You are on page 1of 58

Catalysis Reviews

Science and Engineering

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/lctr20

A review on state-of-the-art catalysts for methane


partial oxidation to syngas production

T.J. Siang, A.A. Jalil, S.Y. Liew, A.H.K. Owgi & A.F.A. Rahman

To cite this article: T.J. Siang, A.A. Jalil, S.Y. Liew, A.H.K. Owgi & A.F.A. Rahman (2022): A
review on state-of-the-art catalysts for methane partial oxidation to syngas production, Catalysis
Reviews, DOI: 10.1080/01614940.2022.2072450

To link to this article: https://doi.org/10.1080/01614940.2022.2072450

Published online: 23 May 2022.

Submit your article to this journal

Article views: 452

View related articles

View Crossmark data

Citing articles: 1 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=lctr20
CATALYSIS REVIEWS
https://doi.org/10.1080/01614940.2022.2072450

A review on state-of-the-art catalysts for methane partial


oxidation to syngas production
a a,b a
T.J. Siang , A.A. Jalil , S.Y. Liew , A.H.K. Owgia, and A.F.A. Rahmana
a
School of Chemical and Energy Engineering, Faculty of Engineering, Universiti Teknologi Malaysia,
Johor, Malaysia; bCentre of Hydrogen Energy, Institute of Future Energy, Universiti Teknologi Malaysia,
Johor, Malaysia

ABSTRACT ARTICLE HISTORY


The rapid and severe deactivation of methane partial oxidation Received 18 Decmber 2021
catalysts remains a major hindrance restraining its potential in Revised 5 April 2022
commercialization and industrialization for large-scale syngas Accepted 27 April 2022
production. It is imperative to provide an in-depth understand­ KEYWORDS
ing discission about the intrinsic and synergistic interactions of Synergistic interaction;
catalyst components toward the catalyst efficiency during methane oxidation;
reforming reactions. This review presents a contemporary eva­ combustion-reforming;
luation of recent works on synergistic relationship among cata­ syngas; heterogeneous
lyst components (support, active metals and catalyst structure) catalysis; catalyst design
during methane partial oxidation using theoretical and state-of-
the-art experimental procedures. Advancements achieved
through this synergistic relationship not only enhance proper­
ties of existing catalysts but also leading to discovery and
development of novel catalyst systems. Thermodynamics, reac­
tion mechanisms, catalytic performances, catalyst deactivation
induced by carbonaceous deposition and reaction kinetic mod­
eling have been successfully explored and described using
information from these essential interactive factors over these
decades. This viewpoint explains the roles of the interactions
and their functions toward exploration of efficient catalysts
systems for industrial applications.

1. Introduction
The depletion of fossil fuel and anthropogenic CO2 emission has grown
exponentially due to the large-scale industrial combustion driven by the
increasing global energy demand and economic expansion in recent years.
Figure 1 illustrates the contribution of sectors to global CO2 gas emission.[1]
Notably, the energy sector was the main contributor with 73.2%, whereby the
energy use in industry (24.2%) was the leading factor, followed by energy use
in buildings (17.5%) and transport (16.2%). The fossil fuel-based combustion
(i.e., coal, natural gas and petroleum) reportedly resulted in 33.5 gigatons of
global energy-related CO2 emission in 2019 due to intensive population
growth and it predictably grows by an additional 19.2% in 2050.[2,3] Thus,

CONTACT A.A. Jalil aishahaj@utm.my School of Chemical and Energy Engineering, Faculty of Engineering,
Universiti Teknologi Malaysia, 81310 UTM Johor Bahru, Johor, Malaysia; Centre of Hydrogen Energy, Institute of
Future Energy, Universiti Teknologi Malaysia, 81310 UTM Johor Bahru, Johor, Malaysia
© 2022 Taylor & Francis
2 T. J. SIANG ET AL.

[1]
Figure 1. Global co2 gas emission by various sectors. Reproduced with permission from Ref.
Copyright 2020, ourworldindata.org.

there is a worldwide concern regarding the environment degradation and


climate changes as they are closely associated with the present carbon-
intensive energy system.[4] In this context, substituting current fossil fuel-
based energy with less carbon-intensive energy sources and reducing the
substantial dependency on nonrenewable energy are an indispensable and
urgent mission. Syngas (mixture of H2 and CO) emerges as a prospective
alternative, which provides feasible building blocks for the downstream pro­
duction such as hydrogen for fuel cell, methanol in petrochemical industry
and synthetic fuels from Fisher-Tropsch synthesis (FTS).[5]
Although renewable energy is the long-term key to overcome the afore­
mentioned problems, syngas is presently the feasible short-term solution
acting as the cushion to smooth the transition toward an effective low-
carbon energy system in near future.[6] To date, the market of syngas and its
derivatives was predicted at 2,45,557 MWth and was forecasted to achieve
4,06,860 MWth by 2025 with a 10.6% of compound annual growth rate in
between 2020 and 2025.[7] Such a projection was closely related to the devel­
opment of the chemical industry in different applications, in particular, the
CATALYSIS REVIEWS 3

Figure 2. Growth rate of global syngas market by region forecast in 2022‒2027.

chemical industry preferentially uses feedstocks that are cost-efficient for


procurement and production. The syngas market was mainly dominated by
four major applications including chemicals, gaseous fuels, liquid fuels and
power generation.[8] As seen in Figure 2, the Asia-Pacific region predictably
dominates the global syngas market with increasing demand from the chemi­
cal industry, primarily from refineries in China. In fact, the total installed
capacity of electric power generation in China has increased from 2011 GW
(2019) to 2200 GW (2020), with a growth rate of around 9%. Meanwhile, the
di-ammonium phosphate production in India has grown about 15% from
3.89 million tons in 2019 to 4.48 million tons in 2020, which indirectly resulted
in a rising demand of syngas.[8] Hence, these factors are expected to exert
a momentous influence on the syngas market in the coming years.
Additionally, there are five existing commercial-scale gas-to-liquids (GTL)
plants implement FTS coupled with syngas production technologies providing
259 Mbpd of synthetic fuel, whilst three additional GTL facilities with similar
configuration are under development across the world.[9]
Amongst the syngas production technologies, methane steam reforming
(MSR) is presently the conventional technology to be deployed in industrial
applications due to its abundancy of feedstock (i.e., water and methane from
natural gas) and high selectivity of H2.[6,10] However, MSR yields a H2/CO
ratio of 3 with considerable amount of concomitant CO2 emission, making it
unfavorable for direct downstream processes. From industrial and environ­
mental perspectives, the research emphasis has shifted toward methane dry
reforming (MDR) as it transforms greenhouse CO2 into syngas with a ratio of
unity and advanced MSR, namely, methane autothermal reforming, a process
comprises MSR and methane partial oxidation (MPO) where the energy
4 T. J. SIANG ET AL.

requirement of endothermic MSR can be compensated by the heat release


from exothermic MPO and thus improvement of overall reactor efficiency.[11]
Nevertheless, the abovementioned technologies require auxiliary separation
and purification units for adjusting syngas ratio to 2, which imposes the overall
capital cost.[12,13] From the above reasons, MPO has appeared as a prominent
candidate to substitute other methane reforming technologies. In addition, the
MPO technology reportedly offers 10‒15% and 25‒30% lower in the energy
requirement and capital investment, respectively, compared to MSR.[14] In
fact, partial oxidation technology has been implemented in 12 world-wide
industrial plants by the Linde group, namely, partial oxidation (POX) plants,
which convert various kind of C-containing compounds into syngas and other
petrochemicals. Since few decades ago, one of the four world’s largest POX
plants continuously supply 200,000 Nm3/h of syngas to the global gas market
till today.[15,16]
The concept of syngas production via catalytic MPO was initially revealed
by Liander et al.[17] in 1929 during investigation of the ammonia process and
a characteristic study of MPO over Ni-based catalysts was firstly published by
Prettre et al.[18] in 1946. Based on the results of temperature profiles corre­
sponding to the bed exit, they found that the reaction process mainly compri­
sedtwo stages, whereby the CH4 total oxidation generating the CO2 and H2O,
followed by MSR and MDR to yield H2 and CO, knowing as combustion-
reforming route in the latter studies. In 1971, Huszar et al.[19] studied the
single grain of Ni/mullite for MPO and suggested that the reactions were
restrained by the O2 diffusion through the gas film encircling the catalyst due
to the formation of NiAl2O4. On this basis, Gavalas et al.[20,21] investigated the
influences of calcination and pre-reduction temperatures on alumina sup­
ported Ni catalysts and reported that CO2 and H2O gases were the main
products as a result of dominant combustion process. In contrast to the
early success of MSR study, catalytic MPO publications only emerged in
1990s, with most studies focusing on the catalyst materials screening.
Ashcroft et al.[22] screened the catalytic performances of Ru-containing oxides,
namely, RuO2, Ru/Al2O3 and a series of MRu2O7 (M = Gd, Tb, Pr, Sm, Eu, Dy,
Tm, Yb and Lu) for the MPO activity. They reported that the syngas selectivity
and CH4 conversion could be attained in a range of 90 to 99% at reaction
temperature above 750°C and atmospheric environment.
In the late 1990s, the emphasis of study began to shift toward the enhance­
ment of catalytic activity and stability through the fundamental understanding
of elementary steps for MPO reaction. Hickman et al.[23,24] conducted the
MPO simulation considering 19 elementary steps, namely, adsorption, deso­
rption and surface reactions over Rh and Pt particles surface. They attributed
the disparity in product selectivity to the energy barriers of particular reac­
tions, specifically the OH formation was relatively easier to occur on Pt surface
than Rh surface. Conversely, the dissociative CO2 adsorption was reportedly
CATALYSIS REVIEWS 5

much easier to proceed on Rh surface. Subsequently, these intrinsic differences


in elementary reaction contributed to explain the catalytic behaviors of Rh and
Pt coated monoliths. The advantageous effect of noble metal as promoter was
illustrated by Choudhary et al.[25] instead of the use as hosting metal. They
reported that 0.1 wt% noble metal promotion could reduce the ignition
temperature at about 18% and 25% with Pt and Pd promoter, respectively.
Such catalytic behavior of noble metal promoter was also suggested by Ji
et al.[26]
Catalytic MPO technology undeniably offers practical syngas content for
wide-ranging downstream industrial applications with higher cost- and
energy-efficiency, but it encounters several challenges adversely affecting its
catalysis activity as MPO owns a complex system consisting of multiple
reactions.[27] In general, metal group VIII, IX and X such as Ni, Co, Fe, Ru,
Rh, Pd, Ir and Pt, are widely acknowledged as active metals for methane
oxidation studies.[28,29] Despite the non-noble metal catalysts are extensively
investigated for methane oxidation studies, the severe deterioration in catalysis
activity due to carbon deposition and hot spot formation at catalyst bed has
confined their practicability for industrial purposes. In this context, the
emphasis on development of catalyst design respective to noble metal has
been renewed with arising concerns from both academic and industrial realms
since it possesses great resistance toward high temperature, reoxidation and
carbon deposition as well as excellent ability of CH4 scission.[30–33]
Particularly, carbon deposition is unavoidable during CH4 reforming pro­
cesses at high-temperature zone even though in the presence of noble metal
as the active hosting material for a catalyst system.[34,35] Additionally, the
competitive adsorption of oxidants on catalyst surface under the exceedingly
oxidative atmosphere could influence the extent of carbon elimination.[6]
Therefore, exploring a novel catalyst system presents high accessibility of
reactants to active sites at a molecular level to facilitate the surface reaction
enhancing syngas formation and simultaneously suppressing carbon deposi­
tion is essential in MPO.
To date, numerous strategic designs of MPO catalyst have been introduced
and investigated by tuning the extrinsic or intrinsic attributes of the catalyst
components to address the abovementioned setbacks in MPO process.
Advanced design of MPO catalysts essentially involved the deployment of
intrinsic properties of catalyst components and elucidation of their complex
synergetic relations for optimal performance and catalyst lifespan by employ­
ing state-of-the-art experimental and theoretical descriptions. Catalysts of
various configurations, morphologies and topologies have been reviewed
over the years to determine the effects of synergistic component relations on
catalyst properties such as metal-support interaction, reducibility, oxygen
vacancy and mobility, surface basicity/acidity, particle size and distribution,
heat and mass transfer limitations. Several works performed in MPO have
6 T. J. SIANG ET AL.

demonstrated that establishing an excellent heat stable and carbon-resistant


catalyst by adjusting only one parameter is challenging. In order to design
a robust and efficient MPO catalyst, it is essential to take into account the
collaborative interaction of multiple parameters. It has been found that these
interactions have been closely related to ascertaining the MPO thermody­
namics, kinetics and mechanisms. The MPO catalyst system is what concerns
the presented review and the literature has extensively been systematized
based on the actual catalyst designs associated with the fundamental under­
standing of reaction mechanism. There are a number of literature review for
this topic.[36–38] Therefore, an extensive description of MPO technology is
unnecessary, but additional comprehensive summaries have been lacking.
Here, we have first attempted to go from the thermodynamics to reaction
mechanism of MPO, showing how the difference in MPO reaction pathway
due to different oxidation state of metal and interaction with the support and
reactant affecting the catalytic process and desired product selectivity. To
focus the topic of this review, we provide in-depth discussions about how
the state-of-the-art catalyst designs tackle the fundamental reaction mechan­
ism from aspects of metal-support interaction, surface acidity and basicity,
oxygen vacancies, perovskite and surface morphology in order to achieve an
efficient MPO activity. In the following sections, we also summarize the
significant findings of MPO kinetic studies depending on the reaction condi­
tions and some suggestions are given in the conclusion for near future works.

2. Thermodynamics and reaction mechanism


2.1. Thermodynamics, heat and mass transport

Chemical reactions are undeniably dominated by the thermodynamics of their


reaction equilibria. As a complex system such as partial oxidation of methane
consisting of multiple reaction equilibrium, thereby the desired products or
unwanted by-product can be achieved through several reaction routes.[39–42]
In this context, thermodynamic evaluation is useful which serves as a guide for
understanding and regulating complex chemical reactions.[43–45] By consider­
ing the estimated product distribution as a function of temperature (Figure 3)
and Gibbs free energy change calculations for plausible reactions (Table 1 and
Figure 4), the reaction routes for MPO to syngas production can be deter­
mined thermodynamically.[46]
Thermodynamically, negative value of Gibbs free energy change suggests
the propensity of reaction to occur. As summarized in Table 1, the MPO (R3)
can be perceived as of the direct route to produce syngas owing to its most
negative value of Gibbs free energy change (Figure 4). This indicated that both
CH4 and O2 could be readily activated at all temperature and this is widely
supported by numerous studies in past decades.[47–51] However, the question
CATALYSIS REVIEWS 7

Figure 3. (A) Equilibrium composition and (b) equilibrium reactant conversions and H2/CO molar
ratio for methane partial oxidation at various temperatures and atmospheric pressure calculated
by HSC chemistry 6. Reproduced with permission from Ref. [46] Copyright 2020, Elsevier.

of whether it is the only reaction to catalyze has been a keynote in the MPO
investigations since the remaining observed reactions in Table 1 typically play
crucial roles in the MPO to affect the overall thermodynamic equilibrium in
the system. As demonstrated in Figure 3a, the absence of O2 with substantial
equilibrium amount of both H2O and CO2 suggested the dominance of CH4
combustion (R4, also known as CH4 total oxidation) at low-temperature
regions which was responsible for the O2 equilibrium conversion of 100%
(Figure 3b). At increasing temperature, the subsequent oxidizing agents were
further consumed to produce syngas via MSR (R1) and MDR (R2), known as
indirect combustion-reforming route.[46,52–56] This postulation is further cor­
roborated by the calculated Gibbs free energy change profiles of MSR and
MDR in Figure 4a, where the Gibbs free energy changes shifted from positive
into negative value at temperature above 600°C. As a result, it could be
concluded that the catalytic MPO is constituted by co-occurrence of direct
and indirect routes depending on the inherent nature of catalyst system during
reaction.
D and ND represent dominant and not dominant, respectively, while ND‒
D (or D‒ND) represents from not dominant to dominant (or vice versa).
From our previous thermodynamic study,[46] the CO dehydrogenation
(R11), Boudouard reaction (R12), CO2 hydrogenation (R13) and CH4 decom­
position (R14) were identified as main factors for carbon formation during
8

Table 1. Summary of propensity of primary and side reactions occur during reaction based on gibbs free energy change calculations performed by HSC chemistry 6.
Reproduced with permission from Ref. [46] Copyright 2020, Elsevier.
T. J. SIANG ET AL.

Temperature regions
Region 1 Region 2 Region 3 Region 4
Reactions Equations No. ΔG (kJ mol−1) (200‒400°C) (400‒600°C) (600‒800°C) (800‒1000°C)
Reforming reactions
Methane steam reforming (MSR) CH4 þ H2 O ! 3H2 þ CO R1 −0.2446 T + 150.85 ND ND‒D D D
Methane dry reforming (MDR) CH4 þ CO2 ! 2H2 þ 2CO R2 −0.2797 T + 179.38 ND ND‒D D D
Methane partial oxidation (MPO) CH4 þ 0:5O2 ! 2H2 þ CO R3 −0.1921 T – 79.924 ND ND‒D D D
Non-coke forming reactions
CH4 combustion CH4 þ 2O2 ! CO2 þ 2H2 O R4 0.0005 T – 800.78 D D D D
CO2 methanation 4H2 þ CO2 ! CH4 þ 2H2 O R5 0.2096 T – 122.32 D D‒ND ND ND
CO methanation 3H2 þ CO ! CH4 þ H2 O R6 0.2446 T – 150.85 D D‒ND ND ND
Water-gas shift (WGS) H2 O þ CO ! CO2 þ H2 R7 0.0351 T – 28.534 D D‒ND ND ND
Revere water-gas shift (RWGS) CO2 þ H2 ! H2 O þ CO R8 −0.0351 T + 28.534 ND ND ND-D D
Carbon gasification by CO2 CðsÞ þ CO2 ! 2CO R9 −0.177 T + 124.14 ND ND ND-D D
Carbon gasification by H2O CðsÞ þ H2 O ! H2 þ CO R10 −0.1419 T + 95.604 ND ND ND-D D
Coke forming reactions
CO dehydrogenation H2 þ CO ! H2 O þ CðsÞ R11 0.1419 T – 95.604 D D‒ND ND ND
Boudouard reaction 2CO ! CðsÞ þ CO2 R12 0.177 T – 124.14 D D D‒ND ND
CO2 hydrogenation 2H2 þ CO2 ! CðsÞ þ 2H2 O R13 0.1069 T – 67.069 D D‒ND ND ND
CH4 decomposition CH4 ! CðsÞ þ 2H2 R14 −0.1027 T + 55.246 ND ND‒D D D
CATALYSIS REVIEWS 9

Figure 4. Gibbs free energy of (a) reforming reactions, (b) non-coke forming reactions and (c) coke
forming reactions as a function of temperature at pressure of 1 bar. Reproduced with permission
from Ref. [46] Copyright 2020, Elsevier.

MPO. Based on the Gibbs free energy changes profiles of coke-forming


reactions (Figure 4c), the CH4 decomposition was the only reaction favors at
elevated temperatures whereas the rest were pronounced at low-temperature
regions. Similar conclusions were also suggested by Jang et al.,[43] Cao et al.[44]
and Jabbour et al.[57] In addition, concerning the non-coke forming reactions
in Table 1 and Figure 4b, the parallel existence of reverse water-gas shift (R8),
carbon gasification by CO2 (R9) and H2O (R10) has a major influence on the
H2/CO ratio (Figure 3b), depending on the equilibrium at varying temperature
in MPO system.[46]
As the MPO constitutes of direct and indirect routes, the existing
large quantity of oxidizing agents (i.e., O2, CO2 and H2O) may have
large implications associated with heat and mass transport on the results
attained by experimental works. Typically, the occurrence of hot spots
hinders the assessment of experimental data and the depiction of
mechanism. Wang et al.[58] investigated the changes in temperature
and product distribution along with the reactor length for MPO process,
as shown in Figure 5. They assigned the initial decline in temperature
(from 0 to 1 cm) to the presence of endothermic MSR and MDR,
resulting in the formation of H 2 and CO, followed by a decreased
concentration of CO 2 and H2O (Figure 5b). The subsequent rise in
temperature (1 to 6 cm of reactor length) and the syngas promotion
affirmed that partial and total oxidation of CH4 occurred in the hot zone
at the front part of catalyst bed whilst the MSR and MDR occurred after
the section. This observation was in line with the findings of Li et al.,[59]
where the temperature difference in bed surface temperature (front part
of catalyst bed) was much varied than other sections.
10 T. J. SIANG ET AL.

Figure 5. (a) Schematic drawing of the reactor; (b) the axial temperature and species composition
profiles in the reactor during the MPO reaction at 900°C and GHSV of 37,000 mL g−1 h−1).
Reproduced with permission from Ref. [58] Copyright 2018, Elsevier.

The formation of hot spot is mainly contributed by the exothermic reac­


tions, especially to CH4 total oxidation during MPO process. Holmen and the
coworkers [27] suggested a further insight into the correlation between mass
diffusivity and thermal diffusivity contributing to the formation of hot spots.
They found that heat diffusion was slightly slower than mass diffusion in gas
phase as a function of temperature and thus leading to formation of hot spots
due to temperature gradients. The formation of hot spots on gas temperature
gradient and catalyst surface complicates the evaluation of kinetics of MPO
and adversely affect the catalytic stability. This postulation was supported by
the computational modeling and simulations for partial oxidation of CH4 or
hydrocarbons in combination with typical kinetics.[60–66] In order to solve the
issue of temperature gradients in kinetic evaluations, numerous efforts have
been dedicated to studies for minimizing the formation of hot spots such as (i)
introducing catalytic materials with excellent thermal resistance and conduc­
tivity to distribute excess heat [58,67,68]; (ii) adding diluted gas feed (> 90%
dilutant) to attain isothermal gas temperatures during reaction [69,70]; (iii)
coupling MPO with endothermic MSR or MDR to utilize the excess heat for
producing syngas [71,72]; and (iv) chemical looping of MPO from preventing
excessive heat produced by CH4 total oxidation.[73,74]

2.2. Reaction mechanism and carbon formation


Considering the reported MPO studies involving density functional theory
(DFT) calculations,[31,75–78] transient pulse [79–82] and in situ experimental
evaluations,[83–86] the MPO mechanism can be generally categorized as direct
and indirect routes to produce syngas, as shown in Figure 6. The speculation of
direct MPO route has been established since 1990s, whereby dissociative
CATALYSIS REVIEWS 11

Figure 6. Pictorial illustration of direct and indirect routes for CH4 partial oxidation process.

adsorption and activation of both CH4 and O2 occur at metal surface and
support surface, respectively. Even though the energy required for dissociation
of CH4 molecules is closely related to the catalyst surface properties, it is
widely accepted that the CH4 dissociation is always the rate determining step
to initiate MPO reaction. The dissociated CHx species preferably adsorbs on
a site completing tetravalency, with adsorbed CH3 locates on top of metal
atoms and CH2 binding between two adjacent metal atoms forming a bridged
adsorption, which leads to the formation of adsorbed hydrogen and adsorbed
carbon species. The adsorbed hydrogen species react with adjacent one to
produce H2 gas and subsequently leaves the catalyst surface. Meanwhile, the
O2 molecules dissociatively adsorb on the support surface with O-only coor­
dination to form adsorbed surface oxygen species, which is normally a fast
process. With the participation of adsorbed surface oxygen species, adsorbed
carbon species at metal surface can be converted into CO gas. Apart from
aforementioned direct CH4 decomposition, adsorbed CH4 molecules (at metal
surface) can be dissociated by reacting with adsorbed surface oxygen species
(at support surface) to form adsorbed CHx and OH species. Besides the main
12 T. J. SIANG ET AL.

syngas product, some byproducts such as H2O and CO2 are possibly formed
during direct MPO route, in which adsorbed hydrogen species react with
adsorbed OH groups to yield H2O gas whereas the CO gas is adsorbed on
support surface and combine with adsorbed surface oxygen species to generate
CO2 gas.
In the case of indirect MPO mechanism route, as CH4 and O2 gases
continuously move through the catalyst bed, it is commonly believed that
CH4 combustion is dominating the front section, followed by CH4 reforming
processes (i.e., MSR and MDR) in the remaining section of catalyst bed.
Similarly, the CH4 molecules are dissociatively adsorbed on the metal surface
to form CHx intermediates whereas the adsorbed surface oxygen species are
originated from O2 molecules adsorbed and activated at support surface.
Subsequently, adsorbed surface oxygen has possibility to react with adsorbed
CHx species and thus formation of adsorbed CHxO intermediate. Some
authors suggest the adsorbed CHxO intermediate to be the precursor to CO2
and H2O formation under the exceedingly oxidative environment.[78,87] Other
authors suggested that concentration of adsorbed surface oxygen on catalyst
surface is the key factor for direct route of MPO to produce H2 and CO, thus
preventing the tendency of total oxidation.[78,80,86] In a series of classical
computational and experimental investigations about the short-contact-time
MPO process,[60,88] Schmidt and the coworkers verified the existence of short
oxidation zone at the catalyst entrance, followed by reforming zone in the
remaining catalyst bed even at high space velocity. They also found that the
syngas selectivity was drastically reduced at high space velocity. This could
indicate that concentration of adsorbed surface oxygen on catalyst surface is
independent from space velocity.
In the following section of catalyst bed (CH4 reforming zone), the subse­
quent H2O produced from CH4 combustion further adsorb and dissociate on
the support surface and as a result, the formation of hydroxyl groups. Notably,
the adsorption and dissociation of CO2 can be categorized into three different
ways reliant on the surface structure and defects of support, namely, (i) O-only
coordination, of which both O atoms bind with surface (ii) C-only coordina­
tion and (iii) O and C coordination where both O and C atoms bond with
catalyst surface whilst another O atom expose to gas phase. In general, these
dissociation steps of both H2O and CO2 are considerably rapid. Similarly, the
CH4 adsorb and dissociate into adsorbed CHx species on metal surface in the
reforming zone. With the participation of adsorbed surface oxygen or hydro­
xyl groups, the adsorbed CHx species can be transformed into adsorbed CHx
O species or adsorbed CO. In fact, some authors reached an agreement that the
adsorbed CHxO species is the main contributory factor to the H2 and CO
formation whereas other authors argued that the CO formation is a result of
reduction between carbonates from adsorbed CO2 and carbon deposit on
metal surface. In addition, other research groups claimed that CO is directly
CATALYSIS REVIEWS 13

originated from adsorbed CO2 instead of adsorbed CHxO species. To date,


there is no clear consensus regarding the details for MPO mechanism on
catalyst surface since it is highly depending on the intrinsic catalyst nature
and reaction conditions. Nevertheless, the formation or dissociation of
adsorbed CHxO species to produce H2 and CO is reportedly the rate deter­
mining step and this is a dovetailed agreement among most of CH4 reforming
studies.
Given the MPO constitutes of reforming processes consuming CH4 as the
main feedstock, the carbonaceous deposition toward catalyst deactivation is
another challenging topic for the system. Generally, the formation of carbo­
naceous species on catalyst surface is unavoidable during hydrocarbon
reforming reactions, attributed to the parallel hydrocarbon decomposition at
elevated temperature.[34,35,89] Interestingly, the degree of catalyst deactivation
is closely related to the reactiveness of carbon deposit with different structural
type as function of temperature.[12,89] To determine the mechanism of carbon
growth on metal surface, Wu et al.[90] conducted an in situ transmission
electron microscopy (TEM) experiment to provide a further insight into the
carbon encapsulation. As seen in Figure 7, the multiple partial carbon layers
started to nucleate and grow from the surface of Pt nanoparticle without the
completed encapsulation on the nanoparticle. Over time, the surface of nano­
particles has altered due to the formation of new steps through new nucleated
and developed carbon layers. Based on the DFT calculation from Xu and

Figure 7. Two time-lapsed series of TEM images of (a) 0 min, (b) 50 min, (c) 68 min, (d) 0 min, (e)
61 min and (f) 76 min of Pt nanoparticles as a function of reaction time. white arrows indicate the
location of step formation as carbon layers encapsulate the nanoparticle. Reproduced with
permission from Ref.[90] Copyright 2016, Elsevier.
14 T. J. SIANG ET AL.

Saeys,[91] this observation could be attributed to the surface atomic carbon


diffused into the subsurface layer of the metal particles and subsequently grew
to form graphene islands. Eventually, the developed graphitic layers comple­
tely encapsulated the active sites of metal particles, leading to a drastic loss in
catalytic performance and catalyst deactivation. Several general reviews on
carbon deposition of the actual catalyst system can be found in the
literature.[34,35,89,92,93] Thus, the development of catalyst system has always
been emphasized on the carbon elimination in order to prolong the catalyst
lifespan for MPO activity.

3. Correlation between oxidation state of metal and MPO pathway


Fundamentally, the existence of active sites activates the dissociation of reac­
tant forming reactive intermediates and subsequently combine to produce
product via surface reaction.[94] In MPO reaction, the metals typically serve
as active sites for CH4 dissociation while acidic/basic supports are responsible
for the dissociation of oxidants (i.e., O2, H2O and CO2).[95,96] However, the
MPO also constitutes indirect combustion-reforming pathway where CH4
total oxidation occurs at front part of catalyst bed and subsequent reforming
processes at the following catalyst bed and thus, the changes in oxidation state
of the metal for surface reaction is of interest toward the MPO
performance,[97–100] which serves as a guide for catalyst design. In this section,
the MPO pathway and performance associated with the state of catalyst have
been discussed based on advanced characterization techniques in recent years.
To determine the correlation between MPO performance and oxidation
state changes in Pd, Stötzel et al.[101] studied the oscillatory behavior of MPO
over the dynamic structural changes in Pd catalysts by using quick-scanning
extended X-ray absorption fine structure (QEXAFS) approach. Combining the
results of linear combination analysis and MS signal for observed oscillations
during MPO reaction, they noticed that the period of hydrogen generation
(green box in Figure 8) was highly correlated to the duration of metallic Pd0
formation, particular to the lower section of the catalyst bed where the metallic
Pd0 form lasted longer due to the dominance of CH4 reforming. This was
indicative of the presence of combustion-reforming process. Additionally, this
finding revealed the specific turning point of interchangeable oxidized/
reduced state of Pd (green circle in Figure 8) and the corresponding formation
of different gaseous products. A similar observation is also reported by Velasco
and his coworkers [102] in an investigation of MPO over Ni and Ru catalysts.
The concentration of metallic surface area of catalysts has appreciable influ­
ence on the MPO route. In the high concentration of metallic surface area, the
temperature difference in the catalyst bed was relatively lower, indicating the
concurrence of CH4 total oxidation and CH4 reforming where the combustion
heat was moderately absorbed by endothermic CH4 reforming for syngas
CATALYSIS REVIEWS 15

Figure 8. Results of systematic linear combination analysis in comparison to the ms signal for
observed oscillations at the beginning and the end of the catalytic fixed bed (upper part, x/l = 0;
lower part, x/l = 1). Reproduced with permission from ref.[101] Copyright 2011, American Chemical
Society.

production. Conversely, in the low concentration of metallic surface area, the


CH4 reforming reactions were consecutive to the CH4 combustion due to
insufficient metallic sites for simultaneously catalyzing both reactions, leading
to a more severe temperature gradient.
Rocha et al.[33] conducted in situ X-ray absorption near edge structure-
partial oxidation of methane (XANES-POM) experiment to examine the
changes in Pt oxidation changes along the reactor length and reaction dura­
tion. They found that metallic Pt0 phase was the key to yield the syngas and
this trend was more pronounced at elevated temperature. However, partial of
the metallic Pt0 particles underwent oxidation to form PtO2 at oxidative
environment, which negatively affected the syngas selectivity. This observation
was assigned to the co-occurrence of MPO and CH4 total oxidation inside the
catalyst bed. Such observation was in line with the findings of Velasco et al.,[102]
Grunwaldt et al.[103] and Mishra et al.[32] The obtained syngas selectivity was
apparently higher in sequential bed compared to single bed, accredited to the
16 T. J. SIANG ET AL.

facilitated CH4 activation owing to large amount of preserved metallic phase


under less competitive adsorption behavior of gas reactants, specifically in the
large amount of existing oxidizing agents.

4. State-of-the-art catalyst design


Modern advances in catalyst design and attempts at yielding synthetic
long chain hydrocarbons have revitalized interest in MPO. In this context,
numerous efforts have been devoted to effectually overcome the above­
mentioned difficulties in heat and mass transport for MPO system, oxida­
tion state changes in metal toward the MPO performance and catalyst
deactivation induced by carbon deposition on catalyst surface. Figure 9
demonstrates the brief overview about various strategic designs of MPO
catalyst in recent years. In this subsection, in-depth discussions of key

Figure 9. Schematic illustration of the specific strategics for MPO catalyst design.
CATALYSIS REVIEWS 17

findings with respect to metal-support interaction, surface acidity and


basicity, oxygen vacancies, perovskite structure and surface morphology
have been provided.

4.1. Metal-support interaction


Given the duration of metallic phase of a catalyst system is the key factor for
a long-term syngas production in MPO, enhancing the metal-support inter­
action is an effectual method to preserve the metallic active sites by minimiz­
ing the oscillation of metal oxidation state during reaction. In heterogeneous
catalysis, metal-support interaction is of fundamental importance as it is
responsible for the catalytic performance, selectivity and stability. A strong
metal-support interaction generally occurs where the support is reducible
playing crucial role to govern the catalytic activity and stability via the electron
transfer between metal-support interfaces or the formed intermediate
phases.[104] In fact, in the presence of metal-support interaction, the reactants
can migrate either from metal or support to the edge/boundary where the
reaction occurs, thus creating spillover effect to facilitate the reaction rate. In
addition, numerous MPO studies confirmed that the enhanced metal-support
interaction can largely suppress the deposition of carbonaceous species to
evade the encapsulation or separation of the metal particles from the support,
hence improving the catalyst lifespan [96, 105‒.106]
Selection or modification of support material is a common yet effective
approach to tune the extent of metal-support interaction for a catalyst system.
Spinel oxides, chemical formula of AB2O4, is a typical candidate for catalytic
heterogeneous system due to its excellent confinement effect toward doped
metal particles, of which the tetrahedral sites are preferentially occupied by the
trivalent cations whereas the both divalent and trivalent cations fill the octa­
hedral sites. Gil-Calvo et al.[107] synthesized NiAl2O4 spinel-based catalysts
with various Ni/Al molar ratios (from 0.13 to 0.50) and examined their
intrinsic activity in MPO. At optimized Ni/Al molar ratios around 0.25 to 0.30,
CH4 conversion of 54‒57% and H2/CO ratio at around 2.3 for 30 h on-stream
could be obtained, accredited to the enhanced metal-support interaction, of
which the higher specific activity of crystallites originated from Ni2+ cations
occupied at octahedral sites in the spinel-like aluminate despite the lower
metallic surface areas and larger crystallite sizes. Such advantageous support
structure toward metal-support interaction is also visibly observed in Ni/CeO2
catalyst.[108‒109] Kang et al.[105] synthesized the Ni-Fe2O3/Al2O3 oxygen car­
rier for examining the chemical looping MPO and MDR activity. The forma­
tion of Ni aluminate in the oxygen carrier, as the product of strong metal-
support interaction, was the key intermediate for inhibiting the carbon deposi­
tion during the chemical looping process. In fact, the existing Ni aluminate
18 T. J. SIANG ET AL.

Figure 10. Schematic diagram of CH4 interaction with pristine CeO2 surface and Ni-CeO2 surface
having surface defects with undercoordinated o atoms. Reproduced with permission from Ref.
[108]
Copyright 2015, American Chemical Society.

substantially enhanced the rate of CO production via carbon gasification,


leading to the insignificant amount of carbon deposit (0.135 molC/molsyngas)
located on the carrier surface.
An appropriate amount of Ni loading on CeO2 reportedly improved the
interstitial and surface defect by creating ‒O‒Ni‒O‒Ce super structure with
exposed exceedingly under-coordinated O atoms, which played an important
role in MPO activity, particular to the activation of CH4 molecules at low
temperature.[108,110] As shown in Figure 10, the equivalent symmetry of C‒H
bond (around 109.23°) for adsorbed CH4 on the under-coordinated O atoms
of ‒O‒Ni‒O‒Ce super structure could be tuned to nonequivalent form (>
109.23°), thus greatly facilitated the C‒H bond cleavage. This evolution
explains the excellent catalyst performance of the 7.5%Ni/CeO2, of which the
CH4 conversion of 98%, H2 and CO selectivity of 71% and 73%, respectively,
for 50 h on-stream without significant catalyst deactivation. Similarly, a notable
low-temperature CH4 activation over Pt/CeO2 during MPO was also observed
by Singha et al.,[109] attributed to the remarkable enhanced metal-support
interaction.
Addition of second metal forming a bimetallic catalyst system is also
a promising method to greatly enhance the metal-support interaction by
tuning surface state of the alloy particles.[96,106,111,112] Particularly, incorpora­
tion of secondary metal with higher oxygen affinity can efficiently reduce the
oxygen coverage of primary metal, thus prolong its preservation of metallic
phase during oxidation reaction. To investigate the relationship between
metal-support interaction and activation of reactants, Luo et al.[96] prepared
Pd/TiO2, Au/TiO2 and Au-Pd/TiO2 catalysts and assessed their catalytic
performances in MPO activity. The Pd/TiO2 and Au/TiO2 catalysts failed to
activate either the CH4 or O2 at 500‒650°C but conversely, Au-Pd/TiO2
catalyst successfully dissociated both CH4 and O2 to form H2 and CO. They
accredited this to the enhanced metal-support interaction via electron transfer
CATALYSIS REVIEWS 19

between metal particles (i.e., Au and Pd) and TiO2, enabling the Au particles to
improve the adsorption and scission of O2 whilst Pd particles facilitated the
CH4 dissociation and as a result, Au-Pd catalyzed MPO could be occurred at
low temperature. However, an opposite behavior was observed by Emamdoust
et al.[113] in their work of MPO over SiO2 supported Ni and NiCe catalysts.
From the results of temperature programmed surface reaction for CH4 (CH4-
TPSR) and O2 (O2-TPSR), they found that the temperature for CH4 and O2
activation increased by 100°C and 50°C, respectively, in the presence of ceria.
This behavior was attributed to the rapid reoxidation of metallic Ni0 to NiO
form of NiCe/SiO2 catalyst at low temperature due to its excellent redox
properties, which led to the requirement of higher temperature to retain the
metallic Ni0 phase during MPO reaction. Based on the findings of Li et al.,[106]
addition of Re to a Ru/Al2O3 catalyst not only strengthened the metal-support
interaction to preserve the metallic Ru0 phase during reaction but also surpris­
ingly altered the reaction mechanism. In situ diffuse reflectance infrared
Fourier transform (DRIFT) studies revealed that the primary role of Ru was
to reduce the Re element to a lower-valent ReOx species and the subsequent
reduced Re species preferentially produced the syngas through direct MPO
route by forming formate intermediate in short contact time.
Apart from modification of support structure and addition of secondary
metal, tailoring the nanoparticles size and shape of hosting metal is of crucial
to the MPO performance as it not only affect the extent of metal-support
interaction, most importantly it leads to different reaction mechanism
depending on metal oxidation states associated with nanoparticles size [114–
116,]. In the study about size effect of Rh nanoparticles on syngas selectivity in
MPO activity, Kondratenko et al.[114] reported that the MPO activity
decreased by 20% with a rise in Rh nanoparticle size from 1.1 to 5.5 nm.
They attributed this behavior to the kinetic interactions of CH4, O2 and CO2
with Rh catalysts associated with particle size. Based on the microkinetic
evaluation of transient experiments, increased size of Rh nanoparticle led to
a drop in rate constants for all reactants and hence, affecting both primary
(CH4 total oxidation) and secondary (CH4 reforming) routes in MPO system.
The enhanced metal-support interaction reportedly not only facilitates the
activation of reactants but also effectively hinder the formation of carbonac­
eous species on catalyst surface. In the study of Rh supported hydroxyapatite
catalysts in MPO evaluation, the RhOx in small particles possessed a great
carbon resistance owing to strong metal-support interactions (where the
detected carbon amount was lesser than 1 wt%), which governed the long-
term MPO test in terms of CH4 conversion of 86% for 30 h.[117]
A further insight into the elementary reaction steps in association with
metal nanoparticle size has been recently revealed by computational works.
Based on the DFT calculation, Guo et al.[78] found that the size of Rh particle
has considerable influence on the oxidation state and thus leading to two
20 T. J. SIANG ET AL.

different mechanisms for syngas formation, known as CH4 dissociation and


CHx species oxidation. The Rh catalysts with low coordination number
(smaller Rh cluster or single Rh atom) reportedly possessed higher resistance
toward carbon deposition and oxidation. This was due to that in CH4 dis­
sociation route, the dissociated CHx* and C* species were relatively active,
requiring abundant active sites and generally adsorbed in several hollow site of
Rh. Conversely, CHx species oxidation pathway required less active sites as the
generated CH2O* species was a relatively saturated species, which usually
adsorbed on bridge or top sites of Rh particles. Thus, small-sized Rh catalysts
with low coordination number contributed to the CHx* oxidation process and
suppress the CH4* dissociation except the first C–H bond scission step, as
shown in Figure 11.

4.2. Surface acidity and basicity


The surface acidity/basicity of catalyst has been reported to act as an important
role on the catalytic performance for reforming of hydrocarbons.[118–120] The
tuning of acid-basic properties of catalyst can be contributed by the

Figure 11. The calculated pathways for MPO (δE/EA) on the small-sized RH catalyst model. red
colored path represents the optimum reaction route. Reproduced with permission from Ref.[78]
Copyright 2020, Elsevier.
CATALYSIS REVIEWS 21

combination of support, active metal and promoter. According to the


literature,[121] an increase in acidity strength favorably enhances the CH4
cracking and polymerization of carbon-containing compounds, which con­
tributes to the catalyst deactivation induced by carbon deposition. In this
context, basicity strength plays a crucial role for carbon gasification by facil­
itating the adsorption and activation of oxidizing agents.[93,120] The effect of
carbon deposition is insignificant on reforming activity if rate of carbon
gasification is equal to or superior to that of carbon formation.[35]
In the investigation about the basicity of Ni-based catalysts (i.e., Ni/Al2O3,
Ni/MgO/Al2O3, Ni/MgAl2O4 and Ni/Sorbacid) on syngas ratio and carbon
deposition in MPO reaction, Özdemir et al.[122] revealed that an increase in
basicity enormously reduced the H2/CO ratios (decrement around 73.8‒
82.6%) owing to the promotion in both reverse Boudouard reaction and
reverse water gas shift (RWGS) during MPO reaction at 500 to 600°C.
Additionally, increasing basicity via a rise in Mg content visibly suppressed
the deposition of graphitic carbon, retaining the reactivity of carbon gasifica­
tion at lower temperature with oxidant. Song et al.[123] prepared Ni/ZrO2
catalysts with employment of different NaOH concentrations (from 5 to
40 wt%) during the synthesis and examined them in MPO to determine the
correlation between basicity strength and the catalytic performance.
Interestingly, they found that low concentration of NaOH promoted the
development of tetragonal ZrO2 during synthesis process, resisting the carbon
formation due to less surface acidic sites. Such a difference in ZrO2 structure
significantly affect the CH4 reaction rate and the amount of coke formation,
which resulted in a very stable CO selectivity of 86% lasted for 150 h. However,
a higher concentration of NaOH resulted in the transformation of tetragonal
ZrO2 to monoclinic ZrO2 crystallization which deteriorated the MPO activity.
This observation indicates that an overly strong basic properties of a catalyst
could adversely affect its catalytic performance in the MPO activity.
López-Fonseca and the coworkers studied the effect of MgO/NiAl2O4 on
the MPO performance by tuning the surface properties of catalysts.[118] The
rising MgO addition drastically altered the strength and density of surface
basic sites of catalyst, leading to an excellent resistance to the carbon deposit
and metal sintering. Specifically, MgO/NiAl2O4 possessed weak basic sites
with well dispersed Mg species was optimum catalyst for MPO activity. They
reported that the extent of basicity strength had a considerable influence on
the MPO mechanism with respect to CO or CO2 formation. A highly stable
and strong basic sites were predominantly consisted of O2- anions. The surface
deposited carbon normally reacts with two type of oxygen species, namely,
mobile oxygen species and chemisorbed O2- species. The carbon reacts with
the mobile oxygen species to form CO, whilst CO2 is the product of reaction
between carbon and chemisorbed O2- species. However, as seen in Figure 12,
the observed CO/CO2 ratio and H2 yield substantially declined with rising
22 T. J. SIANG ET AL.

Figure 12. Correlation between basicity strength of MgO/NiAl2O4 catalysts and mpo activity in
terms of CO/CO2 ratio and H2 yield. Reproduced with permission from Ref.[118] Copyright 2016,
Elsevier.

strong basic sites density. They assigned this observation to the increased
quantity of chemisorbed O2- species at extremely strong basic sites density
and as a result, deterioration on MPO activity. Similarly, Ding et al.[124]
reported the advantages of basicity by introduction of Yb dopant to Ni/Al2
O3 catalyst for MPO activity. The addition of Yb not only enhanced the
dispersibility of Ni particles but also facilitated the CO2 adsorption on catalyst
surface, thus improving the carbon balance and hindering the carbon deposi­
tion. As a result, the Ni-Yb/Al2O3 catalyst reportedly exhibited CH4 conver­
sion of 98% and H2 selectivity of 83%, respectively, for a 170 h-on-stream
MPO reaction.

4.3. Oxygen vacancies


Oxygen vacancy is a paradigm depiction of intrinsic defects on reducible
catalyst surface, which is generally developed during synthesis process after
exposes to a reductive environment and elevated temperature.[38,58] The dif­
fusivity of oxygen vacancies largely affects the physicochemical properties of
a catalyst, which determines its catalytic performance during reaction. In fact,
reducible supports with oxygen vacancies are able to store oxygen ions at
oxidizing atmosphere and release the subsequent oxygen ions for reaction at
reducing surrounding, thus regenerating the oxygen vacancies via the redox
CATALYSIS REVIEWS 23

cycles.[125] This can be attributed to the high oxygen storage capacity (OSC) of
reducible materials improves the mobility of oxygen atoms from crystal lattice
to surface, which facilitates the corresponding cyclic process.[14] In MPO, the
existence of oxygen vacancies acts as active sites for dissociation of oxidants,
hence enhancing the pool of labile oxygen ions for surface reaction on
catalyst.[38] Additionally, it has been established that the tuning of metal
state for catalytic activity can be achieved by shifting the Fermi level with
appropriate combination of reducible oxide support and metal
atoms.[30,126,127] As a result, the formation energy of oxygen vacancies could
be altered and subsequently affect the reactivity of the oxygen atoms from the
surface lattice of a reducible oxide, which significantly influenced the reaction
kinetic regimes for C‒H bond activation in CH4 molecules.[56,72,128]
Amongst the reducible oxide supports, CeO2 is widely recognized as the
most suitable representative for evaluating effect of oxygen vacancies on
a catalysis activity owing to its high OSC and great diffusivity of releasing
and restoring oxygen.[14,127] On this basis, several studies about CeO2 support
modification including CeO2 mixed oxides,[129–131] CeO2 with different topo­
logical structure [132,133] and the corresponding in situ evaluations [134,135] have
been developed. Toscani et al.[134] conducted an in situ dispersive X-ray
absorption spectroscopy (DXAS) study for MPO over CuO–NiO/CeO2
–ZrO2 catalyst in order to determine the relationship between the active Ni
and redox behavior of CeO2–ZrO2. They reported that such a redox behavior
of Ce associated with formation of oxygen vacancies could create a H2 spil­
lover effect on catalyst surface leading to a lower reduction temperature for
catalyst activation, retaining the preservation of metallic Ni0 phase for long-
term MPO reaction. A similar behavior was also observed in their recent
in situ DXAS evaluations of MPO over NiO/CeO2-Sm2O3
nanocomposites,[135] which further highlights the significance of tuning oxy­
gen vacancy concentration in CeO2-supported Ni catalysts to optimize the
catalytic activity. A detailed mechanism of a catalyst support with high con­
centration of oxygen vacancies toward preservation of metallic phase and
promotion of reforming activity has been recently proposed by Yentekakis
et al.[136] for Rh/alumina-ceria-zirconia. As shown in Figure 13a, the adsorbed
oxygen atoms on oxygen vacancies could produce an effective double layer
(Oδ-, δ+) via O2- back-spillover effect on the Rh particles, leading to changes in
chemisorptive properties and Rh work function. Notably, Oδ- layer not only
enhanced the Rh work function to strengthen the chemisorptive bond of
electron donor (CHx species) but also weakening that of electron acceptor
(atomic O from oxidants). As a result, the Rh‒O bonds in RhOx state were
weakened and destabilized, thus readily transforming into metallic Rh0 form
during reaction. Similar findings in the presence of oxygen vacancy-rich
support are also suggested by works of Das et al.[133] and Al-Fatesh et al.[137]
24 T. J. SIANG ET AL.

Figure 13. (a) Model of the bifunctional support-mediated promotion by supports with high
oxygen ion lability. Reproduced with permission from Ref.[136] Copyright 2018, Elsevier. (b)
Proposed reaction processes of chemical looping mpo over Ni-doped WO3 oxygen carriers.
Reproduced with permission from Ref.[86] Copyright 2017, American Chemical Society.

Apart from the “effective double layer model” to preserve the metallic phase
for a catalyst system, the stored oxygen atoms on oxygen vacancies reportedly
contribute to the carbon removal via gasification and hence suppression on the
carbonaceous deposition on catalyst surface. Although the redox characteristic
of reducible oxygen carriers is of interest for MPO, recent studies revealed that
the regeneration of reduced oxygen carriers by oxidants favored the CO2
formation at the expense of CO selectivity.[86,138] In this context, exploration
of oxygen carriers to provide optimum lattice oxygen to selectively produce
syngas in chemical looping MPO is essential.[74,79] Chen et al.[86] designed
WO3-based oxygen carriers as the primary source of lattice oxygen and further
enhanced the CH4 conversion ability and availability of lattice oxygen by
introducing NiO nanoclusters. The characterization results revealed that the
embedment of Ni cations into the WO3 lattice structure largely reduced the
tungsten‒oxygen bond strength, thus increased the lattice oxygen availability.
As seen in Figure 13b, NiOx species efficiently dissociated CH4 molecules to
form CHx species and subsequently reacted with O2- species from WO3-based
CATALYSIS REVIEWS 25

oxygen carriers to produce syngas and oxygen vacancies during CH4 reduction
stage. In O2 oxidation stage, the metallic Ni0 phase was oxidized to NiOx
species and active lattice oxygens were restored on oxygen vacancies in WO3-
based oxygen carriers for the next redox cycle. The drastic change in reactivity
of oxygen carriers to chemical looping MPO with an addition of a low quantity
of metal dopant on Fe2O3 oxygen carriers is also reported by Qin et al.[139]
With the Cu dopant, they found that the CH4 conversion was 4‒5 fold high
than its counterparts under same reaction conditions due to the facilitation of
oxygen vacancies formation in Fe2O3 oxygen carriers lowering barrier of CH4
activation energy. Shen et al.[140] investigated the effect of regeneration dura­
tion on the syngas selectivity over BaFe3Al9O19 hexaaluminate (BF3A) as the
oxygen carries in chemical looping MPO. They found that the syngas selec­
tivity was highly reliant on the variation in oxidation state of Fe in association
with different regeneration periods where the concentration of oxygen vacan­
cies was determined by the oxidation state of Fe. A decrease in regeneration
duration (from 15 min to 4.2 min) resulted in a high tendency of Fe2+ and Fe0
formation, creating more oxygen vacancies for MPO and MDR to produce
syngas.

4.4. Perovskite
Given the merits of oxygen vacancy toward oxidation reactions, perovskite
oxides (ABO3 where A = rare-earth element; B = 3d transition metals),
mixtures of semiconducting and oxygen carrier materials, have been exten­
sively studied in reforming reactions, in particular to chemical looping
approach.[141,142] In fact, the concentration of oxygen vacancy in perovskite
structure is highly related to the total charge of A and B cations.[142] The
reduction degree of transient metal cations and high oxygen storage capa­
city are the two attractive features of perovskites make it applicable in
chemical looping MPO as both catalyst and oxygen carrier.[71,72,143]
Interestingly, the partial or total combustion reportedly depends on the
nature of the oxide ion involved and type of oxygen interaction with the
perovskite surface. Thus, the knowledge of surface oxygen binding energies
and the corresponding kinetics is of significance to catalysis. In this context,
Mihai et al.[144] have conducted a kinetic study for CH4 total combustion,
partial oxidation and CH4 decomposition on LaFeO3 perovskite with the
manipulation of the O content and crystal size during the chemical looping
MPO process. They found that the syngas selectivity was highly reliant on
the quantity of O available and withdrew from the LaFeO3 perovskite
structure and its crystal size. Kinetic determining step in the CH4 conver­
sion was ascertained as the surface reaction step related to Fe–O bond that
substantially depending on Fe–O bond strength, which determined the
LaFeO3 crystal size. The larger LaFeO3 crystal size reportedly possessed
26 T. J. SIANG ET AL.

a lower Fe–O bond strength, which exhibited a greater syngas selectivity


and higher amount of O removal. Factually, the changes in the amount of
O removal was the key factor which apparently influences the nature of
active sites. As seen in Figure 14, three active sites have been determined
where Fe sites (i) highly binding surface O (FeI) resulted in formation of
CO2; (ii) moderately binding surface O (FeII) led to formation of H2 and
CO; and (iii) highly binding oxygen vacancy sites (FeIII) resulted in forma­
tion of carbon. Based on these findings, the propensity of partial or total
oxidation of CH4 was closely associated with surface O coverage of LaFeO3
perovskite and a high selectively syngas production through MPO was
probable at moderate O coverage by adjusting the degree of oxidation for
reduced perovskite in chemical looping process. The three stages of Fe sites
coordinated with different surface O content are further verified by the
works of Yang et al.[145] combining the microkinetic analysis and DFT
computations. Their results revealed that the overall energy barrier of
total oxidation could be considerably increased in the presence of oxygen
vacancy sites, which led to a reduction in CO2 selectivity. Therefore, Fe
coordination environment in association with the surface oxygen vacancy
concentration was the most important determinant of catalyst selectivity.

Figure 14. Schematic diagram of the total combustion, partial oxidation and methane decom­
position on different active sites of LaFeO3 perovskite. (o)s ‒ surface oxygen species, □ – oxygen
vacancy site, l – number of total active sites on the surface. Reproduced with permission from
Ref.[144] Copyright 2012, Elsevier.
CATALYSIS REVIEWS 27

Since the amount of O removal participating oxidation reactions is signifi­


cantly related to the extent of Fe–O bond energy, the tailoring of elemental
composition toward the electronic structure in perovskite structure is of
interest. Jiang et al.[146] prepared a series of LnFeO3 (where Ln = Pr, Gd, Sm
and La) and reported that perovskites with different A-site metals exhibiting
various geometric tilting degrees of the FeO6 octahedra. Based on the DFT
results, a severe geometric tilting of FeO6 octahedra could be induced by A-site
cations having a smaller radius, led to a weakened Fe–O orbital hybridization;
thus, the Fe–O covalency reduced in the order, LaFeO3 > PrFeO3 > SmFeO3
> GdFeO3. This reflected the drop in CH4 conversion from LaFeO3 (65%) to
GdFeO3 (46%), which corresponded to the Fe–O covalency. Additionally, the
syngas yield trended in the same direction as the decreased oxygen mobility
and surface oxygen activity, suggesting that the Fe–O covalency was the crucial
factor in influencing the syngas yield. In fact, the weakened Fe–O covalency
indicating a rising charge-transfer energy, which caused a higher energy
barrier in both oxygen vacancy formation and the oxygen mobility.
Consequently, the attenuating oxygen migration resulted in a sluggish diffu­
sion rate of lattice oxygen in perovskite structure and thus adversely affected
the participation of surface O content in MPO activity. In order to investigate
the oxygen mobility effect on catalytic activity of perovskites in chemical
looping MPO, Chang et al.[143] synthesized La1-xSrxFeO3-δ (x = 0, 0.2 and
0.5) and La0.5Sr0.5Fe1-xCoxO3-δ (x = 0 and 0.5) perovskites as well as examined
them in both experimental and DFT assessments. In perovskites, the oxygen
mobility determines the concentrations of both oxygen vacancy and surface
oxygen, which has profound effect on the product selectivity. For the La0.5Sr0.5
Fe1-xCoxO3-δ perovskites, the oxygen amount participated in CH4 total oxida­
tion were two-fold higher than that of La1-xSrxFeO3-δ. They noticed that the
quantity of active oxygen grew with the increasing A-site replacement (Sr2+ for
La3+) and more pronounced to the B-site substitution (Co3+ for Fe3+), indi­
cative of the prompt replenishment of lattice oxygen resulting in high surface
oxygen density. In addition, the Co‒O bonds were much weaker than Fe‒O
bonds and as a result, a rapid oxygen diffusion and ion mobility in the bulk,
explaining the inclination of La0.5Sr0.5Fe1-xCoxO3-δ toward total oxidation
products. Thus, La0.8Sr0.2FeO3 perovskites with comparatively low oxygen
migration were suggested as prominent catalyst to yield syngas through partial
oxidation process.
To completely avail the characteristic of perovskite in chemical loop­
ing process, Shen et al.[140] coupled the CO2 splitting with MPO over
LaFeCo1-xO3 with various Fe:Co ratios (3:1, 1:1 and 1:3). As seen in
Figure 15a, the oxygen carriers were reduced at CH4 stage and produced
different gaseous products depending on the reduction degree of
LaFeCo1-xO3 perovskites. Similar to the reports of Chang et al.[143] and
Mihai et al.,[144] the readily and rapid release of adsorbed oxygen on
28 T. J. SIANG ET AL.

Figure 15. (a) Proposed reaction mechanism of CH4 partial oxidation coupled with CO2 splitting over
LaFexCo1-xO3 for syngas production. Reproduced with permission from Ref.[71] Copyright 2020, Elsevier.
(b) Scheme of possible paths for lattice oxygen mobility in La1-xCexFeO3 (x = 0, 0.5, 1) redox catalysts. (a)
Calculated (dft+u) oxygen vacancy formation energy (δef,vac). (b) Calculated (dft+u) oxygen migration
energy (em). (c) The bulk diffusion coefficient (dchem) of oxygen in redox catalysts at 850°C determined by
analyzing the conductivity relaxation kinetics after an abrupt change in the atmosphere from 1 vol.%
CH4/CO2 to 10 vol.% CH4/CO2 in electrical conductivity relaxation measurement. (d) The arrhenius plots
of the dchem over a temperature of 650‒850°C measured by switching between two CO/CO2 gas streams
with different po2 in electrical conductivity relaxation measurement. (e) Correlation between syngas
yield and degree of FeO6 octahedral distortion. Reproduced with permission from Ref.[72] Copyright
2020, American Chemical Society.
CATALYSIS REVIEWS 29

oxygen carriers was responsible for total oxidation at beginning CH4


stage owing to the sufficient O content. As the adsorbed oxygen con­
sumed, oxygen vacancy sites were produced which enabled the migration
of lattice oxygen to the catalyst surface leading to MPO for syngas
generation. An increase in Fe addition was beneficial for the syngas
selectivity where the H2 and CO selectivity was 99% and 53%, respec­
tively, over LaFeCo 1-x O 3 (Fe:Co = 3:1). This was attributed to the
incorporation of Fe improved the binding capability between Co‒O
bonds and thus suppressing the release rate for oxygen in perovskite.
Although CO2 was being generated as a result of total oxidation at first
CH4 stage, the restoration of lattice oxygen could only be done during
subsequent CO2 stage based on the thermodynamic limitations.[147] The
use of CO2 as regeneration feedstock reportedly prevented the deep
oxidation of the metals in perovskites due to the suppression on the
formation of electrophilic O2- or O− species. At CO2 oxidation stage,
they found that the O atoms from CO2 transported to the reduced form
of Co and Fe metals by La2O3. Subsequently, the Co and Fe alloys
underwent the regeneration and redistribution of structures and thus
recovered to their perovskite forms and ready for next CH4 cycle. In
order to gain an insight into the lattice distortion toward the oxygen
mobility and surface oxygen exchange capability of perovskite for MPO
coupled with CO2 splitting, Zhang et al.[72] conducted the DFT study of
the LaFeO3 perovskites with various La:Ce composition and different
probable oxygen migration paths (Figure 15b). The computed formation
energy of oxygen vacancy (∆Ef,vac) for La0.5Ce0.5FeO3 was the lowest
(3.05 eV) compared to that of LaFeO3 (4.73 eV) and CeFeO3 (4.8 eV)
specimens. Due to the oxygen diffusion by hopping process in perovs­
kite-type metal oxides, increasing the FeO6 octahedral distortion might
produce more oxygen vacancy sites and thus improvement of oxygen
transfer in the lattice.[148] Such great FeO6 distortion in La0.5Ce0.5FeO3
also contributed to the lowest oxygen ion transport energy (Em), regard­
less the oxygen migration paths. Additionally, the bulk diffusion coeffi­
cient (D chem ) of La 0.5 Ce 0.5 FeO 3 was ten-fold higher than others,
indicating its outstanding bulk oxygen mobility as a result from FeO6
octahedral distortion. This postulation was further corroborated by the
lowest apparent activation energy of lattice oxygen transfer for La0.5Ce0.5
FeO3 in the Arrhenius plots of the Dchem. Furthermore, CH4-TPR and
CH4 pulses findings suggested that FeO6 octahedral distortion improved
surface reactivity of lattice oxygen and CH4 activation capability of La1-x
CexFeO3 perovskites. Hence, the enhanced bulk oxygen transport due to
FeO6 distortion in La0.5Ce0.5FeO3 was the key to explain its excellent
catalytic performance, whose syngas yield was 3‒8 times larger than its
counterparts.
30 T. J. SIANG ET AL.

4.5. Surface morphology

Since the MPO is a fast exothermic reaction, which readily generates hot
spot in catalyst bed and as a result, catalyst deactivation due to collapse
of catalyst structure and agglomeration of active metal is always an
unavoidable setback. In this context, modification of catalyst morphol­
ogy has emerged as an effective approach to facilitate the external and
internal gas diffusion of reactants and products in the catalyst bed in
order to minimize the effect of hot spot formation for achieving high
syngas yield.[67,149] Li et al.[150] investigated the effect of mesoporous
material for the MPO performance by preparing Co-Ni-MCM-41 cata­
lysts. The Co-Ni-MCM-41 exhibited 88% CH4 conversion and 87% H2
selectivity, accredited to the excellent anchoring effect of mesoporous
structure reducing the metal sinterization. In addition, the mesoporous
channels in MCM-41 hindered the nucleation of carbonaceous species to
prevent the catalyst deactivation through encapsulation. In the study of
chemical looping MPO, Liu et al.[151] also reported the advantages of
mesoporous Fe2O3@SBA-15 nanostructure in the catalysis activity by
attaining a near 100% CO selectivity at temperature as low as 750°C.
They attributed this superior CO selectivity to the mesoporous nanos­
tructure of SBA-15 facilitated the reaction rate on surface site, of which
low-coordinated oxygen atoms enormously enhanced the cleavage of Fe‒
O bond and CO formation compared to conventional Fe2O3 micropar­
ticle, as illustrated in Figure 16a. In addition, the HRTEM images of Fe2
O 3 @SBA-15 (Figure 16b) before and after redox cycles showed no
noticeable morphological discrepancy, suggesting the highly stable struc­
ture of Fe2O3@SBA-15 at high temperature.
Apart from modified mesoporous siliceous materials, Ding and his cow­
orkers [152,153] employed 3D honeycomb-like silica as substrate to prepare
mesoporous metal oxide (viz., Yb2O3, CeO2, La2O3 and ZrO2) supported Ni
catalysts and assessed their catalytic performances for MPO. Such morpholo­
gical structure with great confinement effect provided a strong metal-support
interaction and highly dispersed active sites, governing the CH4 conversion of
90 to 92% and H2/CO ratio of 2 to 3, particular to mesoporous CeO2 and ZrO2
supported Ni catalysts. Based on their DFT computation, they found that the
defect sites of CeO2 and ZrO2 presenting the surface hydroxyl groups could be
substantially restored via reoxidation CeO2 and ZrO2 (due to redox cycles) and
H2O activation, thus promoting the H2 generation. The beneficial effects of
modulating the reducible support structure are also suggested by Singha
et al.[154] in their MPO investigation over 2Pt-CeO2 nanoporous (2Pt-
CeO2NP) spheres. In brief, the high Pt dispersion was readily obtained in the
presence of bimodal CeO2 nanoporous spheres through controlled Pt deposi­
tion technique. Contributed to its bimodal pore structure, an average Pt
CATALYSIS REVIEWS 31

Figure 16. Schematic diagram of (a) chemical looping of mpo process and (b) structure and co
selectivity in conventional oxygen carrier vs Fe2O3@SBA-15 oxygen carrier. (b) HRTEM images of (a)
fresh Fe2O3@SBA-15 oxygen carrier and (b) spent Fe2O3@SBA-15 oxygen carrier after 75 redox
cycles. Reproduced with permission from Ref.[151] Copyright 2019, Springer Nature.

nanoparticles size of 1.27 nm supported on nanoporous spheres of 150 nm


with 5 to 15 nm CeO2 particles could be achieved. The combination effect of
bimodal nature and high oxygen storage capacity resulted in a lower activation
temperature of CH4 at 350°C meanwhile anchoring the Pt nanoparticles
resided at smaller pores, hence preventing sinterization and fluctuation of
32 T. J. SIANG ET AL.

MPO activity due to rapid molecular transportation. The 2Pt-CeO2NP report­


edly exhibited a high CO selectivity and a stable H2/CO ratio around 1.95 at
800°C with negligible activity loss induced by carbon deposit.
Encapsulation of metal nanoparticles with a porous shell (i.e., alumina or
silica), namely, core-shell structured nanocatalysts have been extensively scru­
tinized in reforming studies owing to its great resistance toward sintering and
carbon deposition. Particularly, the metal nanoparticles isolated inside the
inert shells have a relatively uniform environment around the core surface
which governs the reaction rate for product formation [155‒.156] Li et al.[157]
prepared Ni nanoparticles of narrow size distribution (6‒45 nm) encapsulated
inside meso- and microporous silica, namely, Ni@porous silica core–shell
nanocatalysts via in situ NiO reduction. The notable discrepancy in catalytic
performance and stability of these core–shell nanocatalysts in MPO reaction
was highly related to the variation in Ni core size, shell porosity and degree of
core–shell interaction depending on microcapsular cavity structure. The opti­
mized Ni-350@meso-SiO2 catalyst (calcined at 350°C) with Ni core size of
6 nm and SiO2 shell of 3 to 4 nm mesopores presented a remarkable CH4
conversion of 93% and H2 selectivity of 92% at 750°C. Similar findings are also
suggested by Ding et al.[155] and Dou et al.[158] The results demonstrated that
SiO2 shell played an essential role in suppression on carbon deposition, by
blocking the active edge and corner atoms to prevent the nucleation of carbon
atoms, which accounts for excellent catalytic stability for long-term MPO
reaction. As abovementioned in the previous section, perovskite appears as
a promising redox catalyst owing to its astonishing restoration in catalyst
structure during cycles of reduction/oxidation processes in chemical looping
MPO. Interestingly, this feature has been further highlighted when coupled
with the core-shell structure concept. In a chemical looping MPO investiga­
tion, Shafiefarhood et al.[159] prepared core–shell Fe2O3@LaxSr1-xFeO3 redox
catalyst, Fe2O3-LaxSr1-xFeO3 perovskite and composite redox catalysts using
inert (Al2O3 and MgAl2O4) and ionic conductive (ZrO2 and Y2O3) materials
for assessing their catalytic activity. The core–shell Fe2O3@LaxSr1-xFeO3 was
10 to 200 times much active than that of inert and ionic conductive materials
in MPO reaction, accredited to the small Fe2O3 core size (50 nm) facilitating
lattice oxygen mobility and electron conduction through the mixed conductive
LaxSr1-xFeO3 shell (10 nm). For the core–shell redox catalyst, the encapsula­
tion of Fe2O3 with a stable perovskite shell lessened the metallic Fe on the
surface under a reducing environment, which resulted in a lower carbon
formation compared to composite redox catalysts. This postulation is further
corroborated by Neal et al.[160] in their investigation of dynamic MPO over
core–shell Fe2O3@LaxSr1-xFeO3 in the absence of gaseous oxygen.
In recent years, fibrous catalysts emerge as rising stars in MPO process due
to its high void volume of the fibrous structure improving the accessibility of
gaseous reactants to the active sites at high gas space velocities to achieve high
CATALYSIS REVIEWS 33

syngas yields without significant carbon deposition.[53,67,149,161,162] To deter­


mine the correlation of surface morphology and catalytic performance, Ma
et al.[149] synthesized Ni/Al2O3 with different morphological structure,
namely, fibrous and monolithic configuration. In comparison with spherical
catalyst, the Ni particles were well dispersed on the surface of fibrous catalyst,
which created a relatively stronger metal-support interaction. Apart from that,
the fibrous Ni/Al2O3 exhibited remarkable syngas yield, which was around 21
times higher than that of monolithic Ni/Al2O3. This was ascribed to the
superior mass transfer of reactants and product in the presence of fibrous
structure. By combining the concepts of redox characteristic (supply O atom
to suppress carbon deposit) and surface morphology (enhance accessibility of
reactants to active sites) in MPO, Zhao et al.[161] prepared Ni-CeAlO3-Al2O3
/FeCrAl-fiber catalyst by using layered double hydroxides approach. The
catalyst not only achieved a remarkable CH4 conversion (86%) and H2/CO
selectivity (96.1/91.5%) at high gas hourly space velocity (100 L g−1 h−1) for
350 h reaction run but also greatly hindered the deposition of carbonaceous
species in the presence of “CeAlO3-CeO2” redox cycle. In our previous
works,[53,162] by coupling the features of oxygen vacancies and dendritic
fibrous morphology, KCC-1 supported noble metals catalysts possessing
unique hierarchical nanopores and three-dimensional (3D) center-radial
nanochannels (Figures 17 a and d) were successfully synthesized for MPO
reaction. The results revealed that the coexistence of oxygen vacancies and
fibrous morphology largely enhanced the metal-support interaction led to
a reduction in the apparent activation energy of CH4 and O2- species, thus
accelerating the surface reaction for syngas production (Figure 17e). In addi­
tion, the cyclic adsorption/desorption of O2- species, well known as Mars–van
Krevelen mechanism, on the high concentration of oxygen vacancies effi­
ciently hindered the carbon deposition via gasification process. In the evalua­
tion of catalytic stability and regenerability of KCC-1 supported noble metals
catalysts,[162] CH4 conversion of 55‒82% and H2/CO ratio around 1.52‒2
could be readily achieved for 16 h-on-stream MPO reaction despite the 3 h
in situ oxidation process with O2 after the first 8 h reaction run, suggesting its
excellent resistance to the reoxidation of active metal in the presence of KCC-1
with high oxygen ion lability, which was dovetailed with the findings of
Yentekakis et al.[136]
Recent advancements have been described as a ground-breaking develop­
ment of MPO catalysts based on detailed material characterization with
employment of theoretical simulation and real-time reaction monitoring at
the atomic and molecular levels. Intensive efforts have been put on the
emphasis of fundamental understanding toward interactions among active
metals, supports, gaseous reactants and operating conditions in order to
develop better catalysts systems. As highlighted in Table 2, these advances of
34 T. J. SIANG ET AL.

Figure 17. (a and b) FESEM images and (c and d) OF fresh fibrous KCC-1 support at different
magnifications. (b) pictorial depiction of MPO mechanistic over fibrous M/KCC-1 (m = ru, pd, rh)
catalysts. Reproduced with permission from Ref. [53] Copyright 2020, Elsevier.

new paradigm development to yield efficient MPO catalysts are mostly based
on new insights into synergistic correlations between catalyst components,
catalysis activity and product selectivity.

5. Kinetic study of MPO


Kinetic studies are conducted to provide an empirical reaction rate expression
based on a theoretical reaction mechanism that best fits the experimental data
in order to describe the reaction rate and determine the chemical process.[94]
Fundamentally understanding of the inherent mechanistic MPO routes and
Table 2. Highlights of recent advancements in development of MPO catalysts.
Reaction conditions
CH4 GHSVb CH4 H2/
Temp.a :O2 (L g−1 conversion CO
−1
Catalyst (oC) ratio h ) (%) ratio Key findings Ref.
[96]
Au-Pd/TiO2 600 2:1 40,000c 25.9 3.51 Synergy effects between Au and Pd enhanced the metal-support interaction, where Au modulated the behavior of
oxygen and Pd was able to activate the C‒H bonds in CH4 molecules at low-temperature region
[107]
NiAl2O4 650 2:1 384 72.0 2.62 The remarkable performance at low-temperature region was due to strong metal-support interaction as a result of
crystallites derived from Ni2+ cations to metallic phase at octahedral sites in the oxide lattice
[108]
Ni/CeO2 800 2:1 50 98.4 - The enhanced catalytic activity was attributed to the enhancement of metal-support interaction owing to the
formation of ‒O-Ni‒O‒Ce super structure, which led to interstitial point defects and surface defects facilitating CH4
activation
[109]
Pt-CeO2 350 2:1 80 23.2 1.85 The nanosized Pt and Ce particles created a strong metal-support interaction which was the key factor for CH4
activation at low temperature around 350°C
[122]
Ni/MgO 800 2:1 157.5 85.3 2.09 The increase of basicity promoted interaction of surface-adsorbed carbon with oxygen, accelerating the carbon
gasification and achieved the long-term stability
Ni-Yb/Al2O3 700 2:1 350 90.4 - The Yb promotion not only reduced the Lewis acidity of Al2O3 support but also improved the CO2 adsorption and thus, [124]
enhancing resistance toward carbon deposition
Ni0.5WOx/Al2O3 800 1.33:1 28 57.6 2.08 Introduction of Ni2+ cations into WO3 bulk could weaken the tungsten‒oxygen bond strength, which improved the [86]
lattice oxygen diffusion and surface reaction rates of chemical looping MPO
[140]
BaFe3Al9O19 900 1:1 15 90.1 1.97 In chemical looping MPO, a shorter regeneration duration could largely improve syngas yield. This was due to the
reduction of Fe3+ in Al sites that active for CH4 combustion whereas increasing reduced Fe2+ to Fe0 accounting for
MPO
Ru/CeO2-ZrO2 800 2:1 50 96.9 - The oxygen storage capacity of CeO2-ZrO2 solid solution provided large amount of surface adsorbed oxygen for surface [163]
reactions and strengthened the metal-support interaction with nanosized Ru particles.
[143]
La0.8Sr0.2FeO3 900 1:1.39 432d 95.1 2.09 The superior activity was attributed to the lower oxygen mobility, which highly related to the bulk oxygen
concentration and rate of lattice oxygen diffusion. La0.8Sr0.2FeO3 with lower oxygen mobility favored to MPO
reaction
LaFeO3 900 2:1 960d 96.8 2.02 The surface oxygen vacancy concentration associated with Fe coordination was the key factor governing the syngas [145]
selectivity, of which CH4 oxidation could be transformed from total oxidation (O-excessive surface) to partial
CATALYSIS REVIEWS

oxidation (O-deficient surface)


(Continued)
35
36

Table 2. (Continued).
Reaction conditions
T. J. SIANG ET AL.

CH4 GHSVb CH4 H2/


Temp.a :O2 (L g−1 conversion CO
Catalyst (oC) ratio h−1) (%) ratio Key findings Ref.
[146]
LaFeO3 850 1:2 36 65.1 1.55 The syngas yield was reportedly controlled by the Fe‒O covalency associated with A-site cation in perovskite. The
larger La cations reduced the energies of both oxygen vacancies formation and oxygen anions migration, thus
facilitating oxygen transfer and the surface reaction
Ni@meso-SiO2 750 2:1 72 93.2 - The excellent CH4 conversion was assigned to the degree of core-shell interaction determined by Ni cores size and the [157]
porosity of SiO2 shells. Such microcapsular structure governed the catalytic activity and stability for 500 min
[155]
Ni/ZrO2@SiO2 750 2:1 50 98.9 - The higher oxygen storage and release capacity of Ni/ZrO2@SiO2 effectively prolonged the catalyst lifespan by
suppressing carbon formation. Specifically, silica shell possessed a great resistance toward carbon deposition by
blocking the active edge and corner atoms for carbon growth
Ni-CeAlO3-Al2O3 700 2:1 100 86 - The remarkable syngas selectivity was ascribed to the thin-sheet fiber structure enabling short contact time of reactant [161]
/FeCrAl-fiber on active sites at high GHSV. In addition, the ‘CeAlO3-CeO2’ cycles allow the facile oxygen storing and releasing
process, which enhanced the surface oxygen activity and inhibited the carbon filaments growth
Rh/KCC-1 800 2:1 30 80.3 1.93 The excellent catalytic stability after exposure of oxidation process was attributed to the dendritic fibrous structure [162]
which allowed high accessibility of reactants to active sites, thus rapid restoration of catalysis activity. Additionally,
the presence of oxygen vacancies not only provided labile electrons for enhancing the metal-support interaction and
preserving the metallic phases, but also generated labile oxygen ions for suppressing carbon growth
a
Temp. represents temperature.
b
GHSV is gas hourly space velocity.
c
The unit is h−1.
d
The unit is mL g−1 h−1 CH4.
CATALYSIS REVIEWS 37

ascertaining the kinetic parameters based on the best-fit kinetic model are vital
to optimize the catalyst design and reactor design, leading to enhancement of
MPO respective of cost and energy efficiencies. However, kinetic studies of
MPO are relatively scarce compared to other reforming processes such as MSR
or MDR, and MPO kinetic evaluations might become an important topic in
the near future. In kinetic investigations of LaCr0.85Ru0.15O3 catalyst,
Melchiori et al.[164] applied modified simple rate expressions, whose reaction
orders respect to partial pressure were assumed to match their corresponding
reaction stoichiometric values, to capture the rates of reactant consumption
and product formation (Table 3). To ease the calculations, all reaction rate
constants of 4-reactions mechanism were simply estimated from Arrhenius
equation and water gas shift reaction was assumably an equilibrium process.
Based on the estimated pre-exponential factor and activation energy values,
they evaluated the reaction rates of these 4-reactions mechanism as a function
of the axial coordinate in order to figure out the dependency and changes in
these reactions along with the reactor tube. The results showed that both H2
and CH4 oxidation possessed comparable reaction rates along the reactor
length whilst the MPO was dominant at the front part of reactor under O2-
exessive surrounding. As CH4 and O2 partial pressures reduced along with the
axial of reactor length, CO oxidation rate was remained constant but the H2
oxidation rate became higher owing to the increasing H2 partial pressure. The
analogous method is also used by Bawornruttanaboonya et al.[165] to calculate
the activation energy of reactions considered in the indirect MPO route over
noble
k and P represent the reaction rate constant and partial pressure, whilst the
K is the equilibrium constant.
metal-based catalysts. Besides the abovementioned four reaction mechan­
isms, the MSR and MDR were included for the modeling, as shown in Table 3,
which fitted well the experimental data for predicting the CH4 conversion and
product formations.
Unlike the pre-assumed reaction order values in kinetic works of Melchiori
et al.,[164] some authors simplified the empirical power law model to estimate
the CH4 activation energy and reaction order of each reactant based on the
experimental data as shown in Table 4. The empirical power law model for the
MPO process is described in Equation (1). The -rCH4 and k represent the
reaction rate of CH4 and reaction rate constant with P being the reactant
partial pressure, whereas the α and β are corresponding to the reaction orders.
McCormick et al.[166] conducted the MPO experiment under various operat­
ing conditions and fitted the data obtained from vanadyl pyrophosphate
(VPO) catalyst to a power law model. The results showed that the reaction
order of CH4 was higher than that of O2, suggesting a greater dependence on
the CH4 partial pressure for syngas production. The approximately zero
reaction order of O2 was probably a result of Mars–van Krevelen redox cycles
38
T. J. SIANG ET AL.

Table 3. The multiple reaction mechanisms involved for the MPO process and the estimated kinetic parameters.
Pre-exponential factor, a Apparent activation energy, EA
(mol kgcat−1 s−1) (kJ mol−1)
No. Reactions Rate models ref.[164] r2 ref.[165] r2 ref.[164] r2 ref.[165] r2
1 5 0.99 - - 84.8 0.99 - -
CH4 partial oxidation:CH4 þ 0:5O2 ! 2H2 þ CO r1 ¼ k1 PCH4 PO0:52 1:21 � 10
2 H2 oxidation:H2 þ 0:5O2 ! H2 O r2 ¼ k2 PH2 PO0:52 2:61 � 104 0.99 2:8 � 102 0.99 49.4 0.99 43 0.99
3 CO oxidation:CO þ 0:5O2 ! CO2 2 0.99 0.99 17.1 0.99 45 0.99
r3 ¼ k3 PCO PO0:52 3:04 � 10 1:3 � 102

4 Water gas shift reaction:H2 O þ CO Ð H2 þ CO2 r4 ¼ k4 PCO PH2 O K1 PCO2 PH2 1:71 � 106 0.99 5:7 � 102 0.99 102.2 0.99 84 0.99

8:269 � 102 7:389 � 105


log10 K ¼ 1:247 þ þ
T T2
8 10
1:902 � 10 1:850 � 10
þ 4
5 T3 r ¼ k PT P - - 0.99 - - 70 0.99
CH4 dry reforming:CH4 þ CO2 ! 2H2 þ 2CO 5 5 CH4 CO2 6:0 � 102
6 CH4 steam reforming:CH4 þ H2 O ! 3H2 þ CO r6 ¼ k6 PCH4 PH2 O - - 6:0 � 102 0.99 - - 70 0.99
CATALYSIS REVIEWS 39

Table 4. Summary of kinetic parameters derived from power law models of MPO over different
catalysts.
Reaction order
Activation energy, EA
(kJ mol−1) α β
Temperature
Catalysts (oC) CH4 CH4 O2 References
[166]
VPO 300‒425 102 0.73 0.08
[167]
FePO4 450‒700 164 0.66 0.45
[167]
FePO4/SiO2 450‒700 144 0.61 0.28

with fast catalyst reoxidation. The similar lower reaction order for O2 due to
redox properties of catalyst is also suggested by the findings of Alptekin
et al.[167] From Table 4, the reaction order of CH4 was greater than that of
O2 which indicates that CH4 adsorption on the active sites of catalyst is the
dominant step for MPO.
α β
rCH4 ¼ kPCH P
4 O2
(1)

Given the MPO consists of indirect combustion-reforming pathway,


determining the detailed kinetic parameters based on rate laws is
a challenging task. Nguyen et al.[168] used empirical power law model
to approximately deduce the global kinetic parameters for ascertaining
the actual rate determining cycle over Ni/La2O3, which included the
impact of reactant concentration under various operating conditions.
They compare the power rate laws at low conversion and the global
rate constants of total oxidation and dry reforming routes to affirm the
rate determining cycle of indirect MPO reaction. The computed results
showed that the reaction rate of CH4 total oxidation increased from 0.05
to 0.19 mol m−3 s−1 with increasing temperature (500 to 600°C) for all
different CH 4 /O 2 ratio cases. Interestingly, they found that the O 2
dependency was negligible (zero β values) and CH4 was the only primary
reactant significantly affected the reaction rate, implying that adsorption
of CH4 on catalyst surface dominated the CH4 total oxidation reaction.
Similarly, for the case of MDR, a rise in temperature (from 375‒450°C)
brought a positive effect to reaction rate of MDR (from 0.06 to
0.47 mol m−3 s−1) but led to reduction in the reaction orders of CH4 and
CO2. This was attributed to the competitive adsorption of both reactants
on the metallic Ni0 surface. However, the values of CH4 reaction order
were noticeably higher than CO2, suggesting a greater dependance on
CH4 partial pressure for surface reaction.
Based on the estimated reaction parameters, a conceptual diagram of
indirect MPO kinetics over Ni/La2O3 was proposed as demonstrated in
Figure 18. The reaction rate constant of CH4 total oxidation was around
two-fold slower than that of MDR and MSR, confirming its role as the rate
40 T. J. SIANG ET AL.

Figure 18. (a) Kinetically coupled catalytic CH4 dry and steam reforming cycles over Ni0 assisted by
CH4 total oxidation cycle. Reproduced with permission from Ref.[168] Copyright 2014, Elsevier. (b)
Effect of O2 partial pressure on the ratio of steady-state coverage by molecularly adsorbed
O species to atomic o species estimated using kinetic parameters in Table 6. The numbers in the
plot represent the diameter of rh np on γ-Al2O3 in nm. (c) Apparent rate constant of CH4 activation
over differently sized rh np on γ-al2o3 as determined from CH4/Ar = 1:1 (white circle), CH4/O2
/Ar = 2:1:2 (gray circle) and CH4:Co2/Ar = 1:1:1 (black circle). Reproduced with permission from Ref.
[114]
Copyright 2014, American Chemical Society.

determining cycle during indirect MPO process. This revealed that the MSR
and MSR cycles promptly consumed the H2O and CO2 with a rate regu­
lated by the sluggish kinetic cycle of CH4 total oxidation for overall MPO
process. To intrinsically describe the rate of two-step sequences combus­
tion-reforming mechanism in the indirect process of MPO, Nguyen and the
coworkers[169] further developed the classical power law modeling based on
Quasi-Steady State Approximation (QSSA) theory, as demonstrated in
Table 5. Using the developed power rate law expressions (Equations. 2
and 3 in Table 5), both computed and experimental values of CH4 conver­
sion as a function of contact time were utterly consistent in sequential CH4
total oxidation and MDR cycles. These best-fitted rate expressions further
affirmed that the rate of CH4 total oxidation cycle was the rate determining
cycle for the overall indirect MPO process.
Table 5 Representative 2-step sequences of combustion-reforming kinetic rate models for indirect MPO over Ni0/La2O3 [170].
Assumption Mechanism Kinetic rate model Eq.
� �
CH4 total oxidation occurs k7 k7 (2)
CH4 þ La O !fCH4 g � � � � � O LaðRDSÞ ½L�k7 k9 PCH4
over La2O3 support. k8 rMTO ¼ k
La þ La La þ La OH 1þk7 ðPCH4 Þ
fCH4 g � � � � � O O ! CH3 O 9

CH4 molecular k9
CH3 O La ! :::::
adsorption on La2O3 as (other unspecified steps
RDS. which are not crucial steps kinetically
resulting in H2O and CO2 products)
k10
2La þ O2 ! 2La O
CH4 dry reforming occurs k11 k11 PCH4 (3)
CH4 þ 2X ! CH3 X þH X ðRDSÞ k14 PCO
2 �
over Ni0 where the CO2 k12
rMDR ¼ ½L�k14 PCO2 �
k PCH
being originated from CH3 X þ X ! CHy X þH X; 1 � y � 3 1þk11 P 4
14 CO2

CH4 total oxidation. k13


CH X þ X !C XþH X
k14
Both single-site CH4 CO2 þ X ! CO2 X ðRDSÞ
k15
and CO2 molecular CO2 X þ X ! CO XþO X
adsorptions on Ni0 k16
active sites as RDS. CO2 X ! CO þ X
k17
C XþO X ! CO X þX
k18
CO X ! CO þ X
k19
H XþH X ! H2 þ 2X
RDS and X are being rate determining step and active site whereas [L] and □ represent the concentration of active site and oxygen vacancy site.
CATALYSIS REVIEWS
41
42 T. J. SIANG ET AL.

Table 6. Micro-kinetic models employed in kinetic modeling of activation for CH4, O2 and CO2.
Reproduced with permission from Ref.[114] Copyright 2014, American Chemical Society.
Model Elementary steps Rate constants (s−1)
1 CH4 þ X ! products eff
kreac ðCH4 Þa
2 O2 þ X ! O2 X eff
kads ðO2 Þa
O2 X ! O2 þ X kdes ðO2 Þ
O2 X þ X ! 2O X eff
kdiss ðO2 Þa
2O X ! O2 XþX kass ðO2 Þb
3 CO2 þ X ! CO2 X eff
kads ðCO2 Þa
CO2 X ! CO2 þ X kdes ðCO2 Þ
CO2 X þ X ! CO X þ O X eff
kdiss ðCO2 Þa
CO X þO X ! CO2 XþX kass ðCO2 Þb
eff eff
Note the kads ¼ CX � kads and kdiss ¼ CX � kdis where the CX is active site concentration.

The power law modeling has been extensively applied to estimate the
apparent kinetic parameters due to its simplicity excluding the understanding
of mechanism. However, this empirical model might inaccurate and inapplic­
able for wide-ranging reaction conditions, which fails to reflect the reaction
elementary steps on catalyst surface associated with its properties. On this
basis, in order to correlate the interaction kinetic of reactant gases and particle
sizes of Rh/γ-Al2O3, Kondratenko et al.[114] derived the microkinetic models
(Table 7) and employed them in kinetic assessments of experimental responses
for CH4, O2 and CO2.

Table 7. Estimated kinetic parameters of O2 and CO2 activation over Rh/γ-Al2O3 with different
nanoparticle (NP) sizes.[114].
For o2 activation:
eff eff
kads ðO2 Þ kdes ðO2 Þ kdiss ðO2 Þ kass ðO2 Þ
–1
Catalyst (s ) (s−1) (s−1) (m3 mol−1 s−1)
Rh NP (1.1 nm)/γ-Al2O3 7:4 � 103 9.7 7:3 � 101 3:7 � 102
Rh NP (2.5 nm)/γ-Al2O3 5:2 � 103 1:4 � 10 2
4:9 � 101 7:4 � 102
Rh NP (2.9 nm)/γ-Al2O3 4:3 � 103 7:4 � 101 1:1 � 102 2:6 � 102
Rh NP (3.7 nm)/γ-Al2O3 2:6 � 103 3.2 2:0 � 102 1:3 � 101
Rh NP (5.5 nm)/γ-Al2O3 1:0 � 103 2.7 1:1 � 102 1:0 � 101

For CO2 activation:


Catalyst eff kdes ðCO2 Þ eff kass ðCO2 Þ
kads ðCO2 Þ kdiss ðCO2 Þ
–1
(s ) (s−1) (s−1) (m3 mol−1 s−1)

Rh NP (1.1 nm)/γ-Al2O3 1:0 � 103 1:2 � 102 1:0 � 101 6:6 � 103
Rh NP (2.5 nm)/γ-Al2O3 9:1 � 102 4:4 � 102 5.4 3:8 � 103
Rh NP (2.9 nm)/γ-Al2O3 9:7 � 102 2:0 � 102 6.4 5:5 � 103
Rh NP (3.7 nm)/γ-Al2O3 8:9 � 102 1:3 � 102 6.0 5:3 � 103
Rh NP (5.5 nm)/γ-Al2O3 6:8 � 102 5:4 � 102 6.4 3:2 � 102
CATALYSIS REVIEWS 43

The elementary steps involved for calculations are (i) adsorption/deso­


rption of oxygen molecules and (ii) reversible dissociation of molecularly
adsorbed O species into two atomically adsorbed O species. As demonstrated
eff
in Table 8, the apparent adsorption constants (kads ) for both O2 and CO2
activations decreased with a growth of Rh nanoparticle (NP) sizes, accredited
to the higher oxygen affinity possessed by smaller Rh NP. This indicated that
the degree of adsorption and dissociation of O2 molecule over various Rh NP
sizes was considerably affected by interaction between the rate constants of
individual reaction steps. To illustrate this speculation, calculations of O2
partial pressure influence on the ratio of steady-state coverage by molecularly
adsorbed O species to atomic O species (Θ½O2 �=Θ½O�) has been performed
(Figure 18b) using kinetic values in Table 8.
The Θ½O2 �=Θ½O� ratio substantially grew as the partial pressure of O2
increased, suggesting a drop in pressure could lead to a reduction of adsorbed
O2 molecule coverage. Particularly, the smaller Rh NP exhibited the greater
Θ½O2 �=Θ½O2 �Θ½O� Θ½O� ratio, indicative of Rh NP with larger size favored
the adsorbed oxygen dissociation process. This was attributed to the higher
concentration of reaction center with unique metal configuration in larger Rh
NP facilitating the π bonds activation in O2 molecules and thus enhancement
in reforming reaction rate. This postulation was supplemented by the apparent
first-order kinetic constants values for C‒H bond scission in CH4 molecules
with and without the existence of oxidizing agents (i.e., O2 or CO2) as
a function of Rh NP sizes in Figure 18c.

6. Concluding remarks and future perspectives


This review implicates the recent advances toward development of a flexible
MPO catalyst system for sustainable application in the quest for its wide-
ranged commercialization and industrialization. It emphasized the synergistic
interactions among catalyst constituent components and their significance in
improving catalytic activity and stability. Theoretical and experimental find­
ings are pointing to intrinsic variations in the mechanistic steps across differ­
ent catalyst system. These disparities are closely related to the CH4
dissociation, binding site preferences, surface oxygen concentration and oxy­
gen mobility from catalyst lattice structure and support species. In this context,
the extrinsic and intrinsic properties of a catalyst system, namely, metal-
support interaction, surface acidity and basicity, oxygen vacancies, perovskite
and surface morphology have been widely examined and tailored toward
performance improvement in the optimistic design of MPO catalysts.
Fundamental understanding of these interactions on catalyst performance
have been consistently conducted with the assistance of state-of-the-art
44 T. J. SIANG ET AL.

characterization and experiments using advanced equipment. Apart from that,


computational DFT was applied to compliment experimental data by provid­
ing insight into surface chemistries of catalyst system at atomic level.
Tuning of metal-support interaction of catalyst system can be realized
through appropriate selection or modification of active metals and support
materials. A strong metal-support interaction not only preserves the metallic
active sites by minimizing the oscillation of metal oxidation state during
reaction but also creates spillover effect to facilitate the surface reaction rate.
The adjustment between acidity and basicity strengths can substantially pre­
vent the catalyst deactivation, in which acid sites promote the carbon deposit
originated from CH4 cracking and basic sites facilitate adsorption and activa­
tion of oxidizing agents for carbon gasification. The adverse effect of carbon
deposit is insignificant if rate of carbon gasification is equal to or superior to
that of carbon formation. In addition, oxygen vacancies act as active sites for
cyclic oxygen storing and releasing processes to enhance surface reaction. The
adsorbed oxygen atoms on oxygen vacancies reportedly produce an effective
double layer via O2- back-spillover effect on the metal particles, which weakens
and destabilizes the M‒O bonds in metal oxides and converting them into
metallic phases during reaction. Meanwhile, the partial or total combustion
depends on the nature of the oxide ion involved and type of oxygen interaction
with the perovskite surface. Hence, tuning perovskite lattice structure can
significantly control the rate of oxygen mobility determined by the oxygen
vacancies and surface oxygen concentrations. Modification of catalyst mor­
phology (i.e., core-shell or dendritic fibrous structure) can enhance accessi­
bility of reactant gases to active sites within short contact time, which impedes
the hot spot formation and growth of carbonaceous species, thus achieving
high catalytic activity.
Although enormous progress has been made in optimizing catalyst compo­
nents to improve catalytic performance, activity and stability have always been
antagonistic. There is a regular trade-off between these two primary perfor­
mance measures during catalyst design, depending on the industrial and
economic perspectives. Hence, developing a catalyst system with superior
activity and long-term stability is the key to achieving commercialization
and industrialization of syngas generation using MPO technology. In addition,
the existing power law models can only demonstrate the dominance or
importance of reactant partial pressure in overall catalytic activity but it is
unable to capture the projection of product formation involving complex and
multiple mechanism steps. Success development of mechanism-derived
kinetic models with accurate and consistent kinetic parameters is greatly
advantageous in catalyst and reactor design for improving the commercializa­
tion feasibility of the MPO technology. Research is still primarily focused on
the development of materials and tuning of their properties and interactions to
achieve the desired level of activity and long-term stability. Nevertheless, large
CATALYSIS REVIEWS 45

research efforts are also required to devoted on catalyst lifespan associated


with regenerative and restorative ability, cost efficiency of catalyst synthesis
approaches and extensive mechanism-derived kinetics and computational
DFT studies to cover all aspects of catalytic heterogeneous MPO reaction.

Abbreviations

DFT Density functional theory


DRIFT Diffuse reflectance infrared Fourier transform
DXAS Dispersive X-ray absorption spectroscopy
FTS Fisher-Tropsch synthesis
GTL Gas-to-liquids
MDR Methane dry reforming
MPO Methane partial oxidation
MSR Methane steam reforming
OSC Oxygen storage capacity
POX Partial oxidation
QEXAFS Quick-scanning extended X-ray absorption fine structure
QSSA Quasi-Steady State Approximation
RWGS Revere water-gas shift
TEM Transmission electron microscopy
WGS Water-gas shift
XANES-POM X-ray absorption near edge structure-partial oxidation of methane

Acknowledgments
The authors are grateful for the financial aid by the Transdisciplinary Research Grant
(No. 06G52 and 06G53) from Universiti Teknologi Malaysia and the Universiti Teknologi
Malaysia Zamalah Scholarship (Tan Ji Siang) for this work.

Disclosure statement
No potential conflict of interest was reported by the author(s).

ORCID
T.J. Siang http://orcid.org/0000-0003-3944-674X
A.A. Jalil http://orcid.org/0000-0003-0811-3168
S.Y. Liew http://orcid.org/0000-0001-5802-066X

References
[1] Hannah Ritchie and Max Roser. CO2 and Greenhouse Gas Emissions. https://ourworl
dindata.org/co2-and-other-greenhouse-gas-emissions; 2020 (accessed Mar 22, 2022).
[2] Friedlingstein, P.; Jones, M.; O’sullivan, M.; Andrew, R.; Hauck, J.; Peters, G.;
DBakker, O. Global Carbon Budget 2019. Earth Syst. Sci. Data. 2019, 11(4), 1783‒
1838. DOI: 10.5194/essd-11-1783-2019.
46 T. J. SIANG ET AL.

[3] Peters, G. P.; Andrew, R. M.; Canadell, J. G.; Friedlingstein, P.; Jackson, R. B.;
Korsbakken, J. I.; Peregon, A.; Peregon, A. Carbon Dioxide Emissions Continue to
Grow Amidst Slowly Emerging Climate Policies. Nat. Clim. Change. 2020, 10(1), 3–6.
DOI: 10.1038/s41558-019-0659-6.
[4] Kan, S. Y.; Chen, B.; Wu, X. F.; Chen, Z. M.; Chen, G. Q. Natural Gas Overview for
World Economy: From Primary Supply to Final Demand via Global Supply Chains.
Energy Policy. 2019, 124, 215–225. DOI: 10.1016/j.enpol.2018.10.002.
[5] Dos Santos, R. G.; Alencar, A. C. Biomass-derived Syngas Production via Gasification
Process and Its Catalytic Conversion into Fuels by Fischer Tropsch Synthesis: A Review.
Int. J. Hydrogen Energy. 2020, 45(36), 18114‒18132. DOI: 10.1016/j.
ijhydene.2019.07.133.
[6] Minh, D. P.; Siang, T. J.; Vo, D. V. N.; Phan, T. S.; Ridart, C.; Nzihou, A.; Grouset, D.
Hydrogen Production from Biogas Reforming: An Overview of Steam Reforming, Dry
Reforming, Dual Reforming, and Tri-reforming of Methane. In Hydrogen Supply Chains;
Azzaro-Pantel, C., Ed.; Academic Press: London, 2018; pp 111–166.
[7] Research and Markets Ltd. Syngas & Derivatives Market by Production Technology,
Gasifier Type, Feedstock (Coal, Natural Gas, Petroleum Byproducts, Biomass/waste),
application (chemicals, fuel, and electricity), and region - Global Forecast to 2025,
https://www.researchandmarkets.com/reports/; 2021 (accessed Oct 25, 2021).
[8] Mordor Intelligence. Syngas market‒Growth, Trends, and Forecast (2022-2027). https://
www.mordorintelligence.com/industry-reports/syngas-market; 2021 (accessed Mar 22,
2022).
[9] Lee, N. Smaller-scale and Modular Technologies Drive GTL Industry Forward. http://
www.gasprocessingnews.com/features/201706/smaller-scale-and-modular-technologies
-drive-gtl-industry-forward.aspx/; 2020 (accessed Oct 25, 2021).
[10] Abdullah, B.; Abd Ghani, N. A.; Vo, D. V. N. Recent Advances in Dry Reforming of
Methane over Ni-based Catalysts. J. Clean. Prod. 2017, 162, 170–185. DOI: 10.1016/j.
jclepro.2017.05.176.
[11] Wang, B.; Albarracín-Suazo, S.; Pagán-Torres, Y.; Nikolla, E. Advances in Methane
Conversion Processes. Catal. Today. 2017, 285, 147‒158. DOI: 10.1016/j.
cattod.2017.01.023.
[12] Siang, T. J.; Pham, T. L.; Van Cuong, N.; Phuong, P. T.; Phuc, N. H. H.; Truong, Q. D.;
Vo, D. V. N. Combined Steam and CO2 Reforming of Methane for Syngas Production
over Carbon-resistant Boron-promoted Ni/SBA-15 Catalysts. Micro. Meso. Mater. 2018,
262, 122–132. DOI: 10.1016/j.micromeso.2017.11.028.
[13] Subraveti, S. G.; Roussanaly, S.; Anantharaman, R.; Riboldi, L.; Rajendran, A. Techno-
economic Assessment of Optimised Vacuum Swing Adsorption for Post-combustion
CO2 Capture from Steam-methane Reformer Flue Gas. Sep. Purif. Technol. 2020, 256,
117832. DOI: 10.1016/j.seppur.2020.117832.
[14] Pantaleo, G.; La Parola, V.; Deganello, F.; Singha, R. K.; Bal, R.; Venezia, A. M. Ni/CeO2
Catalysts for Methane Partial Oxidation: Synthesis Driven Structural and Catalytic
Effects. Appl. Catal. B–Environ. 2016, 189, 233–241. DOI: 10.1016/j.apcatb.2016.02.064.
[15] The Global Syngas Technologies Council. Partial Oxidation. https://globalsyngas.org/
syngas-technology/syngas-production/partial-oxidation/; 2018 (accessed Oct 25, 2021).
[16] Linde. Partial Oxidation. Linde Engineering. https://www.linde-engineering.com/en/
process-plants/hydrogen_and_synthesis_gas_plants/gas_generation/partial_oxidation/
index.html/; 2020 (accessed Oct 25, 2021).
[17] Liander, H The Utilisation of Natural Gases for the Ammonia Process. Trans. Faraday
Soc. 1929, 25, 462‒472. DOI: 10.1039/TF9292500462.
CATALYSIS REVIEWS 47

[18] Prettre, M.; Eichner, C.; Perrin, M. The Catalytic Oxidation of Methane to Carbon
Monoxide and Hydrogen. Trans. Faraday Soc. 1946, 42, 335b–339. DOI: 10.1039/
TF946420335B.
[19] Huszar, K.; Racz, G.; Szekely, G. The Catalytic Oxidation of Methane to Carbon
Monoxide and Hydrogen. Acta Chim. Acad. Sci. Hung. 1971, 70, 287. DOI: 10.1039/
TF946420335b.
[20] Gavalas, G. R.; Phichitkul, C.; Voecks, G. E. Structure and Activity of NiOα-Al2O3 and
NiOZrO2 Calcined at High Temperatures: I. Structure. J. Catal. 1984, 88(1), 54‒64. DOI:
10.1016/0021-9517(84)90049-6.
[21] Gavalas, G. R.; Phichitkul, C.; Voecks, G. E. Structure and Activity of NiOα-Al2O3 and
NiOZrO2 Calcined at High Temperatures: II. Activity in the Fuel-rich Oxidation of
Methane. J. Catal. 1984, 88(1), 65‒72. DOI: 10.1016/0021-9517(84)90050-2.
[22] Ashcroft, A. T.; Cheetham, A. K.; Foord, J. A.; Green, M. L. H.; Grey, C. P.; Murrell, A. J.;
Vernon, P. D. F. Selective Oxidation of Methane to Synthesis Gas Using Transition
Metal Catalysts. Nature. 1990, 344(6264), 319‒321. DOI: 10.1038/344319a0.
[23] Hickman, D. A.; Schmidt, L. D. Steps in CH4 Oxidation on Pt and Rh Surfaces:
High-temperature Reactor Simulations. AIChE J. 1993, 39(7), 1164‒1177. DOI:
10.1002/aic.690390708.
[24] Hickman, D. A.; Schmidt, L. D. Production of Syngas by Direct Catalytic Oxidation of
Methane. Science. 1993, 259(5093), 343‒346. DOI: 10.1126/science.259.5093.343.
[25] Choudhary, V. R.; Prabhakar, B.; Rajput, A. M. Beneficial Effects of Noble Metal
Addition to Ni/Al2O3 Catalyst for Oxidative Methane-to-syngas Conversion. J. Catal.
1995, 157(2), 752‒754. DOI: 10.1006/jcat.1995.1342.
[26] Ji, Y.; Li, W.; Xu, H.; Chen, Y. A Study on the Ignition Process for the Catalytic Partial
Oxidation of Methane to Synthesis Gas by MS-TPSR Technique. Catal. Lett. 2001, 71(1),
45‒48. DOI: 10.1023/A:1016648106910.
[27] Enger, B. C.; Lødeng, R.; Holmen, A. A Review of Catalytic Partial Oxidation of Methane
to Synthesis Gas with Emphasis on Reaction Mechanisms over Transition Metal
Catalysts. Appl. Catal. A-Gen. 2008, 346(1‒2), 1‒27. DOI: 10.1016/j.apcata.2008.05.018.
[28] Panov, G. I.; Starokon, E. V.; Ivanov, D. P.; Pirutko, L. V.; Kharitonov, A. S. Active and
Super Active Oxygen on Metals in Comparison with Metal Oxides. Catal. Rev. -Sci. Eng.
2021, 63(4), 597–638. DOI: 10.1080/01614940.2020.1778389.
[29] Mackie, J. C Partial Oxidation of Methane: The Role of the Gas Phase Reactions. Catal.
Rev. -Sci. Eng. 1991, 33(1–2), 169–240. DOI: 10.1080/01614949108020299.
[30] Zhu, Y.; Zhang, S.; Shan, J. J.; Nguyen, L.; Zhan, S.; Gu, X.; Tao, F. In Situ Surface
Chemistries and Catalytic Performances of Ceria Doped with Palladium, Platinum, and
Rhodium in Methane Partial Oxidation for the Production of Syngas. ACS Catal. 2013, 3
(11), 2627–2639. DOI: 10.1021/cs400070y.
[31] Meng, Y.; Ding, C.; Gao, X.; Li, Z.; Li, Z.; Li, Z.; Li, Z. Adsorption of Pd on the Cu (1 1 1)
Surface and Its Catalysis of Methane Partial Oxidation: A Density Functional Theory
Study. Appl. Surf. Sci. 2020, 513, 145724. DOI: 10.1016/j.apsusc.2020.145724.
[32] Mishra, A.; Shafiefarhood, A.; Dou, J.; Li, F. Rh Promoted Perovskites for Exceptional
“Low Temperature” Methane Conversion to Syngas. Catal. Today. 2020, 350, 149‒155.
DOI: 10.1016/j.cattod.2019.05.036.
[33] Rocha, K. D. O.; Macedo, W. C.; Marques, C. M.; Bueno, J. M. Pt/Al2O3La2O3 Catalysts
Stable at High Temperature in Air, Prepared Using a “One-pot” Sol–gel Process:
Synthesis, Characterization, and Catalytic Activity in the Partial Oxidation of CH4.
Chem. Eng. Sci. 2021, 229, 115966. DOI: 10.1016/j.ces.2020.115966.
[34] Bartholomew, C. H Mechanisms of Catalyst Deactivation. Appl. Catal. A–Gen. 2001, 212
(1–2), 17–60. DOI: 10.1016/S0926-860X(00)00843-7.
48 T. J. SIANG ET AL.

[35] Argyle, M. D.; Bartholomew, C. H. Heterogeneous Catalyst Deactivation and


Regeneration: A Review. Catal. 2015, 5(1), 145–269. DOI: 10.3390/catal5010145.
[36] Sengodan, S.; Lan, R.; Humphreys, J.; Du, D.; Xu, W.; Wang, H.; Tao, S. Advances in
Reforming and Partial Oxidation of Hydrocarbons for Hydrogen Production and Fuel
Cell Applications. Renew. Sust. Energ. Rev. 2018, 82, 761‒780. DOI: 10.1016/j.
rser.2017.09.071.
[37] Bashan, V.; Ust, Y. Perovskite Catalysts for Methane Combustion: Applications, Design,
Effects for Reactivity and Partial Oxidation. Int. J. Energ. Res. 2019, 43(14), 7755‒7789.
DOI: 10.1002/er.4721.
[38] Elbadawi, A. H.; Ge, L.; Li, Z.; Liu, S.; Wang, S.; Zhu, Z. Catalytic Partial Oxidation of
Methane to Syngas: Review of Perovskite Catalysts and Membrane Reactors. Catal. Rev. -
Sci. Eng. 2021, 63(1), 1‒67. DOI: 10.1080/01614940.2020.1743420.
[39] Basile, F.; Fornasari, G.; Trifiro, F.; Vaccari, A. Partial Oxidation of Methane: Effect of
Reaction Parameters and Catalyst Composition on the Thermal Profile and Heat
Distribution. Catal. Today. 2001, 64(1‒2), 21‒30. DOI: 10.1016/S0920-5861(00)00505-8.
[40] Zhu, J.; Zhang, D.; King, K. D. Reforming of CH4 by Partial Oxidation: Thermodynamic
and Kinetic Analyses. Fuel. 2001, 80(7), 899‒905. DOI: 10.1016/S0016-2361(00)00165-4.
[41] Xiao, T. C.; Hanif, A.; York, A. P.; Nishizaka, Y.; Green, M. L. Study on the Mechanism
of Partial Oxidation of Methane to Synthesis Gas over Molybdenum Carbide Catalyst.
Phys. Chem. Chem. Phys. 2002, 4(18), 4549‒4554. DOI: 10.1039/B204347E.
[42] Horn, R.; Williams, K. A.; Degenstein, N. J.; Bitsch-Larsen, A.; Dalle Nogare, D.;
Tupy, S. A.; Schmidt, L. D. Methane Catalytic Partial Oxidation on Autothermal Rh
and Pt Foam Catalysts: Oxidation and Reforming Zones, Transport Effects, and
Approach to Thermodynamic Equilibrium. J. Catal. 2007, 249(2), 380‒393. DOI:
10.1016/j.jcat.2007.05.011.
[43] Jang, W. J.; Jeong, D. W.; Shim, J. O.; Kim, H. M.; Roh, H. S.; Son, I. H.; Lee, S. J.
Combined Steam and Carbon Dioxide Reforming of Methane and Side Reactions:
Thermodynamic Equilibrium Analysis and Experimental Application. Appl. Energy.
2016, 173, 80‒91. DOI: 10.1016/j.apenergy.2016.04.006.
[44] Cao, P.; Adegbite, S.; Zhao, H.; Lester, E.; Wu, T. Tuning Dry Reforming of Methane for
FT Syntheses: A Thermodynamic Approach. Appl. Energy. 2018, 227, 190‒197. DOI:
10.1016/j.apenergy.2017.08.007.
[45] Chein, R. Y.; Hsu, W. H. Thermodynamic Analysis of Syngas Production via Chemical
Looping Dry Reforming of Methane. Energy. 2019, 180, 535‒547. DOI: 10.1016/j.
energy.2019.05.083.
[46] Siang, T. J.; Jalil, A. A.; Abdulrasheed, A. A.; Hambali, H. U.; Nabgan, W.
Thermodynamic Equilibrium Study of Altering Methane Partial Oxidation for
Fischer–Tropsch Synfuel Production. Energy. 2020, 198, 117394. DOI: 10.1016/j.
energy.2020.117394.
[47] Boucouvalas, Y.; Zhang, Z.; Verykios, X. E. Partial Oxidation of Methane to Synthesis
Gas via the Direct Reaction Scheme over Ru/TiO2 Catalyst. Catal. Lett. 1996, 40(3–4),
189‒195. DOI: 10.1007/BF00815281.
[48] Goula, M. A.; Lemonidou, A. A.; Grünert, W.; Baerns, M. Methane Partial Oxidation to
Synthesis Gas Using Nickel on Calcium Aluminate Catalysts. Catal. Today. 1996, 32(1‒
4), 149‒156. DOI: 10.1016/S0920-5861(96)00168-X.
[49] Nakagawa, K.; Ikenaga, N.; Suzuki, T.; Kobayashi, T.; Haruta, M. Partial Oxidation of
Methane to Synthesis Gas over Supported Iridium Catalysts. Appl. Catal. A-Gen. 1998,
169(2), 281‒290. DOI: 10.1016/S0926-860X(98)00020-9.
CATALYSIS REVIEWS 49

[50] Basile, F.; Fornasari, G.; Gazzano, M.; Kiennemann, A.; Vaccari, A. Preparation and
Characterisation of a Stable Rh Catalyst for the Partial Oxidation of Methane. J. Catal.
2003, 217(2), 245‒252. DOI: 10.1016/S0021-9517(03)00021-6.
[51] Kim, P.; Kim, Y.; Kim, H.; Song, I. K.; Yi, J. Synthesis and Characterization of
Mesoporous Alumina with Nickel Incorporated for Use in the Partial Oxidation of
Methane into Synthesis Gas. Appl. Catal. A-Gen. 2004, 272(1‒2), 157‒166. DOI:
10.1016/j.apcata.2004.05.055.
[52] Liu, T.; Snyder, C.; Veser, G. Catalytic Partial Oxidation of Methane: Is a Distinction
between Direct and Indirect Pathways Meaningful? Ind. Eng. Chem. Res. 2007, 46(26),
9045‒9052. DOI: 10.1021/ie070062z.
[53] Siang, T. J.; Jalil, A. A.; Hamid, M. Y. S.; Abdulrasheed, A. A.; Abdullah, T. A. T.;
Vo, D. V. N. Role of Oxygen Vacancies in Dendritic Fibrous M/KCC-1 (M= Ru, Pd, Rh)
Catalysts for Methane Partial Oxidation to H2-rich Syngas Production. Fuel. 2020, 278,
118360. DOI: 10.1016/j.fuel.2020.118360.
[54] Nguyen, T. H.; Łamacz, A.; Beaunier, P.; Czajkowska, S.; Domański, M.; Krztoń, A.; Van
Le, T.; Djéga-Mariadassou, G. Partial Oxidation of Methane over Bifunctional Catalyst
I. In Situ Formation of Ni0/La2O3 during Temperature Programmed POM Reaction over
LaNiO3 Perovskite. Appl. Catal. B-Environ. 2014, 152, 360‒369. DOI: 10.1016/j.
apcatb.2014.01.053.
[55] Makarshin, L. L.; Sadykov, V. A.; Andreev, D. V.; Gribovskii, A. G.; Privezentsev, V. V.;
Parmon, V. N. Syngas Production by Partial Oxidation of Methane in a Microchannel
Reactor over a Ni–Pt/La0.2Zr0.4Ce0.4Ox Catalyst. Fuel Process. Technol. 2015, 131, 21‒28.
DOI: 10.1016/j.fuproc.2014.10.031.
[56] Cheng, Z.; Qin, L.; Guo, M.; Xu, M.; Fan, J. A.; Fan, L. S. Oxygen Vacancy Promoted
Methane Partial Oxidation over Iron Oxide Oxygen Carriers in the Chemical Looping
Process. Phys. Chem. Chem. Phys. 2016, 18(47), 32418‒32428. DOI: 10.1039/
C6CP06264D.
[57] Jabbour, K Tuning Combined Steam and Dry Reforming of Methane for “Metgas”
Production: A Thermodynamic Approach and State-of-the-art Catalysts. J. Energy
Chem. 2020, 48, 54‒91. DOI: 10.1016/j.jechem.2019.12.017.
[58] Wang, F.; Li, W. Z.; Lin, J. D.; Chen, Z. Q.; Wang, Y. Crucial Support Effect on the
Durability of Pt/MgAl2O4 for Partial Oxidation of Methane to Syngas. Appl. Catal.
B-Environ. 2018, 231, 292‒298. DOI: 10.1016/j.apcatb.2018.03.018.
[59] Li, B.; Maruyama, K.; Nurunnabi, M.; Kunimori, K.; Tomishige, K. Temperature Profiles
of Alumina-supported Noble Metal Catalysts in Autothermal Reforming of Methane.
Appl. Catal. A-Gen. 2004, 275(1‒2), 157‒172. DOI: 10.1016/j.apcata.2004.07.047.
[60] Deutschmann, O.; Schmidt, L. D. Modeling the Partial Oxidation of Methane in a
Short-contact-time Reactor. AIChE J. 1998, 44(11), 2465‒2477. DOI: 10.1002/
aic.690441114.
[61] Salehi, M. S.; Askarishahi, M.; Godini, H. R.; Schomäcker, R.; Wozny, G. CFD
Simulation of Oxidative Coupling of Methane in Fluidized-bed Reactors: A Detailed
Analysis of Flow-reaction Characteristics and Operating Conditions. Ind. Eng. Chem.
Res. 2016, 55(5), 1149‒1163. DOI: 10.1021/acs.iecr.5b02433.
[62] Partopour, B.; Dixon, A. G. Reduced Microkinetics Model for Computational Fluid
Dynamics (CFD) Simulation of the Fixed-bed Partial Oxidation of Ethylene. Ind. Eng.
Chem. Res. 2016, 55(27), 7296‒7306. DOI: 10.1021/acs.iecr.6b00526.
[63] Carrera, A.; Beretta, A.; Groppi, G. Catalytic Partial Oxidation of Iso-octane over Rh/α-
Al2O3 in an Adiabatic Reactor: An Experimental and Modeling Study. Ind. Eng. Chem.
Res. 2017, 56(17), 4911‒4919. DOI: 10.1021/acs.iecr.7b00255.
50 T. J. SIANG ET AL.

[64] Zhang, Q.; Liu, Y.; Chen, T.; Yu, X.; Wang, J.; Wang, T. Simulations of Methane Partial
Oxidation by CFD Coupled with Detailed Chemistry at Industrial Operating Conditions.
Chem. Eng. Sci. 2016, 142, 126‒136. DOI: 10.1016/j.ces.2015.11.010.
[65] Banke, K.; Hegner, R.; Schröder, D.; Schulz, C.; Atakan, B.; Kaiser, S. A. Power and
Syngas Production from Partial Oxidation of Fuel-rich methane/DME Mixtures in an
HCCI Engine. Fuel. 2019, 243, 97‒103. DOI: 10.1016/j.fuel.2019.01.076.
[66] Gopaul, S. G.; Dutta, A. Dry Reforming of Multiple Biogas Types for Syngas Production
Simulated Using Aspen Plus: The Use of Partial Oxidation and Hydrogen Combustion
to Achieve Thermo-neutrality. Int. J. Hydrogen Energy. 2015, 40(19), 6307‒6318. DOI:
10.1016/j.ijhydene.2015.03.079.
[67] Ma, Y.; Ma, Y.; Chen, Y.; Ma, S.; Li, Q.; Hu, X.; Dong, D.; Buckley, C. E.; Dong, D. Highly
Stable Nanofibrous La2NiZrO6 Catalysts for Fast Methane Partial Oxidation. Fuel. 2020,
265, 116861. DOI: 10.1016/j.fuel.2019.116861.
[68] Guo, J.; Ding, C.; Ma, Z.; Ma, L.; Wang, J.; Shangguan, J.; Zhang, K.; Zhao, M.; Li, Y.;
Wang, M. Highly Dispersed and Stable Pt Clusters Encapsulated within ZSM-5 with Aid
of Sodium Ion for Partial Oxidation of Methane. Fuel. 2021, 289, 119839. DOI: 10.1016/j.
fuel.2020.119839.
[69] Sheppard, T.; Hamill, C. D.; Goguet, A.; Rooney, D. W.; Thompson, J. M. A Low
Temperature, Isothermal Gas-phase System for Conversion of Methane to Methanol
over Cu–ZSM-5. Chem. Commun. 2014, 50(75), 11053‒11055. DOI: 10.1039/
C4CC02832E.
[70] Zhu, J.; Sushkevich, V. L.; Knorpp, A. J.; Newton, M. A.; Mizuno, S. C.; Wakihara, T.; van
Bokhoven, J. A.; Liu, Z.; van Bokhoven, J. A. Cu-erionite Zeolite Achieves High Yield in
Direct Oxidation of Methane to Methanol by Isothermal Chemical Looping. Chem.
Mater. 2020, 32(4), 1448‒1453. DOI: 10.1021/acs.chemmater.9b04223.
[71] Shen, X.; Sun, Y.; Wu, Y.; Wang, J.; Jiang, E.; Xu, X.; Jia, Z.; Jia, Z. The Coupling of CH4
Partial Oxidation and CO2 Splitting for Syngas Production via Double Perovskite-type
Oxides LaFexCo1−xO3. Fuel. 2020, 268, 117381. DOI: 10.1016/j.fuel.2020.117381.
[72] Zhang, X.; Pei, C.; Chang, X.; Chen, S.; Liu, R.; Zhao, Z. J.; Gong, J. FeO6 Octahedral
Distortion Activates Lattice Oxygen in Perovskite Ferrite for Methane Partial Oxidation
Coupled with CO2 Splitting. J. Am. Chem. Soc. 2020, 142(26), 11540‒11549. DOI:
10.1021/jacs.0c04643.
[73] Cruellas, A.; Melchiori, T.; Gallucci, F.; van Sint Annaland, M. Advanced Reactor
Concepts for Oxidative Coupling of Methane. Catal. Rev. -Sci. Eng. 2017, 59(3), 234‒
294. DOI: 10.1080/01614940.2017.1348085.
[74] Liu, R.; Pei, C.; Zhang, X.; Chen, S.; Li, H.; Zeng, L.; Mu, R. Chemical Looping Partial
Oxidation over FeWO <sub>/<sub>sio2 Catalysts. Chinese Journal of Catalysis. 2020, 41
(7), 1140‒1141. DOI: 10.1023/A:1016648106910.
[75] Guo, D.; Wang, G. C. Partial Oxdation of Methane on Anatase and Rutile Defective TiO2
Supported Rh4 Cluster: A Density Functional Theory Study. J. Phys. Chem. C. 2017, 121
(47), 22630. DOI: 10.1021/acs.jpcc.7b07489.
[76] Wang, L.; Li, Z.; Wang, Z.; Chen, X.; Song, W.; Zhao, Z.; Wei, Y.; Zhang, X. Hetero-
metallic Active Sites in Omega (MAZ) Zeolite-catalyzed Methane Partial Oxidation:
A DFT Study. Ind. Eng. Chem. Res. 2021, 60(6), 2400‒2409. DOI: 10.1021/acs.
iecr.0c05457.
[77] Mahyuddin, M. H.; Tanaka, T.; Shiota, Y.; Staykov, A.; Yoshizawa, K. Methane Partial
Oxidation over [Cu2(μ-O)]2+ and [Cu3(μ-O)3]2+ Active Species in Large-pore Zeolites.
ACS Catal. 2018, 8(2), 1500‒1509. DOI: 10.1021/acscatal.7b03389.
CATALYSIS REVIEWS 51

[78] Guo, D.; Wen, J. H.; Wang, G. C. Coordination Dependence of Carbon Deposition
Resistance in Partial Oxidation of Methane on Rh Catalysts. Catal. Today. 2020, 355,
422‒434. DOI: 10.1016/j.cattod.2019.07.017.
[79] Shafiefarhood, A.; Zhang, J.; Neal, L. M.; Li, F. Rh-promoted Mixed Oxides for “Low-
temperature” Methane Partial Oxidation in the Absence of Gaseous Oxidants. J. Mater.
Chem. A. 2017, 5(23), 11930‒11939. DOI: 10.1039/C7TA01398A.
[80] Shafiefarhood, A.; Hamill, J. C.; Neal, L. M.; Li, F. Methane Partial Oxidation Using FeOx
@La0.8Sr0.2FeO3−δ Core–shell Catalyst–transient Pulse Studies. Phys. Chem. Chem. Phys.
2015, 17(46), 31297‒31307. DOI: 10.1039/C5CP05583K.
[81] Mallens, E. P. J.; Hoebink, J. H. B. J.; Marin, G. B. The Reaction Mechanism of the Partial
Oxidation of Methane to Synthesis Gas: A Transient Kinetic Study over Rhodium and
A Comparison with Platinum. J. Catal. 1997, 167(1), 43‒56. DOI: 10.1006/
jcat.1997.1533.
[82] Mudu, F.; Olsbye, U.; Arstad, B.; Diplas, S.; Li, Y.; Fjellvåg, H. Aluminium Substituted
Lanthanum Based Perovskite Type Oxides, Non-stoichiometry and Performance in
Methane Partial Oxidation by Framework Oxygen. Appl. Catal. A-Gen. 2016, 523,
171‒181. DOI: 10.1016/j.apcata.2016.05.013.
[83] Costa, D. S.; Gomes, R. S.; Rodella, C. B.; da Silva Junior, R. B.; Frety, R.; Neto, É. T.;
Brandão, S. T. Study of Nickel, Lanthanum and Niobium-based Catalysts Applied in the
Partial Oxidation of Methane. Catal. Today. 2020, 344, 15‒23. DOI: 10.1016/j.
cattod.2018.10.022.
[84] de Sousa, L. F.; Toniolo, F. S.; Landi, S. M.; Schmal, M. Investigation of Structures and
Metallic Environment of the Ni/Nb2O5 by Different in Situ treatments–Effect on the
Partial Oxidation of Methane. Appl. Catal. A-Gen. 2017, 537, 100‒110. DOI: 10.1016/j.
apcata.2017.03.015.
[85] Araújo, J. C.; Oton, L. F.; Bessa, B.; Neto, A. B.; Oliveira, A. C.; Lang, R.; Otubo, L.;
Bueno, J. M. The Role of Pt Loading on La2O3-Al2O3 Support for Methane Conversion
Reactions via Partial Oxidation and Steam Reforming. Fuel. 2019, 254, 115681. DOI:
10.1016/j.fuel.2019.115681.
[86] Chen, S.; Zeng, L.; Tian, H.; Li, X.; Gong, J. Enhanced Lattice Oxygen Reactivity over
Ni-modified WO3-based Redox Catalysts for Chemical Looping Partial Oxidation of
Methane. ACS Catal. 2017, 7(5), 3548‒3559. DOI: 10.1021/acscatal.7b00436.
[87] Thomas, D. J.; Willi, R.; Baiker, A. Partial Oxidation of Methane: The Role of Surface
Reactions. Ind. Eng. Chem. Res. 1992, 31(10), 2272‒2278. DOI: 10.1021/ie00010a003.
[88] Hohn, K. L.; Schmidt, L. D. Partial Oxidation of Methane to Syngas at High Space
Velocities over Rh-coated Spheres. Appl. Catal. A-Gen. 2001, 211(1), 53‒68. DOI:
10.1016/S0926-860X(00)00835-8.
[89] Minh, D. P.; Pham, X. H.; Siang, T. J.; Vo, D. V. N. Review on the Catalytic
Tri-reforming of methane-Part I: Impact of Operating Conditions, Catalyst
Deactivation and Regeneration. Appl. Catal. A–Gen. 2021, 621, 118202. DOI: 10.1016/
j.apcata.2021.118202.
[90] Wu, J.; Helveg, S.; Ullmann, S.; Peng, Z.; Bell, A. T. Growth of Encapsulating Carbon on
Supported Pt Nanoparticles Studied by in Situ TEM. J. Catal. 2016, 338, 295‒304. DOI:
10.1016/j.jcat.2016.03.010.
[91] Xu, J.; Saeys, M. Improving the Coking Resistance of Ni-based Catalysts by Promotion
with Subsurface Boron. J. Catal. 2006, 242(1), 217–226. DOI: 10.1016/j.jcat.2006.05.029.
[92] Torrez-Herrera, J. J.; Korili, S. A.; Gil, A. Recent Progress in the Application of Ni-based
Catalysts for the Dry Reforming of Methane. Catal. Rev. -Sci. Eng. 2021, 1–58. DOI:
10.1080/01614940.2021.2006891.
52 T. J. SIANG ET AL.

[93] Abdulrasheed, A.; Jalil, A. A.; Gambo, Y.; Ibrahim, M.; Hambali, H. U.; Hamid, M. Y. S.
A Review on Catalyst Development for Dry Reforming of Methane to Syngas: Recent
Advances. Renew. Sus. Energy Rev. 2019, 108, 175–193. DOI: 10.1016/j.rser.2019.03.054.
[94] Djéga-Mariadassou, G.; Boudart, M. Classical Kinetics of Catalytic Reactions. J. Catal.
2003, 216(1–2), 89–97. DOI: 10.1016/S0021-9517(02)00099-4.
[95] Lee, S. J.; Jun, J. H.; Lee, S. H.; Yoon, K. J.; Lim, T. H.; Nam, S. W.; Hong, S. A. Partial
Oxidation of Methane over Nickel-added Strontium Phosphate. Appl. Catal. A–Gen.
2002, 230(1–2), 61–71. DOI: 10.1016/S0926-860X(01)00995-4.
[96] Luo, Z.; Kriz, D. A.; Miao, R.; Kuo, C. H.; Zhong, W.; Guild, C.; He, J.; Willis, B.;
Dang, Y.; Suib, S. L., et al. TiO2 Supported Gold–palladium Catalyst for Effective Syngas
Production from Methane Partial Oxidation. Appl. Catal. A–Gen. 2018, 554, 54–63. DOI:
10.1016/j.apcata.2018.01.020.
[97] Mattos, L. V.; De Oliveira, E. R.; Resende, P. D.; Noronha, F. B.; Passos, F. B. Partial
Oxidation of Methane on Pt/Ce–ZrO2 Catalysts. Catal. Today. 2002, 77(3), 245–256.
DOI: 10.1016/S0920-5861(02)00250-X.
[98] Yan, Q. G.; Wu, T. H.; Weng, W. Z.; Toghiani, H.; Toghiani, R. K.; Wan, H. L.;
Pittman, C. U., Jr Partial Oxidation of Methane to H2 and CO over Rh/SiO2 and
Ru/SiO2 Catalysts. J. Catal. 2004, 226(2), 247–259. DOI: 10.1016/j.jcat.2004.05.028.
[99] Fleys, M.; Simon, Y.; Swierczynski, D.; Kiennemann, A.; Marquaire, P. M. Investigation
of the Reaction of Partial Oxidation of Methane over Ni/La2O3 Catalyst. Energy Fuels.
2006, 20(6), 2321–2329. DOI: 10.1021/ef0602729.
[100] Rabe, S.; Nachtegaal, M.; Vogel, F. Catalytic Partial Oxidation of Methane to Synthesis
Gas over a Ruthenium Catalyst: The Role of the Oxidation State. Phys. Chem. Chem.
Phys. 2007, 9(12), 1461–1468. DOI: 10.1039/B617529E.
[101] Stötzel, J.; Frahm, R.; Kimmerle, B.; Nachtegaal, M.; Grunwaldt, J. D. Oscillatory
Behavior during the Catalytic Partial Oxidation of Methane: Following Dynamic
Structural Changes of Palladium Using the QEXAFS Technique. J. Phys. Chem. C.
2012, 116(1), 599–609. DOI: 10.1021/jp2052294.
[102] Velasco, J. A.; Fernandez, C.; Lopez, L.; Cabrera, S.; Boutonnet, M.; Järås, S. Catalytic
Partial Oxidation of Methane over Nickel and Ruthenium Based Catalysts under Low O2
/CH4 Ratios and with Addition of Steam. Fuel. 2015, 153, 192–201. DOI: 10.1016/j.
fuel.2015.03.009.
[103] Grunwaldt, J. D.; Hannemann, S.; Schroer, C. G.; Baiker, A. 2D-mapping of the Catalyst
Structure inside a Catalytic Microreactor at Work: Partial Oxidation of Methane over
Rh/Al2O3. J. Phys. Chem. B. 2006, 110(17), 8674–8680. DOI: 10.1021/jp060371n.
[104] Pan, C. J.; Tsai, M. C.; Su, W. N.; Rick, J.; Akalework, N. G.; Agegnehu, A. K.;
Cheng, S. Y.; Hwang, B. J. Tuning/exploiting Strong Metal-support Interaction (SMSI)
in Heterogeneous Catalysis. J. Taiwan Inst. Chem. Eng. 2017, 74, 154–186. DOI: 10.1016/
j.jtice.2017.02.012.
[105] Kang, D.; Lim, H. S.; Lee, M.; Lee, J. W. Syngas Production on a Ni-enhanced Fe2O3/Al2
O3 Oxygen Carrier via Chemical Looping Partial Oxidation with Dry Reforming of
Methane. Appl. Energy. 2018, 211, 174‒186. DOI: 10.1016/j.apenergy.2017.11.018.
[106] Li, L.; Dostagir, M. D.; Shrotri, A.; Fukuoka, A.; Kobayashi, H. Partial Oxidation of
Methane to Syngas via Formate Intermediate Found for a Ruthenium–rhenium
Bimetallic Catalyst. ACS Catal. 2021, 11(7), 3782‒3789. DOI: 10.1021/acscatal.0c05491.
[107] Gil-Calvo, M.; Jimenez-Gonzalez, C.; de Rivas, B.; Gutiérrez-Ortiz, J. I.; Lopez-Fonseca,
R. Effect of Ni/Al Molar Ratio on the Performance of Substoichiometric NiAl2O4
Spinel-based Catalysts for Partial Oxidation of Methane. Appl. Catal. B-Environ. 2017,
209, 128‒138. DOI: 10.1016/j.apcatb.2017.02.063.
CATALYSIS REVIEWS 53

[108] Pal, P.; Singha, R. K.; Saha, A.; Bal, R.; Panda, A. B. Defect-induced Efficient Partial
Oxidation of Methane over Nonstoichiometric Ni/CeO2 Nanocrystals. J. Phys. Chem. C.
2015, 119(24), 13610‒13618. DOI: 10.1021/acs.jpcc.5b01724.
[109] Singha, R. K.; Ghosh, S.; Acharyya, S. S.; Yadav, A.; Shukla, A.; Sasaki, T.; Venezia, A. M.;
Pendem, C.; Bal, R. Partial Oxidation of Methane to Synthesis Gas over Pt Nanoparticles
Supported on Nanocrystalline CeO2 Catalyst. Catal. Sci. Technol. 2016, 6(12), 4601‒
4615. DOI: 10.1039/C5CY02088C.
[110] Pantaleo, G.; La Parola, V.; Deganello, F.; Calatozzo, P.; Bal, R.; Venezia, A. M. Synthesis
and Support Composition Effects on CH4 Partial Oxidation over Ni–CeLa Oxides. Appl.
Catal. B-Environ. 2015, 164, 135‒143. DOI: 10.1016/j.apcatb.2014.09.011.
[111] Li, D.; Sakai, S.; Nakagawa, Y.; Tomishige, K. FTIR Study of CO Adsorption on Rh/MgO
Modified with CO, Ni, Fe, or CeO2 for the Catalytic Partial Oxidation of Methane. Phys.
Chem. Chem. Phys. 2012, 14(25), 9204‒9213. DOI: 10.1039/C2CP41050H.
[112] Kaddeche, D.; Djaidja, A.; Barama, A. Partial Oxidation of Methane on Co-precipitated
Ni–Mg/Al Catalysts Modified with Copper or Iron. Int. J. Hydrogen Energy. 2017, 42(22),
15002‒15009. DOI: 10.1016/j.ijhydene.2017.04.281.
[113] Emamdoust, A.; La Parola, V.; Pantaleo, G.; Testa, M. L.; Shayesteh, S. F.; Venezia, A. M.
Partial Oxidation of Methane over SiO2 Supported Ni and NiCe Catalysts. J. Energy
Chem. 2020, 47, 1‒9. DOI: 10.1016/j.jechem.2019.11.019.
[114] Kondratenko, V. A.; Berger-Karin, C.; Kondratenko, E. V. Partial Oxidation of Methane
to Syngas over γ-Al2O3-supported Rh Nanoparticles: Kinetic and Mechanistic Origins of
Size Effect on Selectivity and Activity. ACS Catal. 2014, 4(9), 3136‒3144. DOI: 10.1021/
cs5002465.
[115] Berger-Karin, C.; Sebek, M.; Pohl, M. M.; Bentrup, U.; Kondratenko, V. A.; Steinfeldt, N.;
Kondratenko, E. V. Tailored Noble Metal Nanoparticles on γ‒Al2O3 for High
Temperature CH4 Conversion to Syngas. ChemCatChem. 2012, 4(9), 1368‒1375. DOI:
10.1002/cctc.201200096.
[116] Goodman, E. D.; Latimer, A. A.; Yang, A. C.; Wu, L.; Tahsini, N.; Abild-Pedersen, F.;
Cargnello, M. Low-temperature Methane Partial Oxidation to Syngas with Modular
Nanocrystal Catalysts. ACS Appl. Nano. Mater. 2018, 1(9), 5258‒5267. DOI: 10.1021/
acsanm.8b01256.
[117] Boukha, Z.; Gil-Calvo, M.; de Rivas, B.; González-Velasco, J. R.; Gutiérrez-Ortiz, J. I.;
López-Fonseca, R. Behaviour of Rh Supported on Hydroxyapatite Catalysts in Partial
Oxidation and Steam Reforming of Methane: On the Role of the Speciation of the Rh
Particles. Appl. Catal. A-Gen. 2018, 556, 191‒203. DOI: 10.1016/j.apcata.2018.03.002.
[118] Boukha, Z.; Jiménez-González, C.; Gil-Calvo, M.; de Rivas, B.; González-Velasco, J. R.;
Gutiérrez-Ortiz, J. I.; López-Fonseca, R. MgO/NiAl2O4 as a New Formulation of
Reforming Catalysts: Tuning the Surface Properties for the Enhanced Partial
Oxidation of Methane. Appl. Catal. B-Environ. 2016, 199, 372‒383. DOI: 10.1016/j.
apcatb.2016.06.045.
[119] Jang, W. J.; Shim, J. O.; Kim, H. M.; Yoo, S. Y.; Roh, H. S. A Review on Dry Reforming of
Methane in Aspect of Catalytic Properties. Catal. Today. 2019, 324, 15‒26. DOI:
10.1016/j.cattod.2018.07.032.
[120] Aziz, M. A. A.; Jalil, A. A.; Wongsakulphasatch, S.; Vo, D. V. N. Understanding the Role
of Surface Basic Sites of Catalysts in CO2 Activation in Dry Reforming of Methane:
A Short Review. Catal. Sci. Technol. 2020, 10(1), 35‒45. DOI: 10.1039/C9CY01519A.
[121] Chen, L.; Qi, Z.; Zhang, S.; Su, J.; Somorjai, G. A. Catalytic Hydrogen Production from
Methane: A Review on Recent Progress and Prospect. Catal. 2020, 10(8), 858. DOI:
10.3390/catal10080858.
54 T. J. SIANG ET AL.

[122] Özdemir, H.; Öksüzömer, M. F.; Gürkaynak, M. A. Preparation and Characterization of


Ni Based Catalysts for the Catalytic Partial Oxidation of Methane: Effect of Support
Basicity on H2/CO Ratio and Carbon Deposition. Int. J. Hydrogen Energy. 2010, 35(22),
12147‒12160. DOI: 10.1016/j.ijhydene.2010.08.091.
[123] Song, Y. Q.; Liu, H. M.; He, D. H. Effects of Hydrothermal Conditions of ZrO2 on
Catalyst Properties and Catalytic Performances of Ni/ZrO2 in the Partial Oxidation of
Methane. Energy Fuels. 2010, 24(5), 2817‒2824. DOI: 10.1021/ef1000024.
[124] Ding, C.; Wang, J.; Jia, Y.; Ai, G.; Liu, S.; Liu, P.; Zhang, K.; Han, Y.; Ma, X. Anti-coking
of Yb-promoted Ni/Al2O3 Catalyst in Partial Oxidation of Methane. Int. J. Hydrogen
Energy. 2016, 41(25), 10707‒10718. DOI: 10.1016/j.ijhydene.2016.04.110.
[125] Gambo, Y.; Jalil, A. A.; Triwahyono, S.; Abdulrasheed, A. A. Recent Advances and
Future Prospect in Catalysts for Oxidative Coupling of Methane to Ethylene: A
Review. J. Ind. Eng. Chem. 2018, 59, 218‒229. DOI: 10.1016/j.jiec.2017.10.027.
[126] Krcha, M. D.; Mayernick, A. D.; Janik, M. J. Periodic Trends of Oxygen Vacancy
Formation and C–H Bond Activation over Transition Metal-doped CeO2 (1 1 1)
Surfaces. J. Catal. 2012, 293, 103‒115. DOI: 10.1016/j.jcat.2012.06.010.
[127] Gao, Q.; Hao, J.; Qiu, Y.; Hu, S.; Hu, Z. Electronic and Geometric Factors Affecting
Oxygen Vacancy Formation on CeO2 (111) Surfaces: A First-principles Study from
Trivalent Metal Doping Cases. Appl. Surf. Sci. 2019, 497, 143732. DOI: 10.1016/j.
apsusc.2019.143732.
[128] Chin, Y. H.; Buda, C.; Neurock, M.; Iglesia, E. Reactivity of Chemisorbed Oxygen Atoms
and Their Catalytic Consequences during CH4–O2 Catalysis on Supported Pt Clusters.
J. Am. Chem. Soc. 2011, 133(40), 15958‒15978. DOI: 10.1021/ja202411v.
[129] Scarabello, A.; Dalle Nogare, D.; Canu, P.; Lanza, R. Partial Oxidation of Methane on
Rh/ZrO2 and Rh/Ce‒ZrO2 on Monoliths: Catalyst Restructuring at Reaction
Conditions. Appl. Catal. B-Environl. 2015, 174, 308‒322. DOI: 10.1016/j.
apcatb.2015.03.012.
[130] Osman, A. I.; Meudal, J.; Laffir, F.; Thompson, J.; Rooney, D. Enhanced Catalytic
Activity of Ni on η-Al2O3 and ZSM-5 on Addition of Ceria Zirconia for the Partial
Oxidation of Methane. Appl. Catal. B-Environ. 2017, 212, 68‒79. DOI: 10.1016/j.
apcatb.2016.12.058.
[131] Cheephat, C.; Daorattanachai, P.; Devahastin, S.; Laosiripojana, N. Partial Oxidation of
Methane over Monometallic and Bimetallic Ni-, Rh-, Re-based Catalysts: Effects of Re
Addition, Co-fed Reactants and Catalyst Support. Appl. Catal. A-Gen. 2018, 563, 1‒8.
DOI: 10.1016/j.apcata.2018.06.032.
[132] Yang, W.; Wang, X.; Song, S.; Zhang, H. Syntheses and Applications of Noble-metal-free
CeO2-based Mixed-oxide Nanocatalysts. Chem. 2019, 5(7), 1743‒1774. DOI: 10.1016/j.
chempr.2019.04.009.
[133] Das, S.; Jangam, A.; Jayaprakash, S.; Xi, S.; Hidajat, K.; Tomishige, K.; Kawi, S. Role of
Lattice Oxygen in Methane Activation on Ni-phyllosilicate@Ce1-xZrxO2 Core-shell
Catalyst for Methane Dry Reforming: Zr Doping Effect, Mechanism, and Kinetic
Study. Appl. Catal. B-Environ. 2021, 290, 119998. DOI: 10.1016/j.apcatb.2021.119998.
[134] Toscani, L. M.; Zimicz, M. G.; Martins, T. S.; Lamas, D. G.; Larrondo, S. A. In Situ X-ray
Absorption Spectroscopy Study of CuO–NiO/CeO2–ZrO2 Oxides: Redox
Characterization and Its Effect in Catalytic Performance for Partial Oxidation of
Methane. RSC Adv. 2018, 8(22), 12190–12203. DOI: 10.1039/C8RA01528G.
CATALYSIS REVIEWS 55

[135] Toscani, L. M.; Bellora, M. S.; Huck-Iriart, C.; Soldati, A. L.; Sacanell, J.; Martins, T. S.;
Craievich, A.; Fantini, M. C.; Larrondo, S. A.; Lamas, D. G. NiO/CeO2-Sm2O3
Nanocomposites for Partial Oxidation of Methane: In-situ Experiments by Dispersive
X-ray Absorption Spectroscopy. Appl. Catal. A-Gen. 2021, 626, 118357. DOI: 10.1016/j.
apcata.2021.118357.
[136] Yentekakis, I. V.; Goula, G.; Hatzisymeon, M.; Betsi-Argyropoulou, I.; Botzolaki, G.;
Kousi, K.; Kondarides, D. I.; Taylor, M. J.; Parlett, C. M.; Osatiashtiani, A., et al. Effect of
Support Oxygen Storage Capacity on the Catalytic Performance of Rh Nanoparticles for
CO2 Reforming of Methane. Appl. Catal. B-Environ. 2019, 243, 490‒501. DOI: 10.1016/j.
apcatb.2018.10.048.
[137] Al-Fatesh, A. S.; Arafat, Y.; Kasim, S. O.; Ibrahim, A. A.; Abasaeed, A. E.; Fakeeha, A. H.
In Situ Auto-gasification of Coke Deposits over a Novel Ni-Ce/W-Zr Catalyst by
Sequential Generation of Oxygen Vacancies for Remarkably Stable Syngas Production
via CO2-reforming of Methane. Appl. Catal. B-Environ. 2021, 280, 119445. DOI:
10.1016/j.apcatb.2020.119445.
[138] Tang, M.; Xu, L.; Fan, M. Progress in Oxygen Carrier Development of Methane-based
Chemical-looping Reforming: A Review. Appl. Energy. 2015, 151, 143‒156. DOI:
10.1016/j.apenergy.2015.04.017.
[139] Qin, L.; Guo, M.; Liu, Y.; Cheng, Z.; Fan, J. A.; Fan, L. S. Enhanced Methane Conversion
in Chemical Looping Partial Oxidation Systems Using a Copper Doping Modification.
Appl. Catal. B-Environ. 2018, 235, 143–149. DOI: 10.1016/j.apcatb.2018.04.072.
[140] Shen, Q.; Huang, F.; Tian, M.; Zhu, Y.; Li, L.; Wang, J.; Wang, X. Effect of Regeneration
Period on the Selectivity of Synthesis Gas of Ba-hexaaluminates in Chemical Looping
Partial Oxidation of Methane. ACS Catal. 2019, 9(1), 722–731. DOI: 10.1021/
acscatal.8b03855.
[141] Bian, Z.; Wang, Z.; Jiang, B.; Hongmanorom, P.; Zhong, W.; Kawi, S. A Review on
Perovskite Catalysts for Reforming of Methane to Hydrogen Production. Renew. Sust.
Energ. Rev. 2020, 134, 110291. DOI: 10.1016/j.rser.2020.110291.
[142] Bhattar, S.; Abedin, M. A.; Kanitkar, S.; Spivey, J. J. A Review on Dry Reforming of
Methane over Perovskite Derived Catalysts. Catal. Today. 2021, 365, 2–23. DOI:
10.1016/j.cattod.2020.10.041.
[143] Chang, H.; Bjørgum, E.; Mihai, O.; Yang, J.; Lein, H. L.; Grande, T.; Raaen, S.; Zhu, Y. A.;
Holmen, A.; Chen, D. Effects of Oxygen Mobility in La–Fe-based Perovskites on the
Catalytic Activity and Selectivity of Methane Oxidation. ACS Catal. 2020, 10(6),
3707–3719. DOI: 10.1021/acscatal.9b05154.
[144] Mihai, O.; Chen, D.; Holmen, A. Chemical Looping Methane Partial Oxidation: The
Effect of the Crystal Size and O Content of LaFeO3. J. Catal. 2012, 293, 175–185. DOI:
10.1016/j.jcat.2012.06.022.
[145] Yang, J.; Bjørgum, E.; Chang, H.; Zhu, K. K.; Sui, Z. J.; Zhou, X. G.; Holmen, A.;
Zhu, Y. A.; Chen, D. On the Ensemble Requirement of Fully Selective Chemical
Looping Methane Partial Oxidation over La-Fe-based Perovskites. Appl. Catal.
B-Environ. 2022, 301, 120788. DOI: 10.1016/j.apcatb.2021.120788.
[146] Jiang, B.; Li, L.; Zhang, Q.; Ma, J.; Zhang, H.; Yu, K.; Bian, Z.; Zhang, X.; Ma, X.; Tang, D.
Iron–oxygen Covalency in Perovskites to Dominate Syngas Yield in Chemical Looping
Partial Oxidation. J. Mater. Chem. A. 2021, 9(22), 13008–13018. DOI: 10.1039/
D1TA02103F.
[147] Marek, E.; Hu, W.; Gaultois, M.; Grey, C. P.; Scott, S. A. The Use of Strontium Ferrite in
Chemical Looping Systems. Appl. Energy. 2018, 223, 369–382. DOI: 10.1016/j.
apenergy.2018.04.090.
56 T. J. SIANG ET AL.

[148] Muñoz-García, A. B.; Bugaris, D. E.; Pavone, M.; Hodges, J. P.; Huq, A.; Chen, F.; Zur
Loye, H. C.; Carter, E. A. Unveiling Structure–property Relationships in Sr2Fe1.5Mo0.5
O6-δ, an Electrode Material for Symmetric Solid Oxide Fuel Cells. J. Am. Chem. Soc.
2012, 134(15), 6826–6833. DOI: 10.1021/ja300831k.
[149] Ma, Y.; Ma, Y.; Long, G.; Li, J.; Hu, X.; Ye, Z.; Dong, D.; Buckley, C. E.; Dong, D.
Synergistic Promotion Effect of MgO and CeO2 on Nanofibrous Ni/Al2O3 Catalysts for
Methane Partial Oxidation. Fuel. 2019, 258, 116103. DOI: 10.1016/j.fuel.2019.116103.
[150] Li, Y.; Wang, J.; Ding, C.; Ma, L.; Xue, Y.; Guo, J.; Liu, P. Effect of Cobalt Addition on the
Structure and Properties of Ni–MCM-41 for the Partial Oxidation of Methane to Syngas.
RSC Adv. 2019, 9(44), 25508‒25517. DOI: 10.1039/C9RA03534F.
[151] Liu, Y.; Qin, L.; Cheng, Z.; Goetze, J. W.; Kong, F.; Fan, J. A.; Fan, L. S. Near 100% CO
Selectivity in Nanoscaled Iron-based Oxygen Carriers for Chemical Looping Methane
Partial Oxidation. Nat. Commun. 2019, 10(1), 1‒6. DOI: 10.1038/s41467-019-13560-0.
[152] Guo, S.; Wang, J.; Ding, C.; Duan, Q.; Ma, Q.; Zhang, K.; Liu, P. Confining Ni
Nanoparticles in Honeycomb-like Silica for Coking and Sintering Resistant Partial
Oxidation of Methane. Int. J. Hydrogen Energy. 2018, 43(13), 6603‒6613. DOI:
10.1016/j.ijhydene.2018.02.035.
[153] Ding, C.; Wang, J.; Guo, S.; Ma, Z.; Li, Y.; Ma, L.; Zhang, K. Abundant Hydrogen
Production over Well Dispersed Nickel Nanoparticles Confined in Mesoporous Metal
Oxides in Partial Oxidation of Methane. Int. J. Hydrogen Energy. 2019, 44(57), 30171‒
30184. DOI: 10.1016/j.ijhydene.2019.09.202.
[154] Singha, R. K.; Shukla, A.; Yadav, A.; Konathala, L. S.; Bal, R. Effect of Metal-support
Interaction on Activity and Stability of Ni-CeO2 Catalyst for Partial Oxidation of
Methane. Appl. Catal. B-Environ. 2017, 202, 473‒488. DOI: 10.1016/j.
apcatb.2016.09.060.
[155] Ding, C.; Ai, G.; Zhang, K.; Yuan, Q.; Han, Y.; Ma, X.; Wang, J.; Liu, S. Coking Resistant
Ni/ZrO2@SiO2 Catalyst for the Partial Oxidation of Methane to Synthesis Gas. Int.
J. Hydrogen Energy. 2015, 40(21), 6835‒6843. DOI: 10.1016/j.ijhydene.2015.03.094.
[156] Neal, L.; Shafiefarhood, A.; Li, F. Effect of Core and Shell Compositions on MeOx@Lay
Sr1‒yFeO3 Core–shell Redox Catalysts for Chemical Looping Reforming of Methane.
Appl. Energy. 2015, 157, 391‒398. DOI: 10.1016/j.apenergy.2015.06.028.
[157] Li, L.; He, S.; Song, Y.; Zhao, J.; Ji, W.; Au, C. T. Fine-tunable Ni@porous Silica Core‒
shell Nanocatalysts: Synthesis, Characterization, and Catalytic Properties in Partial
Oxidation of Methane to Syngas. J. Catal. 2012, 288, 54‒64. DOI: 10.1016/j.
jcat.2012.01.004.
[158] Dou, Y.; Pang, Y.; Gu, L.; Ding, Y.; Jiang, W.; Feng, X.; Ji, W.; Au, C. T. Core-shell
Structured Ru-Ni@SiO2: Active for Partial Oxidation of Methane with Tunable H2/CO
Ratio. J. Energy Chem. 2018, 27(3), 883‒889. DOI: 10.1016/j.jechem.2017.07.011.
[159] Shafiefarhood, A.; Galinsky, N.; Huang, Y.; Chen, Y.; Li, F. Fe2O3@LaxSr1‒xFeO3 Core‒
shell Redox Catalyst for Methane Partial Oxidation. ChemCatChem. 2014, 6(3), 790‒799.
DOI: 10.1002/cctc.201301104.
[160] Neal, L. M.; Shafiefarhood, A.; Li, F. Dynamic Methane Partial Oxidation Using a Fe2O3
@ La0.8Sr0.2FeO3-δ Core–shell Redox Catalyst in the Absence of Gaseous Oxygen. ACS
Catal. 2014, 4(10), 3560‒3569. DOI: 10.1021/cs5008415.
[161] Zhao, G.; Chai, R.; Zhang, Z.; Sun, W.; Liu, Y.; Lu, Y. High-performance Ni-CeAlO3-Al2
O3/FeCrAl-fiber Catalyst for Catalytic Oxy-methane Reforming to Syngas. Fuel. 2019,
58, 116102. DOI: 10.1016/j.fuel.2019.116102.
CATALYSIS REVIEWS 57

[162] Siang, T. J.; Jalil, A. A.; Abdulrahman, A.; Hambali, H. U. Enhanced Carbon Resistance
and Regenerability in Methane Partial Oxidation to Syngas Using Oxygen Vacancy-rich
Fibrous Pd, Ru and Rh/KCC-1 Catalysts. Environ. Chem. Lett. 2021, 19(3), 2733‒2742.
DOI: 10.1007/s10311-021-01192-0.
[163] Das, S.; Gupta, R.; Kumar, A.; Shah, M.; Sengupta, M.; Bhandari, S.; Bordoloi, A. Facile
Synthesis of Ruthenium Decorated Zr0.5Ce0.5O2 Nanorods for Catalytic Partial
Oxidation of Methane. ACS Appl. Nano Mater. 2018, 1(6), 2953‒2961. DOI: 10.1021/
acsanm.8b00567.
[164] Melchiori, T.; Di Felice, L.; Mota, N.; Navarro, R. M.; Fierro, J. L. G.; van Sint
Annaland, M.; Gallucci, F. J. A. C. A. G. Methane Partial Oxidation over a LaCr0.85
Ru0.15O3 Catalyst: Characterization, Activity Tests and Kinetic Modeling. Appl. Catal.
A-Gen. 2014, 486, 239‒249. DOI: 10.1016/j.apcata.2014.08.040.
[165] Bawornruttanaboonya, K.; Laosiripojana, N.; Mujumdar, A. S.; Devahastin, S. Catalytic
Partial Oxidation of CH4 over Bimetallic Ni-Re/Al2O3: Kinetic Determination for
Application in Microreactor. AIChE J. 2018, 64(5), 1691‒1701. DOI: 10.1002/aic.16037.
[166] McCormick, R. L.; Alptekin, G. O.; Herring, A. M.; Ohno, T. R.; Dec, S. F. Methane
Partial Oxidation over Vanadyl Pyrophosphate and the Effect of Fe and Cr Promoters on
Selectivity. J. Catal. 1997, 172(1), 160‒169. DOI: 10.1006/jcat.1997.1840.
[167] Alptekin, G. O.; Herring, A. M.; Williamson, D. L.; Ohno, T. R.; McCormick, R. L.
Methane Partial Oxidation by Unsupported and Silica Supported Iron Phosphate
Catalysts: Influence of Reaction Conditions and Co-feeding of Water on Activity and
Selectivity. J. Catal. 1999, 181(1), 104‒112. DOI: 10.1006/jcat.1998.2297.
[168] Nguyen, T. H.; Łamacz, A.; Krztoń, A.; Ura, A.; Chałupka, K.; Nowosielska, M.;
Rynkowski, J.; Djéga-Mariadassou, G. Partial Oxidation of Methane over Ni0/La2O3
Bifunctional Catalyst II: Global Kinetics of Methane Total Oxidation, Dry Reforming
and Partial Oxidation. Appl. Catal. B-Environ. 2015, 165, 389‒398. DOI: 10.1016/j.
apcatb.2014.10.019.
[169] Nguyen, T. H.; Łamacz, A.; Krztoń, A.; Djéga-Mariadassou, G. Partial Oxidation of
Methane over Ni0/La2O3 Bifunctional Catalyst IV: Simulation of Methane Total
Oxidation, Dry Reforming and Partial Oxidation Using the Quasi-Steady State
Approximation. Appl. Catal. B-Environ. 2016, 199, 424‒432. DOI: 10.1016/j.
apcatb.2016.06.034.

You might also like