
4.4.1. Copper–zinc/alumina catalysts
which act as oxidant. A positive aspect of this system is stability of meth-
The high selectivity of Cu–ZnO/Al2O3 catalyst when used in metha-
anol formed to suffer overoxidation in this reaction medium [81]. Con-
nol production via the syngas route also motivated its use in the
versely, the major shortcoming of this solvent system is the difficulty
DMTM reactions. Tabata et al. investigated the optimization of methanol
of separating the methanol from the sulfuric acid. Periana et al. have
yield achieved via DMTM reactions with O2 and NO over Cu–ZnO/Al2O3
found that platinum(II) catalysts were more efficient than Hg(II) cata-
catalysts [82,83]. They demonstrated that although the reactions
lysts previously described by them, and oxidize C–H bond of methane
reached conversions of close to 5.5% with or without catalyst
at temperatures as low as 100 °C. Those authors showed that
(Table 3), in the presence of Cu-ZnO/Al2O3, the selectivity and methanol
bipyrimidine ligand stabilize the reduced species of platinum avoiding
yield increased.
its deactivation.
The best result was obtained under 4 bar of pressure and CH4/O2 =
The positive features of this system are facile hydrolysis of the meth-
8.0 (ca. 2.3% methanol yield). Previously, those authors suggested that
yl bisulfate, the use of recyclable oxidant (i.e., SO2 is easily oxidized to
the formation of CH3OH in reactions over Cu–ZnO/Al2O3-catalysts
SO3), and the use of stable complex catalyst (Scheme 7).
occur through the hydrogenation of HCHO, and the water gas shift reac-
tion (WGSR) [84].
Actually, in a subsequent work, Tabata et al. suggested that three
4.4. DMTM route (ii): solid-catalyzed processes of direct oxidation of meth-
steps: (i) formaldehyde hydrogenation, (ii) formaldehyde dissociation,
ane to methanol
and (iii) the water gas shift reaction, simultaneously occur in methane
oxidation to methanol over Cu–ZnO/SiO2 catalyst, when CH4–O2–NO
It is clear that, in terms of competitiveness, any route of direct oxida-
feed is used [85]. Moreover, they verified that when heating the Cu–
tion of methane to methanol (DMTM route) based on a solid-catalyzed
ZnO/SiO2 catalyst bed to 423 to 623 K, the methanol selectivity almost
process poses a major challenge to operating at lower costs without a
did not change during the reactions, unlike that observed in reactions
loss of selectivity. Additionally, it is desirable that the solid catalysts to
over alumina as support. They suggested that this indicates that SiO2
be used in solid-catalyzed DMTM reactions carried out in the gas
was a more favorable support to enhance methanol selectivity than alu-
phase should be active under milder reaction conditions than those
mina at these temperatures [86].
used in the free-catalyst DMTM processes [82].
Compared to the OCM process, which is another solid-catalyzed pro-
cess for the direct conversion of methane in chemicals (i.e., ethane and 4.4.2. Zeolite catalysts
ethylene), the DMTM process has been relatively less explored. The mo- Currently, studying the support effect, as well as the role of Cu cata-
tive for examining DMTM is that, unlike the OCM reactions that are fea- lyst on DMTM reactions, has assumed great significance due to the dis-
sible only in the gas phase, DMTM route via solid-catalyzed process covery of a selective route to oxidize methane to methanol. Groothaert

Table 4
Methane partial oxidation with nitrous oxide over Cu–Fe-ZSM-5 catalystsa.
Adapted from Ref. [97].

Exp. CH4:N2O molar ratio Space velocity × 10−4 (h−1) Reaction temperature (K) CH4 conversion (%) Product selectivityb (%C)

CO2 CO CH3OH HCHO Olef.

1 80:20 22.0 615 1.52 51 7.2 38 3.8 0.6


2 80:20 43.1 615 1.12 35 8.5 50 6.2 0.8
3 80:20 43.1 510 0.25 19 2.5 78 1.0 –
4 60:40 42.2 619 n.a. 87 9.7 1.3 2.4 –
5 18:82 42.2 619 85 100 – – – –
a
Data taken after 60 min on stream.
b
Olef. = mixture of C2 and C3 olefins; n.a. = not available.
52 M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61

Table 5 steps, as follows. First, Cu–ZSM-5 zeolite should be activated, which


Methane partial oxidation with oxygen over various Fe–NaZSM-5 catalystsa. can be performed by flowing the adequate oxidant (i.e., O2 or NO),
Adapted from Ref. [98].
and then generating the active sites. The second step comprises meth-
Exp. Si:Fe molar ratio CH4 conversion (%) Product selectivity (%) ane oxidation over zeolite active sites. Finally, the third step is the ex-
CH3OH HCHO CO2 traction of methanol formed delivering the active sites to a new
b activation step. Unfortunately, this third step has been impractical and
1 45 0.06 74.37 15.65 15.20
2 34 0.07 69.13 14.39 16.48 hampered in turning the whole process into a catalytic cycle [93,94].
3 22 0.10 63.47 14.84 21.68 Recently, a breakthrough in the chemistry of copper zeolites was de-
a
Reaction conditions: temperature (623 K); contact time (0.5 s); oxygen (15.5 vol.%).
scribed; the ability of another Cu zeolite (i.e., named as Cu–MOR) to
b
Temperature (663 K). convert methane to methanol at 473 K through a multistep process
[95]. In that process, the Cu–MOR catalyst is pre-oxidized, and then
after further hydration and methanol recovery, the catalyst is then
reactivated at the reaction end. These results suggest that this zeolite
Table 6 is also able to produce the tricopper oxide active sites [96]. Although
Methane partial oxidation with oxygen over various Fe–HZSM-5 catalystsa. opening a new possibility to oxidize methane to methanol through
Adapted from Ref. [98].
straight reactions, this work did not achieve a complete catalytic cycle.
Exp. Si:Fe molar ratio CH4 conversion (%) Product selectivity (%) The success of copper zeolites on methane oxidation reactions to
CH3OH HCHO CO2 methanol has motivated assessment of the activity of other zeolite cat-
alysts. Indeed, even before the first data related methane oxidation to
1 45 11.22 16.51 25.48 58.01
2 34 19.11 14.17 18.79 67.03 methanol in the presence of Cu–ZSM-5 at room temperature, Anderson
3 22 31.51 10.79 17.06 72.14 and Tsai and Anderson described methane oxidation by nitrous oxide
a
Reaction conditions: temperature (903 K); contact time (2.5 s); oxygen (15.5 vol.%).
over Fe-ZSM-5 catalysts, synthesized by the exchange of Al3+ cations
in the lattice of the Cu–ZSM-5 zeolite by Fe3+ cations [97]. Those au-
thors devoted most of their attention to assessing the activity of Cu–
et al. were the first to report the selective oxidation of methane over Cu– FeZSM-5 on methane oxidation by nitrous oxide at a temperature
ZSM-5 catalyst (i.e., Zeolite Socony Mobil-5, an aluminosilicate zeolite range of 500 to 620 K. Table 4 summarizes the main results.
produced by Mobil Oil Company) [87–89]. They discovered that after Over H–FeZSM-5, the carbon oxides were the major products (not
activating the dried Cu–ZSM-5 zeolite with oxygen or nitrous oxide showed herein). Moreover, over Cu–FeZSM-5, the reaction was highly
(ca. 448 K or room temperature, respectively), and flow methane over selective to methanol, reaching nearly 78% under favorable conditions
the catalyst, methanol was then selectively formed. (entry 3, Table 4). Those authors attributed the exceptional behavior
Even though the reaction is not a truly catalytic process (i.e., it is a of Cu–Fe-ZSM-5 zeolite to probable synergism between the Cu2+ and
stoichiometric reaction, the catalyst does not turn over), it has attracted Fe3+ cations [97].
great attention because it was highly selective when working at a mild Michalkiewicz performed the DMTM reactions over Fe-ZSM-5 cata-
temperature range (ca. 373 to 473 K), a possible avenue to reach the lysts, under room pressure and at temperatures between 623 and 923 K
previously mentioned “Holy Grail” [52]. In fact, recent studies demon- [98]. He assessed the effects of acidic hydrogen replacement by sodium
strated that methanol formation was directly linked to the in situ forma- and variations of the Si/Fe molar ratio. The best result in terms of selec-
tion of tricopper species on the active sites on the cavity of ZSM-5 tivity was achieved in the Fe–NaZSM-5-catalyzed reactions, which re-
zeolite. In situ ultraviolet–visible spectroscopy studies allowed monitor- sulted in 74% methanol selectivity when the Si:Fe molar proportion
ing the methane oxidation progress over pre-activated Cu–ZSM-5 zeo- was 45:1 (Table 5). However, these reactions achieved only poor con-
lite (i.e., with O2 or N2O) [90]. Throughout the reaction, a decrease of versions (ca. 0.06 to 0.10%).
the band intensity at a wave number close to 22,700 cm− 1 was un- Conversely, high conversions were reached in the presence of cata-
equivocally attributed to the copper (II) μ-dioxide active site, which si- lyst Fe-HZSM-5 (ca. 31.51% for Si/Fe ratio = 22:1) (Table 6). Neverthe-
multaneously occurred with an increase in methanol concentration. less, it resulted in poor methanol selectivity for methanol. Those authors
Nevertheless, to carry this process over Cu–ZSM-5 zeolite displays concluded that increases of temperature and contact time favored the
great challenges. First, removing the methanol from the zeolite requires formation over oxidation products, as suggested by the increase of se-
an additional step of solvent extraction, lowering the methanol yield. lectivity to HCHO and CO2.
Moreover, the reaction is not genuinely catalytic; there is a stoichiomet- Panov et al. assessed the methane partial oxidation by nitrous oxide
ric characteristic of the reaction between the methane and the copper of in the processes in which the α-oxygen sites were created at room tem-
the active sites of the zeolite activate cavity [90]. perature or 433 K temperature over FeZSM-5 zeolite [99,100]. In the
Different approaches such as try convert methane to methanol over first study, they increased the amount of α-sites to 100 μmol/g, and in-
Cu–ZSM-5 zeolites even under non-catalytic conditions have consid- vestigated the methane oxidation by α-oxygen pre-deposited from ni-
ered. However, as far we know, they have failed [91]. In fact, to reacti- trous oxide at room temperature. They proposed that the following
vate the Cu+ sites on the zeolite cavity, the methanol must first be reactions occur over Fe–HZSM-5 or Fe–Na-ZSM-5 catalyst (Scheme 8)
removed [92]. Thus, a possible catalytic cycle should consist of three [99]:

Scheme 8. Methane partial oxidation by oxygen over Fe–HZSM-5 or Fe–Na-ZSM-5 catalysts.


Adapted from Ref. [99].
M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61 53

Table 7 Table 8
Product extraction from the Fe-ZSM-5 surface at different TON of the quasicatalytic reac- Methane partial oxidation with oxygen over various Co–HZSM-5 catalystsa.
tion CH4 + N2O + α-sites at 433 K temperature. Adapted from Ref. [103].
Adapted from Ref. [97].
Exp. Samples Cobalt speciation
Sample TON Amount of Amount of product Yield of CH3OH
% Cox+ in Co3O4 % Cox+ in CoO % isolated Co2+
no. CH4 reacted extracted (μmol g−1) product fraction
(μmol g−1) extracted (mol%) 1 Co–ZSM-5-p 25 45 30
CH3OH CH3OCH3 CH3CHO
(mol%) 2 Co–ZSM-5-acid 15 30 55
3 Co–ZSM-5-at-1–20 35 55 10
1 0.6 60 34 1.6 0.3 63 95
4 Co–ZSM-5-at-1–20 acid 40 45 25
2 1.0 100 52 4.0 1.3 63 91
a
3 1.9 190 96 14 3.0 68 85 Reaction conditions: metal catalysts prepared by impregnation. CH4/ O2 / He molar ratio
4 2.7 265 129 25 3.5 70 82 = 23/ 3/ 5; pressure = 15 bar; GHSV = 5000 h-1; metal catalysts prepared by
5 3.6 370 160 49 0.3 70 76 impregnation.

As can be seen, the Fe+2 cations of α-sites are activated by nitrous


They discovered that the reactions occur through a hydrogen ab- oxide generating adsorbed oxidant species (i.e., Fe3 +–O•−)α-sites,
straction mechanism, making methoxy or hydroxy groups bounded to which convert methane to adsorbed methanol. Afterwards, adsorbed
the α-sites. When the same reaction is performed with heating to methanol can be converted to dimethyl ether and water.
433 K, they verified that at CH4:N2O molar ratio equal to 1:1, the reac- Recently, Panov et al. assessed the methane oxidation by N2O at
tions directly provide methanol (Table 7) [100]. 473 K again and showed that the reaction occurred in a quasicatalytic
Spectroscopic techniques such as FT-IR (i.e., Fourier Transform- mode, resulting in product accumulation on the Fe-ZSM-5 surface
Infrared Spectroscopy), 13C CP/MAS NMR (i.e., Solid-state Cross- [102]. In this work, after 4 h reaction the catalyst achieved a turnover
Polarization Magic Angle Spinning Carbon-13 Nuclear Magnetic Reso- number close to 7. In accordance with these authors, methanol was
nance spectroscopy) and an extraction method were used to quantify formed throughout methane oxidation by oxygen(α) over (Fe3+–O)α
product formation on the zeolite surface. Remarkably, the spillover of sites at 473 K. They verified that methanol and dimethyl ether were
methanol from α-sites allowed these sites to be reused in new steps the main products extracted from the catalyst surface. Those authors
of oxygen deposition, making the reaction a “quasicatalytic” process discovered an outstanding result in which water noticeably increased
(i.e., the turnover number was slightly larger than one). Notably, a max- methanol selectivity (ca. 62% at 548 K; including coke). Because DME
imum TON (i.e., turnover number) equal to 3.6 was reached after a 19 h (i.e. dimethyl ether) was absent in the gas phase, they supposed that it
reaction (Table 7). Those authors verified that a fraction of methanol was converted to coke. That work was the first to describe quasicatalytic
underwent dehydration on zeolite surface, yielding mainly dimethyl and catalytic reactions over a broad temperature range [102].
ether, in addition to other minority products that were not extractable Beznis et al. assessed the activity of Co–ZSM-5 solid catalysts on
[100]. reactions of partial oxidation of methane. They found that methanol
Wood et al. studied the mechanism of “catalytic” oxidation reactions production proportionally increased in relation to surface area of
of methane by nitrous oxide over Fe-ZSM-5 zeolite, and concluded that catalyst, which can be increased treating the catalyst with NaOH
the primary products of methane oxidation are methoxy groups bound- [103]. Those authors discovered that the active sites (i.e., cobalt oxidic
ed to active iron (i.e., Fe–OCH3) [101]. IR spectroscopy data show that species, such as Co3O4 and CoO, present on catalyst external surface)
these methoxy groups are formed at 448 K and reach a maximum con- were also proportionally formed in relation to surface area of solid
centration of 548 K; whereas, the release of reaction products begins at catalyst.
523 K. In the same work, the authors concluded that the presence of extra-
In accordance with those authors, the hydrolysis reaction by the framework Al+3 cations blocked the zeolite pores, hampering the forma-
water generated after over oxidation of methane may then convert tion of the Co+2 active species, which are crucial for reaction selectivity
these species to methanol. However, poor selectivity (ca. 2%) was ob- [103].
tained due to methane combustion, in addition to the strong interaction The acidic treatment removed the Al+3 cations, triggering the speci-
of methanol and the zeolite, which leads to the formation Si–OCH3 spe- ation of cobalt oxides, thus affecting their distribution on the catalyst
cies [98]. surface. Table 8 presents the main results, which were obtained via
Panov et al. investigated the “quasicatalytic” reactions of methane H2-TPR gas desorption analysis.
partial oxidation to methanol by nitrous oxide, using IR and 13C CP/ The highest concentration of isolated Co+2 cations was detected in
MAS NMR spectroscopies [102]. They analyzed the zeolite after the reac- the Co–ZSM-5 zeolite treated with HNO3 during 20 h (Exp. 3, Table 8),
tions and identified adsorbed species on the catalyst surface. Those au- indicating that a high aluminum removal is implied in a high presence
thors proposed a possible sequence of equations that describes these of these cations. Furthermore, if carried out after an alkaline treatment,
“quasicatalytic” reactions (i.e., reaction of CH4 + N2O + α sites over the acidic treatment did not eliminate Al+3 cations, and therefore the
Fe-ZSM-5 catalyst), based on kinetic dependences data measured at isolated Co+2 cation concentration is ca. 25% (Exp. 4, Table 8), remain-
433 K (Scheme 9) [102]. ing close to that of non-treated samples (ca. 30%, Exp. 1, Table 8).

Scheme 9. Equations of reaction CH4 + N2O + α sites over Fe-ZSM-5 catalyst.


Adapted from Ref. [102].
54 M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61

lower than 473 K in less than 0.03 s, the methanol selectivity may be
significantly enhanced at low methane conversions.
Stroud et al. reported that when used CuO–MoO3 catalyst at a pres-
sure of 20 bar and a temperature of 758 K, they obtained a yield of
490 g (kg cat)−1 h−1 of oxygenated products. However, in addition to
methanol and formaldehyde, the reaction produced ethanol and acetal-
dehyde because the feed gas contained 6% ethane [106].
Nitrous oxide was also an oxidant assessed in methane oxidation re-
actions to methanol over MoO3/SiO2 catalyst. The best result was
reached in reactions over 1.7% MoO3/SiO2 catalyst, at 1 bar pressure
and 833 K temperature. The reactions under these conditions formed
methanol and formaldehyde with total selectivity equal to 84.6%
[107]. However, the same researchers could not reproduce their own re-
sults in their next work [108]. The methanol selectivity obtained by
them was lower; however, they discovered that an addition of steam
to the feed gas increased methanol selectivity to ca. 78.1%, an effect
that they attributed to the inhibition of carbon oxides.
Fig. 7. Industrial processes installed to synthesize methanol. The partial oxidation of methane to methanol over molybdenum or
Adapted from Johnson Matthey.
tungsten trioxide based catalysts containing different metal oxides as
dopants was investigated, aiming to identify a catalytic composition
Those authors evaluated the catalytic performance of these zeolites, and that would be able to activate methane without resulting in a loss of
the main results are displayed in Fig. 11 [103]. methanol selectivity [109]. Herein, we only display the results of molyb-
The zeolite alkaline treatment provoked an improvement in metha- denum catalysts (Table 9).
nol selectivity (Fig. 11). In contrast, an increase in zeolite acidic changes The Cu/MoO3 catalyst was more selective than the free catalyst gas
the selectivity toward the formaldehyde formation, signifying that the phase oxidation performed in a reactor bed packed with quartz glass
Co+2 isolated ions give preferentiality to this reaction. beads, similar to that observed for the Ga2O3/MoO3 catalyst (Exps. 5
The zeolite acid treatment minimizes the effect of Al+3 cations; and 8, Table 9) [109]. Those authors attributed the high methanol
however, it also modifies the amount of H+ or Na+ cations present in yield to the possible synergism between MoO3 and Ga2O3 oxides.
the zeolites. Consequently, the reaction selectivity is also changed An outstanding result was obtained by comparing the activity of all
(Fig. 12). these different molybdenum based catalysts with data obtained from
Remarkably, the cation nature (i.e., H+ or Na+ cations) was essential an empty reactor; none of the reactions performed in the presence of
to control the reaction selectivity. Fig. 7 reveals that the catalyst obtain- catalysts assessed by them reached the conversion and methanol selec-
ed after alkaline treatment was much more selective to methanol than tivity achieved in reactions made with an empty reactor (Table 9) [109].
other acid or non-treated zeolites. Those authors linked this observation We concluded that due to high temperatures employed, the reactions in
to the fact that after the first alkaline treatment, the Na+–Co–ZSM-5 cat- the gas phase were preferably more accurate rather than those on the
alysts contain more cobalt on the external surface than H+–Co–ZSM-5 catalyst surface.
catalyst, which favors the methanol formation [104]. (See Fig. 8.) Zhang et al. studied the activity of MoO3 or ZrO2/La–Co–O supported
catalysts in the methane partial oxidation reactions to methanol using
4.4.3. Molybdenum catalysts oxygen as an oxidant, under high pressures (ca. 4.2 to 5.0 MPa) and
For the past 70 years, the methane oxidation reactions to methanol temperatures ranging from 653 to 693 K [110]. They designed a
have usually carried out over molybdenum catalysts under high oxygen continuous-flow reactor capable of operating at high pressures, in
pressure. Dowden and Walker investigate the activity of a number which the methane oxidation to methanol by dioxygen over MoOx/
of supported and unsupported molybdenum catalysts, at pressures La–Co–O and MoOx/ZrO2 catalysts was performed (Table 10). We here-
of 50 bar, temperatures ranging from 403 to 573 K, and feed gas in highlight only results in which methanol selectivity was relevant
with 97% methane and 3% oxygen. The most active catalyst was (Table 10) [110].
Fe 2 O 3 (MoO 3 ) 3 , which resulted in 869 g methanol/kg cat h [105]. Under the studied conditions, ZrO2 and MoOx/ZrO2 gave only trace
However, in accordance with these authors, if the desired products amounts of methanol and for such reasons, omitted in Table 10. The
were removed from the catalyst surface and cooled to temperature MoOx/La–Co–O catalyst containing 7 wt.% showed the highest

Fig. 8. Formation of the transition-state complex during facile singlet oxene transfer to methane from a dioxygen-activated tricopper complex.
Adapted from Ref. [66].
M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61 55

Fig. 9. Productive cycling in the oxidation of methane by O2, mediated by the [Cu1+–Cu1+–Cu1+(7-N-Etppz)]1+ complex in the presence of H2O2 as the sacrificial reductant.
Adapted from Refs. [64–66].

selectivity and yield to methanol (ca. 60% and 6%, respectively), in reac- The (O−)/(O− 2 ) ratio was linked by those authors to the higher defi-
tions performed at 4.2 bar and 693 K [110]. ciency of lattice sites commonly described for LaCoO3 and Co3O4 species
Those authors evaluated the reducibility of the supported catalyst compared with other catalysts. These phases are present on MoOx/La–
and verified that 7 wt.% MoOx/La–Co–O catalyst showed higher (O−)/ Co–O catalysts and were identified by XRD spectroscopy analyses
(O−2 ) proportion than 12 wt.% MoOx/ZrO2 ones. They obtained such [110]. Such a deficiency contributes to the migration and transformation
data by measurement of hydrogen uptake of each catalyst. They suggest of O2 from the gas phase to give O− 2 and O

species on the catalyst
that this feature can be crucial to obtain high methanol selectivity surface.
(Table 10).
The activity, yield, and hydrogen consumption data displayed in 4.4.4. Iron sodalite catalysts
Table 11 were calculated considering the catalyst surface area. These re- For the past 90 years, numerous researchers have investigated the
sults support the explanation that the reducibility, (O−)/(O−2 ) ratio, as activity of iron sodalite based catalysts in methane oxidation reactions
well as the presence of MoOx are essential conditions for good methanol to methanol [111]. Lyions et al. carried out reactions using a feed com-
selectivity. position equal to 3:1 of methane to air, under 55 bar pressure at

Fig. 10. Abortive cycling in the oxidation of methane by O2, mediated by the Cu1+–Cu1+–Cu1+(7-N-Etppz)1+ complex in the presence of H2O2 as the sacrificial reductant.
Adapted from Refs. [64–66]].
56 M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61

689 K. They achieved 70% methanol selectivity in reactions with 5.7%


methane conversion [112]. This group designed their iron catalysts
based on biomimetic studies of the mechanism used by cytochrome
P450 and methane monoxygenase, which should contain as active
sites ferryl (i.e., Fe+4 = O), as well as possible active sites in reactions
of DMTM [113,114]. Lyions et al. attributed the high activity of iron so-
dalite catalysts to a possible interaction between iron species placed on
framework and extra framework [112]. However, even though iron so-
dalite catalysts were successful, Hutchings et al. performed experimen-
tal and theoretical studies, and showed that catalyst used in a simple
flow reactor was less effective that in an empty reactor [109,113].
Chun and Anthony also assessed the participation of the reactor's
wall in the reaction of methane oxidation to methanol [115]. They com-
pared their results to ones obtained with different metal oxide catalysts,
under the same reaction conditions (i.e., high pressure and tempera-
ture). Those authors synthesized a number of catalysts employing dif-
ferent methods (i.e., sol gel method, ionic exchange, precipitation),
Fig. 11. Selectivity toward methanol and formaldehyde, and the total amount of produced and tested them on methane oxidation reactions by oxygen under
oxygenates of Co–ZSM-5 zeolites for methane activation with O2 at 423 K: (A) Co–ZSM-5- high pressure (i.e., 50 atm; 95.65% CH4 and 4.35% O2 feed composition)
p; (B) Co–ZSM-5-at-1–20; (C) Co–ZSM-5-at-1–20-acid; and (D) Co–ZSM-5-acid. and temperatures varying from 640 to 750 K. Fig. 13 shows the main re-
Adapted from Ref. [103].
sults as well as the catalyst synthesis conditions.
Although the highest methanol selectivity was achieved using a
Pyrex bed (ca. 38%), in general all the metal oxides exhibited methanol
selectivity close to ca. 30%. In addition, conversions of methane and ox-
ygen were almost constant (ca. 95%), regardless of the catalyst used.
Those authors verified that under conditions of high pressure and tem-
perature, the homogeneous reactions were predominant, occurring
probably in the empty space present on packed bed of reactor, being re-
sponsible for a large fraction of products formed (Fig. 13) [115].
Chun and Anthony also performed several reactions using an empty
Pyrex glass reactor or packed reactor with Pyrex beads at a range of
temperatures from 704 to 757 K. Those authors verified that regardless
of the conditions used, the methanol selectivity ranged from 37 to 42%
[115].
They concluded that the Pyrex glass surface could inhibit the free
radicals formation, which promotes the homogeneous oxidation of
methane. Nevertheless, it is worth noting that this effect was insignifi-
cant when the reactions were performed using a low surface:volume
ratio. When the reactor bed was packed with larger Pyrex beads (ca.
1.41 to 2.00 mm), at high surface to volume ratio (ca. 300 cm−1) and
temperatures of 724 to 730 K, the conversion of methane was drastically
Fig. 12. Selectivity toward methanol and formaldehyde, and the total amount of produced
oxygenates of Co–ZSM-5 zeolites for methane activation with O2 at 423 K: (A) Co–ZSM-5-
reduced from 4.0 to 0.7% (Exp. 4, Table 12). Conversely, at temperatures
p; (B) Co–ZSM-5-at-1–20; (C) Co–H-ZSM-5; and (D) Co–H-ZSM-5-at-1–20. of 704 to 726 K, and a lower ratio of surface to volume with smaller
Adapted from Ref. [103]. Pyrex beads (0.125/0.149 mm), methane maximum conversion was

Fig. 13. (a) Comparison of selectivity for heterogeneous reactions (O2 = 4.35%; pressure = 50 atm; oxygen conversion = 95–100%; residence time = 4.5 s); (b) properties of catalysts.
Adapted from Ref. [115].
M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61 57

Fig. 16. Effect of NO concentration on the conversion of methane and O2, and on the yield
of C1-oxygenates in reactions over V2O5/SiO2 catalyst (reaction conditions: catalyst weight
Fig. 14. Effect of Fe/Fe + Cu molar ratio in the oxygenates' yield.
(V2O5/SiO2 catalyst, 0.22 g); total flow (89.8 mL min−1); feed = O2 (29%); CH4 (41.5%);
Adapted from Ref. [116].
NO (0 to 2.9% in N2)).
Adapted from Ref. [118].
reached, limited by total consumption of oxygen (i.e., 4.0 and 100% of
conversions for CH4 and O2, respectively). This inhibitor effect is a con- Conversely, in these reactions, the selectivity was very distinct;
sequence of the free volume reduction of the reactor that hampers the whereas, the former was more selective for methanol, the latter was
gas phase reactions [115]. more selective for formaldehyde (Exps. 6 and 11, Table 13). Moreover,
those authors concluded that oxygenates' yield depends on the molar
ratio of Fe/Fe + Cu in pyrophosphates (Fig. 14) [116].
4.4.5. Other solid catalysts In the same work, the oxidant effect was investigated at different
Stolcova et al. described the methane oxidation to methanol by ni- temperatures [116]. In the presence of N2O at 673 K, methanol and
trous oxide over Cu–Fe pyrophosphates at high temperatures (ca. formaldehyde were formed at an equal molar ratio, however, with
903 K) and at low pressures (ca. 0.65 bar and 0.26 bar for methane low methane conversion (ca. 4%).
and nitrous oxide, respectively) [116]. Table 13 summarizes the main Fig. 15 shows that increasing the reaction temperature result in an
results obtained on assessing the activity of the copper iron pyrophos- increase of the methane conversion, reaching a maximum of ca. 6%.
phates. We omitted results with methanol selectivity lower than 1.9 However, a concomitant decreasing of the methanol selectivity is also
for simplification. favored, and an increase of selectivity for formaldehyde and carbon ox-
Starting from (Cu+2)(Fe+3)2(P2O−4 7 )2 catalyst, they replaced part of ides simultaneously occurs. Using similar reaction conditions, Stolcova
Fe+3 cations by Cu+2, and otherwise replaced part of Cu+2 cations by et al. replaced the nitrous oxide by oxygen and verified that under
Fe+3, keeping the neutrality of catalysts constant throughout an ade-
quate adjustment of pyrophosphate anions load. The monometallic Cu
catalyst (i.e., Exp. 4, CuFe-2/0–1023 K catalyst, Table 13) was more ac-
tive than iron ones (i.e., Exp. 9, CuFe-0/3–1023 K catalyst, Table 13).

Fig. 15. Effect of temperature on conversion and selectivity of methane partial oxidation to
methanol using N2O over CuFe-1/2–1023 K–4 h catalyst (reaction conditions: pressure = Fig. 17. Effect of temperature and of CH4/O2 ratio on the CH4 conversion in reactions over
CH4 (0.65 bar), N2O (0.26 bar); temperature range (673–923 K), catalyst weight (0.5 g), V2O5/SiO2 catalyst (reaction conditions: catalyst weight (0.30 g); total flow
flow (3.6 L·h−1)). (122 mL min−1); feed (22.9–34.4% CH4, 1.0% NO in N2)).
Adapted from Ref. [116]. Adapted from Ref. [118].
58 M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61

Table 10
Methane partial oxidation over MoO3/La–Co–O catalystsa.
Adapted from Ref. [110].

Exp. Catalyst Conversion CH4 Selectivity (%) Yield


(%)
HCHO CH3OH CO CO2 CH3OH

1 Quartz wool 0.9 – 4.9 43.1 51.8 0.04


2 MoO3 1.6 6.9 Trace 40.8 52.2 Trace
3 La–Co–O 4.2 – – 8.2 91.6 –
4 3 wt.% 8.9 Trace 28.8 11.2 60.0 2.6
MoO3/La–Co–O
5 7 wt.% 11.2 Trace 60.0 21.8 18.2 6.7
MoO3/La–Co–O
6 10 wt.% 11.6 Trace 33.2 31.7 34.8 3.9
MoO3/La–Co–O
a
Reaction conditions: space velocity (14,400 mL/(g catal) h−1); temperature (420 K);
pressure (4.2 bar).

selectivity (Fig. 16) [118]. Those authors showed that although insignif-
icant in the absence of nitric oxide, the free-catalyst reaction can occur if
it is nitric oxide to the reaction, probably because it changes the propor-
tion between methyl and methyl peroxide [118]. For this reason, even
though it has a low surface area, the supported catalyst V2O5/SiO2
Fig. 18. Effect of temperature and of CH4/O2 ratio on the yield of C1-oxygenates in reactions achieved a yield close to 16% when NO was introduced in the feed
over V2O5/SiO2 catalyst (reaction conditions: catalyst weight (0.30 g); total flow stream.
(122 mL min−1); feed (22.9–34.4% CH4, 1.0% NO in N2)). Using NO (ca. 1% NO in N2 feed) and oxygen as an oxidant, they
Adapted from Ref. [118]. assessed the effects of temperature (ca. 823 to 923 K) and methane to
oxygen molar ratio (ca. 1.3 to 2.0%) in the methane oxidation over
V2O5/SiO2 catalyst (Fig. 17) [118].
oxygen and at temperatures ranging of 673 to 923 K, methanol selectiv- In general, an increase of CH4:O2 proportion resulted in a decreasing in
ity begins at lower values than ca. 10%, and drops down to almost zero, CH4 conversion regardless of the reaction temperature, except at 823 K, in
when the conversions reached 30% [116]. They also assessed the possi- which the reaction almost stopped. The highest conversion (ca. 38%) was
ble participation of the lattice oxygen atoms via experiments with reac- obtained at 898 and 923 K (Fig. 17), using CH4:O2 at a 1:3 molar ratio.
tants pulse. Stolcova et al. concluded that after react with methane, the Conversely, the reactions performed at the highest temperature (ca.
oxygen atoms provided by nitrous oxide could quickly replace those 923 K) and with CH4:O2 molar ratio equal to 1:8 provided greatest com-
atoms. It proceeds without compromising the conversion, selectivity, bined yield (i.e., methanol and formaldehyde yields) (Fig. 18).
and active sites during the entire process. Those authors attributed the Heteropolyacids are a versatile class of solid catalysts active
highest activity of CuFe-1/2–1023 K–4 h to a possible synergism be- in different reactions in either homogenous or heterogeneous
tween copper and iron cations on these pyrophosphates catalysts. systems [119]. Kasztelan and Moffat investigated the activity of differ-
In this regard, Sojka et al. investigated a possible synergic effect be- ent silica-supported heteropolyacids (i.e., dodecatungstophosphoric,
tween Cu and Fe incorporated to the ZnO support [117]. They per- dodecamolybdophosphoric, decamolybdo-2-vanadophosphoric, dodeca-
formed the methane oxidation reactions to methanol using air as an tungstosilicic, and dodecamolybdosilicic acids), in the methane oxidation
oxidant at a temperature range of 773 to 1123 K. Those authors sug- reactions by oxygen or nitrous oxide [120]. However, they reached poor
gested that Fe(II)/Fe(III) and Cu(I)/(II) redox coupling, in addition to methanol selectivity (ca. lower than 1.0%) in the reaction performed in
the Lewis acidity of these catalysts, may promote the oxidation of meth- a fixed-bed continuous-flow reactor heated to 843 K.
yl intermediates to methoxyde, which could be oxidized via hydride Mansouri et al. described the use of lacunary Keggin-type
transfer reactions catalyzed by ZnO supported iron copper [117]. heteropolyoxometalates in the methane oxidation to methanol using
Fierro et al. evaluated the activity of supported silica vanadium ox- nitrous oxide (Table 14) or oxygen (Table 15) [121]. They synthesized
ides in the methane oxidation by nitric oxide and oxygen mixtures salts with general formula [PW11MO39](7 − n)−, where M = Co(II),
under room pressure, and verified that an increase of NO concentration Ni(II), and Fe(III) from the K7PW11O39 precursor, and used it as catalysts
to 1.5% sharply improved the methanol and formaldehyde combined in methane oxidation at room pressure and temperatures of 873 to
923 K [121].
When oxidized by nitrous oxide, an increase in the temperature
Table 9
Methane partial oxidation with oxygen over various molybdenum catalystsa. resulted in an increase in methane conversion, regardless of the cat-
Adapted from Ref. [109]. alyst employed. The CsPW 11Fe-catalyzed reactions (ca. 33.6%) at
gave highest methanol selectivity besides 5% methane conversion
Exp. Catalyst Temperature (K) CH4 conversion Main products
selectivity (%)
(Table 14). Although this conversion achieved the highest value
(ca. 11%) when the temperature was raised to 923 K, the methanol
CH3OH CO CO2
selectivity was remarkably reduced to 14.6%. The same effect has ob-
1 MoO3 450 0.3 11 72 17 served for the other catalysts.
2 Cu/MoO3 450 0.6 19 60 7
On the other hand, the lacunary heteropoly were catalysts that were
3 Co/MoO3 450 0.5 14 47 39
4 Fe/MoO3 450 0.5 14 67 19 less effective in the reactions of methane oxidation by oxygen, resulting
5 Ga/MoO3 450 1.3 10 77 13 in lower conversions than the ones reached with nitrous oxide
6 Zn/MoO3 450 (Tables 14 and 15).The methanol selectivity obtained when replacing
7 Empty tube 450 3.5 29 63 7 nitrous oxide by oxygen was similar; the CsPW11Fe-catalyzed reactions
8 Quartz packing 500 8.0 15 68 17
achieved the highest selectivity (ca. 17.8%). However, the reactions
M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61 59

Table 11
Relationship between catalytic performance and activity.
Adapted from Ref. [110].

Exp. Catalyst Modified activity Modified yield Modified H2 O−/O2− ratio


(×107 mol s−1 m2) (×107 mol s−1 m2) consumption
mmol/m−2
C1-oxygenate COx

1 La–Co–O 8.0 0 18.0 2.6 3.3


2 7 wt.% MoOx/La–Co–O 17.8 10.6 (CH3OH) 7.2 2.7 2.0
3 12 wt.% MoOx/ZrO2 3.6 1.7 (HCHO) 1.9 0.1 0

Table 12
Inhibiting effect of Pyrex for homogeneous reactionsa.
Adapted from Ref. [115].

Exp. Range temperature CH4 O2 Selectivity (%)


of catalyst bed (K) Conversion Conversion
CO CO2 HCHO CH3OH
(%) (%)
b
1 704–724 4.0 100 49.2 6.2 1.8 39.7
2c 706–726 4.0 100 51.1 7 Trace 41.8
3c 708–734 4.3 100 50.4 6.2 8.9 34.5
4d 724–730 0.7 20.1 39 18.6 Trace 42.4
4e 754–777 4.0 94.3 56.6 2.9 4.2 34.6
a
Reaction conditions: 4% oxygen; pressure 50 bar.
b
Surface to volume ratio of Pyrex (L·cm−1) = 10.0; residence time = 5.9 s; empty Pyrex tube was used as a reactor.
c
Surface to volume ratio of Pyrex (L·cm−1) = 32.3; residence time = 5.8 s; 1.41/2.00 mm Pyrex beads were packed in a reactor.
d
Surface to volume ratio of Pyrex (L·cm−1) = 300; residence time = 4.7 s; 0.125/0.149 mm Pyrex beads were packed in a reactor.
e
Surface to volume ratio of Pyrex (L·cm−1) = 300; residence time = 4.5 s; 0.125/0.149 mm Pyrex beads were packed in a reactor.

heated to 873 K reached a poor methane conversion of 0.8%. The natural gas). These reactions are carried out over alumina supported
CsPW11Co-catalyzed reactions achieved the highest conversion of Ni catalysts at high temperatures. The methane conversion to syngas
methane (ca. 6%) (Table 15). and posterior synthesis of methanol occur in multistep processes,
By using of voltammetry analysis, Mansouri et al. attributed the which is a very costly and intensive energy route. Although drastic reac-
higher reactivity of nitrous oxide if compared to the oxygen to the tion conditions (ca. pressures of 50 to 100 bar, temperatures of 473 to
higher oxidant characteristic of metal cluster [121]. Based on kinetics 573 K), the methanol selectivity in this process is generally close to
data, those authors suggested different reaction pathways for methane 99%, with yields of 70 to 74%. The high cost of syngas production re-
oxidation by nitrous oxide or oxygen. The nitrous oxide activation quires a large capital investment, in addition to a complex industrial in-
may involve MO2 active sites, which provide the intermediates methoxy frastructure. These are the main disadvantages of this route. Among
to methanol or formaldehyde formation. Conversely, the oxygen activa- alternatives currently explored, the direct partial oxidation of methane
tion could involve a possible hydrogen abstraction step by the lattice ox- to methanol (DMTM route) has assumed significant importance and
ygen atom of MO groups, resulting in organometallic intermediates has become the focus of extensive research. In this review, we describe
(i.e., methyl-metal compounds), which play a crucial role throughout the background, as well as the recent advances obtained, in two DMTM
methane oxidation reactions [121]. routes that occur in the gas phase. The first is the catalyst-free homoge-
neous reactions, and the second is the solid-catalyzed process of meth-
5. Outlook ane oxidation to methanol. Catalyst-free homogeneous route converts
methane to methanol at high temperatures (ca. 700 to 745 K) and
Currently, all industrially consumed methanol has been produced high pressures (ca. 50 bar). In these processes, radical steps in the gas
through the syngas route, where Cu–Zn/Al2O3 is the catalyst phase govern the reaction, resulting in the over oxidation of methanol
(i.e., Synetix process). The more expansive step of this process is syngas to carbon oxides, compromising the reaction selectivity. Consequently,
production through the methane steam reforming (i.e. present in the maximum conversion attained in the reaction equilibrium is

Table 13
Catalytic performance of the CuFe pyrophosphates after different pretreatment and metal loads on the methane oxidationa.
Adapted from Ref. [116].

Exp. Catalyst CuFe-X–Y–Zb CH4 conversion (%) Selectivity (%)

HCHO CH3OH CO CO2 Yield of oxygenate


products

1 CuFe-1/2–1073 K–72 h 3.1 28.5 2.1 56.3 13.1 0.9


2 CuFe-1/2–1073 K–30 h 4.4 26.7 1.9 63.3 8.1 1.3
3 CuFe-1/2–1023 K–30 h 4.5 33.8 2.5 55.2 8.5 1.6
4 CuFe-2/0–1023 K–4 h 4.5 26.4 1.9 50.3 21.4 1.3
5 CuFe-3.25/0.5–1023 K–4 h 4.8 25.1 2.6 53.8 18.5 1.3
6 CuFe-2.5/1–1023 K–4 h 5.1 23.1 2.5 61.3 13.1 1.3
7 CuFe-1.75/1.5-1023 K–4 h 5.4 23.3 2.3 61.9 12.5 1.4
8 CuFe-0.5/2.5–1023 K–4 h 1.5 71.8 9.1 19.1 0 1.2
9 CuFe-0/3–1023 K–4 h 1.3 71.5 10.1 18.4 0 1.1
a
Reaction conditions: pressures = methane (0.65 bar), nitrous oxide (0.26 bar); flow = 3.6 L·h−1; temperature 903 K; catalyst = 0.5 g.
b
CuFe X–Y–Z: X = Cu: Fe molar ratio; Y = calcination temperature; Z = calcination time.
60 M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61

Table 14 methanol (DMTM route) competitive with the multistep process cur-
Heteropolytungstate-catalyzed methane oxidation by N2Oa. rently used, which starts from the syngas as feedstock.
Adapted from Ref. [121].

Exp. Catalyst T (K) CH4 conversion Selectivity (%)


Acknowledgments
CH3OH HCHO CO CO2

1 CsPW11Fe 873 5 33.6 15.6 46.8 4 The authors are grateful to the Brazilian financial agencies CAPES,
923 11 14.6 24 40.4 21 CNPq, and FAPEMIG.
2 CsPW11Co 873 3 25.7 15 50.9 18.3
923 10 10 20.6 42.9 26.4
3 CsPW11Ni 873 2 11.2 4.1 60.7 24 References
923 8 5 8.1 50.8 36
[1] A.P.E. York, T. Xiao, M.L.H. Green, Top. Catal. 22 (2003) 345–358.
a
Reaction conditions: metal catalysts prepared by impregnation. CH4/ O2 / He molar ratio [2] B.G. Hashiguchi, C.H. Hovelmann, S.M. Bischof, K.S. Lokare, C.H. Leung, R.A. Periana,
= 23/ 3/ 5; pressure = 15 bar; GHSV = 5000 h−1; metal catalysts prepared by Methane to methanol conversion, Encyclopedia of Inorganic Chemistry, John Wiley
impregnation. & Sons, Ltd. 2011, pp. 1–41.
[3] T. Mokrani, M. Scurrell, Catal. Rev. 51 (2009) 1–145.
[4] J.H. Lunsford, Catal. Today 63 (2000) 165–174.
responsible by control the methanol selectivity. In general, on the ho- [5] E.F. Sousa-Aguiar, L.G. Apel, Mota. C., Catal. Today 101 (2005) 3–7.
[6] Durante, V. A.; Walker, D. W.; Gussow, S. M.; Lyons, J. E., U. S. Patent No. 4,918,249,
mogeneous reactions reaches only low conversions (ca. 5 to 10%) and to Sun Refining and Marketing Company, 1990.
methanol selectivity of 30 to 40%. A milder oxidant as NO can replace [7] V.I. Mahesh, P.N. Lawrence, P. Alex, L.K. Edwin, S.S. Mohindar, B.D. Dady, Top. Catal.
the oxygen; nonetheless, it is generally less efficient. The solid- 29 (2004) 197–200.
[8] T.V. Choudhary, R.C. Vasant, Angew. Chem. Int. Ed. 47 (2008) 1828–1847.
catalyzed methanol synthesis was the second DMTM route discussed [9] D. Dissanyake, M.P. Rosynek, K.C.C. Kharas, J.H. Lunsford, J. Catal. 132 (1991)
herein. These reactions produce methanol at lower temperatures than 117–132.
those used in the absence of catalyst. The first catalysts most extensively [10] S.S. Bharadwaj, L.D. Schmidt, Fuel Process. Technol. 42 (1995) 109–127.
[11] H.-J. Arpe, Industrial Organic Chemistry, fifth ed. Wiley-VCH, Weinheim, 2010.
assessed on these reactions were iron sodalite and molybdenum- [12] V.R. Choudhary, A.M. Rajput, B. Prabhakar, A.S. Mamman, Fuel 77 (1998)
supported metal compounds. The reactions carried in packed silica reac- 1803–1807.
tor in the presence and absence of the solid metal catalysts revealed that [13] E. Ruckenstein, Y.H. Hu, Appl. Catal. A Gen. 183 (1999) 85–92.
[14] Y.H. Hu, E. Ruchenstein, Ind. Eng. Chem. Res. 37 (1998) 2333–2341.
radical reactions had a significant contribution to the conversion and
[15] Y. Chu, S. Li, J. Lin, J. Gu, Y. Yang, Appl. Catal. A Gen. 134 (1996) 67–80.
methanol selectivity. On the other hand, different of homogeneous reac- [16] D.A.J.M. Ligthart, J.A.Z. Pieterse, E.J.M. Hensen, Appl. Catal. A Gen. 405 (2011)
tions, the using of N2O or NO to oxidize methane to methanol over solid 108–119.
[17] D.A.J.M. Ligthart, R.A. van Santen, E.J.M. Hensen, J. Catal. 280 (2011) 206–220.
catalysts such as V2O5/SiO2 triggered a significant improvement in the
[18] P.J. Dahl, T.S. Christensen, S. Winter-Madsen, S.M. King, Proven autothermal
selectivity and conversion of reactions. Cu–Fe pyrophosphates were ac- reforming technology for modern largescale methanol plants, Nitrogen + Syngas,
tive in the methane oxidation to methanol by N2O, however, the tem- International Proced Conference & Exhibition 2014, pp. 2–3 (Paris).
perature and relative amount of Cu and Fe were crucial aspects in [19] K. Aasberg-Petersen, J.H.B. Hansen, T.S. Christensen, I. Dybkjaer, P.S. Christensen,
C.S. Nielsen, S.E.L.W. Madsen, J.R. Rostrup-Nielsen, Appl. Catal. A Gen. 221 (2001)
conversion and methanol selectivity. Reactions carried over Cu–Fe zeo- 379–387.
lites also achieved high conversions and methanol selectivity; nonethe- [20] P.J. Dahl, T.S. Christensen, S. Winter-Madsen, S.M. King, Process burners for syngas
less, the methane oxidation was stoichiometric in relation to copper or production, Nitrogen + Syngas, International Conference & Exhibition, issue 320
2012, p. 38.
iron catalyst active sites. The greatest difficulty of these reactions is [21] K. Aasberg-Petersen, T.S. Christensen, I. Dybkjaer, J. Sehested, M. Østberg, R.M.
the step of methanol extraction from the catalyst surface. The high Coertzen, M.J. Keyser, A.P. Steynberg, Stud. Surf. Sci. Catal. 152 (2004) 258–405.
methanol polarity hinders its recovery from zeolite surface catalyst. [22] G. Natta, in: P.H. Emmet (Ed.) Catalysis, vol. III, Reinhold, New York, 1953
(Chap. 8).
The activity of another class of solid catalysts on DMTM reactions was [23] Q. Zhang, D. He, Q. Zhu, J. Nat. Gas Chem. 17 (2008) 24–28.
investigated; lacunar tungsten heteropolyacids containing iron [24] J.B. Hansen, P.E.H. Nielsen, Handbook of Heterogeneous Catalysis, first ed. Wiley-
displayed selectivity of 33% to methanol, using N2O as an oxidant and VCH, Weinheim, 1997 2920.
[25] G.D. Li, Q.Y. Liu, Z.Y. Liu, Z.C. Zhang, C.Y. Li, W.Z. Wu, Angew. Chem. Int. Ed. 122
in reactions at 823 K.
(2010) 8658–8661.
Although advances achieved, to turn the methanol in a flexible [26] J. Lehmann, Nature 447 (2007) 143–144.
source of chemicals more than a costly alternative fuel requires a large [27] P.J.A. Tijm, F.J. Waller, D.M. Brown, Appl. Catal. A Gen. 221 (2001) 275–282.
[28] C.D. Chang, A.J. Silvestri, J. Catal. 47 (1997) 249–259.
technological throughput. This revolution certainly will result in great
[29] C. Hammond, S. Conrad, I. Hermans, ChemSusChem 5 (2012) 1668–1686.
expansion of methanol consume market. These changes depend mainly [30] G.A. Olah, Angew. Chem. Int. Ed. 44 (2005) 2636–2639.
on the methanol production process, and consist of reducing syngas [31] G. Ertl, H. Knozinger, J. Weitkamc, Handbook of Heterogeneous Catalysis, Copyright
costs (i.e., multistep route), as well as improving the selectivity of single VCH Verlagsgesellschaft mbH, 1997.
[32] BASF, German Patents 415 686, 441 433,462 837, 1923.
step processes. Indeed, there are enormous challenges to be overcome [33] Davies, P.; Snowdon, F. F. U.S. Patent 3.326.956, ICI, Ltd, 1967.
in order to make the single-step process of methane oxidation to [34] Cornthwaite, D. U. K. Patent 1,296,212 ICI, Ltd, 1972.
[35] L.E.Y. Nonenneman, V. Ponec, Catal. Lett. 7 (1990) 213–218.
[36] M.S. Spencer, Top. Catal. 8 (1999) 259–266.
[37] G.C. Chinchen, P.J. Denny, D.G. Parker, G.D. Short, M.S. Spencer, K.C. Waugh, D.A.
Whan, Prepar. Am. Chem. Soc. Div. Fuel Chem. 29 (1984) 178–186.
Table 15
[38] A.Y. Rozovskii, Y.B. Kagan, G.I. Lin, E.V. Slivinskii, S.M. Loktev, L.G. Liberov, A.N.
Heteropolytungstate-catalyzed methane oxidation by oxygena. Bashkirov, Kinet. Katal. 17 (1976) 1314–1318.
Adapted from Ref. [121]. [39] R. Burch, S.E. Golunski, M.S. Spencer, J. Chem. Soc. Faraday Trans. 86 (1990)
2683–2691.
Exp. Catalyst T (K) CH4 conversion Selectivity (%)
[40] M. Muhler, E. Tornquist, L.P. Nielsen, B.S. Clausen, H. Topsse, Catal. Lett. 25 (1995)
CH3OH HCHO CO CO2 1–8.
[41] K. Klier, Adv. Catal. 31 (1982) 243–313.
1 CsPW11Fe 873 0.8 17.8 16.5 36.8 28.9 [42] K. Klier, V. Chatikavanij, R.G. Herman, G.W. Simmons, J. Catal. 74 (1982) 343–360.
923 4 7.3 23.1 23.5 46.1 [43] P. Khirsariya, R.K. Mewada, Procedia Eng. 51 (2013) 409–415.
2 CsPW11Co 873 1 14.8 16.3 46.6 22.3 [44] M.V. Twigga, M.S. Spencer, Top. Catal. 22 (2013) 191–203.
923 6 6.6 22.5 34.9 36 [45] T. Kobayashi, K. Nakagawa, K. Tabata, M. Haruta, J. Chem. Soc. Chem. Commun.
3 CsPW11Ni 873 3 2.2 3.8 9.8 84.2 (1994) 1609–1610.
923 7 – – 5.8 94.2 [46] A. Caballero, P.J. Perez, Tutorial, Chem. Soc. Ver. 42 (2013) 8809–8820.
[47] M.J. Gradassi, N.W. Green, Fuel Process. Technol. 42 (1995) 65–83.
a
Reaction conditions: The reactions were carried out in a continuous fixed bed reactor [48] Q. Zhang, D. He, Q. Zhu, J. Nat. Gas Chem. 12 (2003) 81–89.
under room pressure with a CH4/O2 molar ratio equal the 2.5/1 at a flow rate of 2 L·h−1 [49] M. He, Y. Sun, B. Han, Angew. Chem. Int. Ed. 52 (2013) 2–16.
during 8 h. [50] J. Haggin, Chem. Eng. News 70 (1992) 33–35.
M.J. da Silva / Fuel Processing Technology 145 (2016) 42–61 61

[51] K. Aasberg-Petersen, J.H.B. Hansen, T.S. Christensen, I. Dybkjaer, P.S. Christensen, [87] T. Takemoto, D. He, Y. Teng, K. Tabata, E. Suzuki, Appl. Catal. A Gen. 225 (2002)
C.S. Nielsen, S. Madsen, J.R. Rostrup-Nielsen, Appl. Catal. A Gen. 221 (2001) 177–184.
379–387. [88] M.H. Groothaert, P.J. Smeets, B.F. Sels, P.A. Jacobs, R.A. Schoonheydt, J. Am. Chem.
[52] H. Schwarz, Angew. Chem. Int. Ed. 50 (2011) 10096–10115. Soc. 127 (2005) 1394–1395.
[53] A. Riaz, G. Zahedi, J.J. Klemes, J. Clean. Prod. 57 (2013) 30–47. [89] P.J. Smeets, M.H. Groothaert, R.A. Schoonheydt, Catal. Today 110 (2005) 303–309.
[54] R.C. Baliban, J.A. Elia, C.A. Floudas, AIChE J. 259 (2013) 505–531. [90] P. Vanelderen, J. Vancauwenbergh, B.F. Sels, R.A. Schoonheydt, Coord. Chem. Rev.
[55] J.H. Edwards, N,.R. Foster, Fuel Sci. Technol. Int. 4 (1986) 365–390. 257 (2013) 483–494.
[56] S. Friedle, E. Reisner, S.J. Lippard, Chem. Soc. Rev. 39 (2010) 2768–2779. [91] J.S. Woertink, P.J. Smeets, M.H. Groothaert, M.A. Vance, B.F. Sels, R.A. Schoonheydt,
[57] Chrstine E. Tinberg, Stephen J. Lippard, Acc. Chem. Res. 44 (2011) 280–288. E.I. Solomon, Proc. Natl. Acad. Sci. 106 (2009) 18908–18913.
[58] R. Balasubramanlanian, S.M. Smith, S. Rawat, L.A. Yatsunyk, T.L. Stemmler, A.C. [92] V.N. Beznis, B.M. Weckhuysen, J.H. Bitter, Catal. Lett. 138 (2010) 14–22.
Rosenzweig, Nature 465 (2010) 115–119. [93] G.E.A. Steghuis, J.G. van Ommen, K. Seshan, J.A. Lercher, Stud. Surf. Sci. Catal. 107
[59] M. Merkx, D.A. Kopp, M.H. Sazinsky, J.L. Blazyk, J. M€uller, S.J. Lippard, Angew. (1997) 403–412.
Chem. Int. Ed. 40 (2001) 2782–2807. [94] P. Vanelderen, R.G. Hadt, P.J. Smeets, E.I. Solomon, R.A. Schoonheydt, B.F. Sels, J.
[60] B.J. Wallar, J.D. Lipscomb, Chem. Rev. 96 (1996) 2625–2657. Catal. 284 (2011) 157–164.
[61] Y. Shiota, G. Juhász, K. Yoshizawa, Inorg. Chem. 52 (2013) 7907–7917. [95] E.M. Alayon, M. Nachtegaral, M. Ranocchiari, J.A. van Bokhoven, Chem. Commun.
[62] R. Balasubramanlanian, A.C. Rosenzweig, Acc. Chem. Res. 40 (2007) 573–580. 48 (2012) 404–406.
[63] P. Vanelderena, R.G. Hadtb, P.J. Smeetsa, E.I. Solomon, R.A. Schoonheydta, B.F. Sels, [96] P.J. Smeets, M.H. Groothaert, R.A. Schoonheydt, Catal. Today 110 (2005) 303–309.
J. Catal. 284 (2011) 157–164. [97] J.R. Anderson, P. Tsai, J. Chem. Soc. Chem. Commun. (1987) 1435–1436.
[64] S.I. Chan, C.Y.-C. Chien, C.S.-C. Yu, P. Nagababu, S. Maji, P.P.-Y. Chen, J. Catal. 293 [98] B. Michalkiewicz, Appl. Catal. A Gen. 277 (2004) 147–153.
(2012) 186–194. [99] E.V. Starokon, M.V. Parfenov, L.V. Pirutko, S.I. Abornev, G.I. Panov, J. Phys. Chem. C
[65] P. Nagababu, S. Maji, M.P. Kumar, P.P.-Y. Chen, S.S.-F. Yu, S.I. Chana, Adv. Synth. 115 (2011) 2155–2161.
Catal. 354 (2012) 3275–3282. [100] E.V. Starokon, M.V. Parfenov, S.S. Arzumanov, L.V. Pirutko, A.G. Stepanov, G.I.
[66] S.I. Chan, Y.-J. Lu, P. Nagababu, S. Maji, M.-C. Hung, M.M. Lee, I.J. Hsu, P.D. Minh, J.C.- Panov, J. Catal. 300 (2013) 47–54.
H. Lai, K.Y. Ng, S. Ramalingam, S.S.-F. Yu, M.K. Chan, Angew. Chem. 125 (2013) [101] B.R. Wood, J.A. Reimer, A.T. Bell, M.T. Janicke, K.C. Ott, J. Catal. 225 (2004) 300–306.
3819–3823. [102] M.V. Parfenov, E.V. Starokon, L.V. Pirutko, G.I. Panov, J. Catal. 318 (2014) 1–13.
[67] P. Nagababu, S.S.-F. Yu, S. Maji, R. Ramu, S.I. Chan, Catal. Sci. Technol. 4 (2014) [103] N.V. Beznis, A.N.C. van Laak, B.M. Weckhuysen, J.H. Bitter, Microporous Mesopo-
930–935. rous Mater. 138 (2011) 176–183.
[68] V.S. Arutyunov, J. Nat. Gas Chem. 13 (2004) 10–22. [104] T.J. Hall, J.S.J. Hargreaves, G.J. Hutchings, R.W. Joyner, S.H. Taylor, Fuel Process.
[69] M.J. Brown, N.D. Parkyns, Catal. Today 8 (1991) 305–335. Technol. 42 (1995) 151–178.
[70] S. Grundner, M.A.C. Markovits, G. Li, M. Tromp, E.A. Pidko, E.J.M. Hensen, A. Jentys, [105] Dowden, D. A.; Walker, G. T. Brit Pat, 1,244,00 L, 1971.
M. Sanchez-Sanchez, J.A. Lercher, Nat. Commun. 6 (7546) (2015) 1–9, http://dx. [106] Stroud, H. J. F. Brit Pat, 1,398,385, 1975.
doi.org/10.1038/ncomms8546. [107] R.X. Liu, M. Iwamoto, J.H. Lunsford, J. Chem. Soc. Chem. Commun. (1982) 78–79.
[71] C.G. Dallos, V. Kafarov, R.M. Filho, Chem. Eng. J. 134 (2007) 209–217. [108] H.F. Liu, R.S. Liew, K.Y. Liu, R.E. Johnson, J.H. Lunsford, J. Am. Chem. Soc. 106 (1984)
[72] S.S. Verma, Energy Convers. Manag. 43 (2002) 1999–2008. 4117–4123.
[73] M.C. Alvarez-Galvan, N. Mota, M. Ojeda, S. Rojas, R.M. Navarro, J.L.G. Fierro, Catal. [109] S.H. Taylora, J.S.J. Hargreaves, G.J. Hutchings, R.W. Joyner, C.W. Lembacher, Catal.
Today 171 (2011) 15–23. Today 42 (1998) 217–224.
[74] G.A. Foulds, B.F. Gray, Fuel Process. Technol. 42 (1995) 129–150. [110] X. Zhang, D. He, Q. Zhang, B. Xu, Q. Zh, Top. Catal. 32 (2005) 215–223.
[75] Q. Zhang, D. He, J. Li, B. Xu, Y. Liang, Q. Zhu, Appl. Catal. A Gen. 224 (2002) 201–207. [111] Durante, V. A.; Walker, D. W.; Gussow, S. M.; Lyons, J. E. U.S. Patent, 4,918,249,
[76] A. Holmen, Catal. Today 142 (2009) 2–8. 1990.
[77] R. Lødeng, O.A. Lindvag, P. Soraker, P.T. Roterud, O.T. Onsager, Ind. Eng. Chem. Res. [112] J.E. Lyons, P.E. Ellis Jr., V.A. Durante, Stud. Surf. Sci. Catal. 67 (1991) 99–116.
34 (1995) 1044–1059. [113] S. Betteridge, C.R.A. Catlow, D.H. Gay, R.W. Grimes, J.S.J. Hargreaves, G.J. Hutchings,
[78] K. Tabata, Y. Teng, T. Takemoto, E. Suzuki, M.A. Banares, M.A. Pena, J.L.G. Fierro, R.W. Joyner, Q.A. Pankhurst, S.H. Taylor, Top. Catal. 1 (1994) 103–110.
Catal. Ver. Sci. Eng. 44 (2002) 1–58. [114] V.A. Durante, D.W. Walker, W.H. Seitzer, J.E. Lyons, Pacifichem 89 (1989) 23–26.
[79] V.S. Arutyunov, Catal. Today 215 (2013) 243–250. [115] J.-W. Chun, R.G. Anthony, Ind. Eng. Chem. Res. 32 (1993) 259–263.
[80] T. Takemoto, K. Tabata, Y. Teng, A. Nakayama, E. Suzuki, Appl. Catal. A Gen. 205 [116] M. Stolcová, R. Polnisera, M. Hronec, M. Mikula, Appl. Catal. A Gen. 400 (2011)
(2001) 51–59. 122–130.
[81] Y. Yamaguchi, Y. Teng, S. Shimomura, K. Tabata, E.J. Suzuki, Phys. Chem. A 103 [117] Z. Sojka, R.G. Herman, K. Klier, J. Chem. Soc. Chem. Commun. 185-186 (1991).
(1999) 8273–8278. [118] J.A. Barbero, M.C. Alvarez, M.A. Bañares, M.A. Peña, Fierro, J. L. G. Chem. Commun.
[82] Y. Teng, H. Sakurai, K. Tabata, E. Suzuki, Appl. Catal. A Gen. 190 (2000) 283–289. 1184-1185 (2002).
[83] R.S. Arutyunov, O.V. Krylov, Russ. Chem. Rev. 74 (2005) 1111–1137. [119] A.L. Cardoso, R. Augusti, M.J. da Silva, J. Am. Oil Chem. Soc. 85 (2008) 555–560.
[84] J.S.J. Hargreaves, G.J. Hutchings, R.W. Joyner, Nature 348 (1990) 428–429. [120] S. Casztelan, J.B. Muffat, J. Catal. 106 (1987) 512–524.
[85] T. Takemoto, D. He, Y. Teng, K. Tabata, E. Suzuki, J. Mol. Catal. A 179 (2002) [121] S. Mansouri, O. Benlounes, C. Rabia, R. Thouvenot, M.M. Bettahar, S. Hocine, J. Mol.
279–286. Catal. A 379 (2013) 255–262.
[86] T. Takemoto, D. He, Y. Teng, A. Nakayama, K. Tabata, E. Suzuki, J. Catal. 198 (2001)
109–115.

Vous aimerez peut-être aussi