Next Article in Journal
Physical, Mechanical and Durability Properties of Ecofriendly Ternary Concrete Made with Sugar Cane Bagasse Ash and Silica Fume
Previous Article in Journal
Atomic Simulation of the Melting and Mechanical Behaviors of Silicon Nanowires
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Oxidative Coupling of Methane: Perspective for High-Value C2 Chemicals

State Key Laboratory of Fine Chemicals, College of Chemical Engineering, Dalian University of Technology, Dalian 116024, China
*
Author to whom correspondence should be addressed.
These authors contributed equally.
Crystals 2021, 11(9), 1011; https://doi.org/10.3390/cryst11091011
Submission received: 19 July 2021 / Revised: 18 August 2021 / Accepted: 20 August 2021 / Published: 24 August 2021
(This article belongs to the Special Issue Catalytic Conversion of Carbonaceous Materials to Fuels and Chemicals)

Abstract

:
The oxidative coupling of methane (OCM) to C2 hydrocarbons (C2H4 and C2H6) has aroused worldwide interest over the past decade due to the rise of vast new shale gas resources. However, obtaining higher C2 selectivity can be very challenging in a typical OCM process in the presence of easily oxidized products such as C2H4 and C2H6. Regarding this, different types of catalysts have been studied to achieve desirable C2 yields. In this review, we briefly presented three typical types of catalysts such as alkali/alkaline earth metal doped/supported on metal oxide catalysts (mainly for Li doped/supported catalysts), modified transition metal oxide catalysts, and pyrochlore catalysts for OCM and highlighted the features that play key roles in the OCM reactions such as active oxygen species, the mobility of the lattice oxygen and surface alkalinity of the catalysts. In particular, we focused on the pyrochlore (A2B2O7) materials because of their promising properties such as high melting points, thermal stability, surface alkalinity and tunable M-O bonding for OCM reaction.

1. Introduction

Methane, being an abundant hydrocarbon resource, provides a kind of comparably cheaper and environmentally friendly fuel [1,2]. The use of natural gas as an important industrial feedstock has increased as compared to petroleum over the past few decades due to the extraction of natural gas from unconventional sources, for example shale deposits [3]. Figure 1 reveals the recoverable shale gas reserves in the world in trillion cubic feet (TCF).
Ethylene is used as a raw material in the manufacture of polymers such as polyethylene, polystyrene, polyvinyl chloride and polyethylene terephthalate, as well as fibers, petrochemical products and organic chemicals. It is also used as the main feedstock (~50%) in the chemical industry. Hence, many global chemical companies are planning to develop economical and feasible processes to produce ethylene directly from methane [3,4].
In general, steam and the thermal cracking of naphtha processes are used for ethylene production in petroleum refineries. Global ethylene production is mainly dependent on these processes [5,6]. However, some of the key problems identified here are energy intensiveness (up to 40 GJ heat per tonne ethylene) and environmental unsustainability [7]. High temperatures (>1200 °C) are required for breaking the C-H bond (bond energy, 440 kJ/mol) in this process and approximately three tons of CO2 is released per ton of ethylene production [8]. Therefore, another way ethylene can be produced from natural gas is via indirect paths by the formation of syngas (CO/H2) through the catalytic Fischer-Tropsch synthesis (FTS) process [9,10]. Alternatively, syngas can be transformed into methanol and then methanol to ethylene [11,12]. In addition to this, methane can also be converted directly to value-added chemicals and fuels through oxidative and non-oxidative catalytic reaction pathways. The non-oxidative processes involve the coupling of methane to olefins; this process suffers from intrinsic thermodynamic limitations and needs higher energy for large-scale applications [13]. Therefore, most of the research for direct conversion has emphasized a process called oxidative coupling of methane (OCM) [14,15,16,17,18,19].
Zhao et al. [14] prepared hierarchically structured porous flower-like materials such as La2O3-H- and Sr-doped (Sr0.1LaOx-H) compounds and highly dense non-porous (La2O3-D and Sr0.1LaOx-D) microspheres. The porous materials showed higher catalytic performances in OCM than non-porous catalysts, presumably due to the creation of the active sites in porous materials, which could enhance the activation of oxygen. Sr0.1LaOx-H showed better OCM results than La2O3-H and reached the optimum selectivity (~48%) and C2 yield (~19%) at 550 °C due to Sr isolating the active sites of the catalyst, which can act as a strong basic site. However, deactivation occurred in the Sr0.1LaOx-H catalyst during OCM in ~30 h, but adding a small amount of Zr can produce better stability in OCM. Lopes et al. [15] prepared La2Ti2O7-based catalysts (LaTi1−xMgxO3+δ) with varied Ti/Mg molar ratios through the partial substitution of Ti by Mg to regulate the quantities of alkaline sites and surface oxygen species, properties crucial for OCM reaction. The characterization data testifies that alkalinity and the surface oxygen vacancy concentration increase as Mg increases. The presence of high amounts of Mg on the surface can increase the incidence of defective sites in the interface formed by La-O-Mg. In general, the rich defects are active and selective toward C2 production in the OCM reaction. Hence, the catalytic results showed that methane conversion and C2 selectivity increased as a function of the Mg amount, and also this catalyst exhibited high thermal stability after 24 h on stream. Sato et al. [16] proposed that, in the presence of an electric field, the CePO4 nanorod catalysts showed high activity and stability even at low temperatures for OCM (C2 yield = 18%) without any external heating. Because, in the electric field, these catalysts can be produced and regenerate the surface-active oxygen species even at low temperatures, which are mainly responsible for OCM, Schmack et al. proposed a meta-analysis method. It is often useful ascertain the links between a catalyst’s physicochemical characteristics and its performance in an OCM reaction [17]. Pirro et al. [18] proposed the process simulator and illustrated it for the OCM to determine the catalyst-dependent kinetics for industrial implementation. For the case study, the feeding of oxygen and cooling were operated under a multi-stage adiabatic configuration. Three catalysts (Sn-Li/MgO, NaMnW/SiO2 and Sr/La2O3) were introduced to estimate the methane conversion and C2+ yield compared to a single stage. They revealed that multi-phase configuration is advantageous compared to a single phase.
The OCM process follows the value-added transformation of natural gas through the catalytic reaction between methane and oxygen for ethylene production. Although more than four decades have passed since the initial work on OCM, this reaction has still not been able to be used as a successful industrial process for ethylene production because of the lower yield (<30%) of C2 products.
In the reaction mechanism of OCM, the first step of abstraction of a hydrogen atom from methane with the help of surface-active oxygen results in the formation of hydroxyl groups, which react with each other to produce H2O, and at the same time create an oxygen vacancy on the catalyst surface. Meanwhile, methyl radicals will combine to form ethane which further undergoes a dehydrogenation reaction to produce the desired product of ethylene, as shown in Figure 2. According to previous reports, active oxygen species on the surface of catalyst consist of peroxide (O22−), lattice oxygen (O2−), carbonate (CO32−), superoxide (O2) and hydroxide (OH) ions. Among all, surface lattice oxygen (O2−) anions play a crucial role in attaining the C2 product selectively, whereas surface electrophilic anions (O, O2 and O22−) usually result in an over-oxidation of the COx product [20,21].
From the pioneering work of Keller and Bhasin [22] and Ito and Lunsford [23], various catalysts have been investigated to attain adequate methane conversion and C2 selectivity/yield. Most of the metal oxides in the periodic table, such as transition metal oxides, rare earth metal oxides, alkali or alkaline earth metal oxides, perovskites and pyrochlore compounds, have been individually or in various combinations tested for OCM reaction [3,24,25,26,27,28]. Zavyalova et al. [29] conducted a comprehensive statistical analysis of about ~1000 research articles related to OCM. They pointed out that both rare-earth metal oxides and alkaline-earth metal oxides, along with alkali metal oxides, are good candidates for OCM reaction. In addition, binary and ternary combinations of metal oxides are also crucial for high and sustainable catalytic activity in the OCM reaction. The selectivity of the catalysts can be further improved by using various dopants such as alkali (Cs, Na) and alkaline earth (Sr, Ba) metals, whereas Mn, W and the Cl anion dopants can be used to enhance the catalytic activity. Figure 3 represents the most promising OCM catalytic systems with C2 yields greater than 25%.
This review presents three types of catalysts for OCM reaction and highlights the features that play key roles in the OCM reactions, such as active oxygen species, the mobility of the lattice oxygen, and the surface alkalinity of the catalysts. The catalyst compositions and their catalytic functions are also discussed. Special attention has been focused on pyrochlore (A2B2O7) materials with promising properties such as high melting points, thermal stability, surface alkalinity and tunable M-O bonding, these being essential for OCM reaction.

2. Catalysts for Oxidative Coupling of Methane

Oxide catalysts are the main catalysts used for OCM reaction. These can be a pristine form or modified with group alkali/alkaline earth metals or transitional metals. These catalytic systems are established under various synthesis methods such as hydrothermal, sol-gel, precipitation, impregnation, thermal decomposition and flame spray pyrolysis to yield tunable catalyst compositions by varying basicity and oxygen vacancies, which are the most important characteristics of OCM catalysts and also key factors in activating methane. These catalysts can be coupled with tunable reaction conditions such as space, velocity and temperature to enhance methane conversion, selectivity and yield [19,30].

2.1. Alkali/Alkaline Earth Metal Doped/Supported on Metal Oxide Catalysts

Pristine or unmodified metal oxide catalysts generally exhibit low C2 selectivity/yield and rapid deactivation in OCM reaction. Modified systems are considered to achieve higher yield at lower temperature regions due to the altered either basicity or oxygen vacancies and/or other defects, which are crucial for CH4 activation [31,32,33,34,35,36]. Ferreira et al. [32] developed CeO2 catalysts modified with earth alkaline metals (M = Mg, Ca, and Sr) and demonstrated that among these three dopants, Ca-doped CeO2 catalyst showed the best performance in the OCM reaction. This is mainly due to the similar ionic radii of Ca2+ and Ce4+ ions and the higher ratio of the oxygen species O2 and O22− to lattice oxygen and a high amount of surface basic sites. Yıldız et al. [33] proved that mesoporous TiO2-rutile supported MnxOy-Na2WO4 catalysts’ better performance than commercial TiO2-rutile- and TiO2-anatase-supported catalysts. In addition, this catalyst is stable during 16 h on stream in OCM. Ivanov et al. [34] synthesized Mg, Al, Ca, Ba and Pb-substituted titanates SrTi1−xAxO3(A = Mg, Al, Ca, Ba, Pb, x = 0.1) and Sr2Ti1−xAxO4(A = Mg, Al, x = 0.1) catalysts by using a mechanochemical method and tested these catalysts for OCM reaction at 850 and 900 °C. Finally, they demonstrated that among these dopants, Mg- and Al-doped SrTiO3 and Sr2TiO4 catalysts showed the best performance in the OCM reaction to provide a higher C2 yield up to 25% and C2 selectivity around 66%, presumably due to (Sr, Mg)O oxide segregate on the surface, which has active oxygen ion-radicals. Peng et al. [35] developed SnO2 modified with alkaline metals and demonstrated that the Li-doped SnO2 catalyst showed the best performance in the OCM reaction among these dopants. This is mainly due to the higher surface intermediate alkaline sites and electrophilic oxygen species. Moreover, this catalyst showed stable catalytic performance up to 100 h without deactivation. Finally, they proposed that the catalytic activity and selectivity towards the desired products were higher for the optimized catalyst (Sn5Li5) than Mn-Na2WO4/SiO2 at a lower temperature (750 °C), which is considered a state-of-the-art catalyst for OCM. Kim et al. [37] carefully explored the role of oxygen species using lanthanum-based perovskite catalysts, which are prepared by a citrate sol-gel method. Their study found that the active surface lattice oxygen species and a facile filling of lattice oxygen vacancies by gas-phase oxygen are responsible for the production of C2 hydrocarbons in OCM.
Elkins et al. [38] studied rare earth oxides (Sm2O3, TbOx, PrOy and CeO2) doped with alkali (Li and Na) and alkaline earth metal (Mg and Ca) supported on MgO. These metal dopants can alter the acid-base property of catalysts, which has an impact on catalysts activity, selectivity and stability. Hence, doped catalysts showed higher C2 yields than undoped catalysts owing to the presence of strong basic sites (see Figure 4). Among all catalysts, Li-TbOx/n-MgO showed better activity, C2 selectivity and C2 yield at 650 °C. However, undoped catalysts showed higher yield at a lower temperature (<600 °C) than the doped catalysts because the active sites in doped catalysts are blocked with by-product CO2 at low temperatures.
Li-doped MgO is one of the best systems for OCM reaction [39,40,41,42,43,44,45] because, in this system, an additional electron is produced when each Li+ cation replaces an Mg+2 cation. This excess of an electron is compensated by the formation of electron holes strongly bonded to lattice oxygen, and O centers are formed with a strong radical character localized to the O atom, as in the case of Li-doped MgO forming [Li+O] centers. [Li+O] catalyzes the CH4 dehydrogenation in the following steps:
[Li+O] + CH4 → [Li+OH] + CH3
2[Li+OH] → [Li+O2− ] + Li+V + H2O
[Li+O2−] + Li+V + 1/2O2 → 2[Li+O]
Here, ‘V’ denotes an oxygen vacancy. The [Li+O] centers are generated in Equation (3) again to fulfill a redox cycle [46]. Ali Farsi et al. [41] reported that 7% Li/MgO catalysts were prepared with two different preparation methods, such as incipient wetness impregnation (IWI) and sol-gel (SG). The catalyst prepared by using the sol-gel method exhibited better methane conversion, higher C2 selectivity, and yield for the desired product than the catalyst prepared by the IWI method. The better performance of the catalyst prepared through the sol-gel method was suggested to relate to the following two factors: (i) it has a significantly larger surface area, and (ii) it comprises a greater quantity of Li incorporated in the MgO matrix. Both of these effects contribute to an increase in the number of active sites in the catalyst. However, at 850 °C, the selectivity of Li/MgO-SG catalyst decreased because of the evaporation of some of the Li from the MgO surface. Qian et al. [42] reported that Li can restructure the MgO surface to expose high-indexed facets such as {110}, {111} and {100} facets; among them, the Mg4c2+ sites of MgO {110} facets are highly active and selective in catalyzing the OCM reaction, which produces high C2 selective products by the formation of methyl radical intermediate without other less-stable carbon-containing radicals (CH2 and CH), which will reduce the methyl radical dissociation and subsequent combustion reaction. Li-doped MgO catalysts with different Li loadings (1.3% and 5.6%) have been prepared by Luo et al. [47]. Different Li contents were observed over both the as-prepared catalysts, but similar Li was present in the used catalysts with different Li surface distribution. Hence, higher CH4 conversion and C2 selectivity were observed over Li (5.6%)/MgO than the Li (1.3%)/MgO catalyst because the loss of Li occurred during the reaction process through the movement of these ions from the bulk material to the catalyst surface and successive desorption was responsible for asymmetrical and coarse structures with bare MgO {110} and {111} facets observed in Li (5.6%)/MgO catalysts but not in the Li (1.3%)/MgO. Hence, the main role of Li in MgO-based catalyst during OCM reaction acts as a structural modifier of the active MgO component rather than as an active center. Due to the poor stability of these Li-supported catalysts, some researchers have recently focused on depressing Li-ions’ evaporation at high temperatures. Matsumoto et al. [48] proposed that the catalytic activity and selectivity towards desired products were higher over crystalline Li2CaSiO4 than Mn-Na2WO4/SiO2, which is considered a state-of-the-art catalyst for OCM. The CH4 conversion and C2 selectivity were obtained at 28.3% and 77.5%, respectively, at the temperature of 750 °C for the OCM reaction. In addition, they exhibited higher catalyst durability for 50 h on stream without deactivation and structural changes. The improvement of OCM activity in Li2CaSiO4 is likely to result from the combination of multiple cations in the crystal lattice. The structure of Li2CaSiO4 consists of a single oxygen site, neighboring with Li, Ca and Si (two Li, two Ca, and one Si atoms). The resulting coordination around oxygen provides a dual character such as strong basicity and lattice stability, which are mainly responsible for CH4 conversion and C2 selectivity, respectively. Elkins et al. [49] reported two different types of metal oxide catalysts such as Li-doped TbOx and Sm2O3 supported on MgO and they found that the Li-Tb2O3/n-MgO catalyst exhibited high activity and selectivity towards desired products at low reaction temperature i.e 650 °C. In addition, the catalyst showed higher stability and lower deactivation rate after 30 h on stream, because the Li addition prompted the reduction of TbOx phase to Tb2O3 and also shifted the Tb 3d core-level electrons to higher binding energy that caused strong Li-TbOx interactions, which in turn became more active toward C2 product formation than the SmOx. Table 1 summaries the available data in literature on various Li doped/impregnated catalysts for OCM.
The Li-supported systems, especially Li-MgO catalysts, are among the most studied OCM Catalysts. Because they show a C2 yield of 20% with higher methane conversion rates at lower temperatures, however, these catalysts suffer from deactivation due to Li vaporization. [Li+O] is specified as the active site for the formation of methyl radical intermediate in OCM reaction, even though many features are still unclear, such as stability of the catalyst, structure activity relationship, and active center; hence, proficient materials are needed to stabilize Li to overcome these issues, and these systems must be stable at high temperatures, where OCM usually occurs. Thus, the synthesis of efficient and Li stable catalysts for OCM is highly challenging. To overcome these issues, in order to improve the catalytic stability of these catalysts, various modifications have been made during the last few decades via the addition of rare-earth oxides or doping with earth alkaline metals or through coordination of multiple ions. As per the previous results, we understand that the mixed-phase oxides have multiple elements with an interface that are believed to show higher activity and stability due to the synergistic effect between the metal oxides. For example, when used Li doped TbOx and Sm2O3 supported on MgO and they found that the Li-Tb2O3/n-MgO catalyst exhibited higher stability and lower deactivation rate after 30 h on stream [49], while Li2CaSiO4 showed the better catalytic activity and selectivity over Mn-Na2WO4/SiO2 and also exhibited higher catalyst durability for 50 h on stream without deactivation and structural changes for the OCM reaction [48].

2.2. Modified Transition Metal Oxide Catalysts

Mn-NaWO4/SiO2 catalyst is widely accepted as the most applicable catalyst, with a higher C2 yield (~27%) in OCM [33,52,53,54,55,56,57,58,59,60]. Li S-B. et al., J. H. Lunsford et al. and R. M. Lambert et al. groups established Mn-NaWO4/SiO2 and the optimized chemical composition of this catalyst is 1.9 wt% Mn-5 wt% NaWO4/SiO2 with Na: W: Mn atom ratio of 2:1:2 [55,61,62]. Ghose et al. [54] prepared different nanostructured complex metal oxides such as Sr–Al complex oxides, La2O3, La–Sr–Al complex oxides, and Na2WO4-Mn/SiO2 by using the solution combustion synthesis (SCS) method. Among these catalysts, Sr3Al2O6 (double perovskite phase) in the Sr–Al oxides is active for OCM. The La2O3 catalyst showed the highest C2 yields (~13.5%) as compared to similar catalysts presented in the literature. However, the addition of La exhibited a higher C2 yield even at lower temperatures of 750 °C. The Na2WO4-Mn/SiO2 catalyst was stable and showed a higher C2 yield (25%) in OCM, which is one of the best results in the literature. Wang et al. [57] carefully explored the effect of Na2WO4-Mn/SiO2 catalysts prepared by different methods such as incipient wetness impregnation (IWI), mixture slurry and the sol-gel method for OCM reaction. As per the different analytical methods, IWI showed that Na, W and Mn are mainly distributed on the catalyst’s surface. However, the other two ways are produced more uniform between the surface and bulk on the catalysts. Hence, the mixture slurry method has excellent stability and is stable during 500 h on stream at 800 °C in the OCM, and during the stream, methane conversion, and C2 selectivity are maintained at 27–31% and 68–71%, respectively. Yunarti et al. [58] developed promising oxide composite material Na2WO4/Mn/Mg0.05Ti0.05Si0.90On, prepared using a one-pot sol-gel method. This catalyst showed a 23.1% C2 yield at 800 °C, an 18% higher C2 selectivity, and a 35% higher C2 yield than the conventional Na2WO4/Mn/SiO2 catalyst, presumably due to Mg and Ti, which are doped in the α-cristobalite SiO2. These ions can inhibit incorporating Na/W/Mn compounds into the α-cristobalite SiO2 structure. As a result, Na/W/Mn compounds are exposed to the catalyst’s surface, which is more accountable for the higher C2 selectivity and yield in the OCM reaction. The catalytic results of Mn-NaWO4/SiO2 indicate synergistic interactions between the various catalyst components (Na, Mn and W), as displayed in Figure 5.
If the catalyst contains only Na or W-oxide, it is inactive and unselective. However, catalysts consisting only of Mn-oxide are quite active towards methane combustion. Na and W-oxide catalysts doping enhance their conversion and C2 selectivity, presumably due to synergy between Na and W components with structural effects and the introduction of surface basic sites. The addition of Na makes it easy to reduce the W-oxide component, which has a positive impact on OCM. Catalysts having both Mn and W oxides showed better C2 selectivity but were slightly less active. The bi-component system (Na and Mn oxide) is 100% selective but has poor conversion, while, the three-component system (Mn-Na-W oxide supported on SiO2) shows significantly higher selectivity than bi-component (Na-W, Na-Mn and, and Mn-W) catalysts, although it is slightly less active than the Na-W oxide catalyst. Each component has a specific role in the reaction, and the Na species are generally known to convert SiO2 support from its primary amorphous phase to crystalline SiO2 (α-cristobalite) phase, which plays an important role in the OCM reaction and also favors in the migration of Mn and W species to the surface of the catalyst [55,56,63]. The common role of Mn is promoting oxygen mobility between surface-adsorbed and lattice oxygen atoms. The surface Mn sites are responsible for OCM activity and C2 hydrocarbon selectivity [53]. OCM reaction mechanism for the supported Mn2O3/Na2WO4/SiO2 catalyst is shown in Figure 6.
Ortiz-Bravo et al. [64] systematically elucidate the electronic and molecular structure of the W and Mn sites on the Mn-Na2WO4/SiO2 catalyst of the corresponding OCM reaction temperature by using different analytical techniques. The Mn-Na2WO4/SiO2 catalyst’s structure is highly temperature-dependent; thus, the association of any OCM activity with crystalline phases observed at room temperature is inadequate. In situ TPO- XRD technique clearly displays that the crystalline phases presented at room temperature in the Mn-Na2WO4/SiO2, Na2WO4/SiO2 and WO3/SiO2 catalysts are absent at OCM temperatures (>700 °C). Upon heating in oxidizing conditions, Na and W-oxide’s crystalline phases altered to α to β then γ-WO3; cubic to orthorhombic, then molten Na2WO4; and the support SiO2 phase can be changed α to β-cristobalite. TPO-Raman spectra clearly validate that the bond order of W sites with tetrahedral (Td) and octahedral (Oh) symmetry varies during the phase transformation. Because all samples retain essentially W6+ valence and Oh-Mn3+ sites are always present on Mn-Na2WO4/SiO2 catalyst, TPO-XANES spectra expose that bond order differences are due to distortion degree variations. Finally, they established that Oh-W6+ sites are inactive in the steady-state OCM tests, but Td-W6+ sites are more active in the presence of Oh-Mn3+ sites, which can be helpful for the activation of methane. Table 2 lists the available data in the literature on MnOx-Na2WO4/SiO2 catalysts for OCM.
Due to their excellent reaction and thermal stability, various supports such as Mg-Ti mixed oxides, SiC, COK-12, SBA-15 were used instead of SiO2 support to improve the catalyst’s performance [60,68,69,70]. In addition, the promotion of Mn/Na2WO4/SiO2 with other metal oxides such as La2O3, SnO2, and CeO2 have also been conducted in OCM [71,72,73]. Aimed to optimize the catalytic material for OCM, Sun et al. [74] prepared a series of CexZr1−xO2 catalysts using a sol-gel method with varied Ce/Zr molar ratios and then modified with Mn2O3-Na2WO4 doping. Among them, Ce0.15Zr0.85O2 containing catalyst exhibited better conversion (25%) and selectivity (67%) even at 660 °C due to generate more O2 species, which are responsible for enhanced activity and selectivity at a lower reaction temperature. Moreover, this catalyst showed stable catalytic performance up to 100 h at 660 °C without decreasing the methane conversion (25%) and C2-C3 selectivity (67%). Recently, Geo Jong Kim et al. proved that TiO2 support showed better performance than SiO2 in Mn/Na2WO4-based catalysts. The authors have modified the conventional OCM catalyst (Mn-NaWO4/SiO2, MNWSi) with TiO2 (MNWTi) as support and CeO2 (MNWCeTi) as a promoter. Here they varied the reaction parameters and compared the activities with conventional catalyst. They found that MNWTi catalyst improved methane conversion and C2 yield. The methane conversion and C2 yield of this catalyst enhanced from 28.5 to 46.0% and from 12.1 to 25.9% at 775 °C, respectively, as compared to with conventional OCM catalyst, presumably due to having MnTiO3 and Mn2O3, with higher Mn species on the surface which are helpful to activate the O2 at low temperature [75]. Temperatures higher than 700 °C are usually required to initiate this reaction, and higher C2 yields could be achieved around 800 °C [52,76]. Therefore, catalyst stability and selectivity of hydrocarbons are limited due to the suppression of active sites at very high temperatures and increased non-selective oxidation products. Therefore, there is still a strong motivation to explore novel catalysts with high reaction performance, particularly at low temperatures, to execute this reaction industrially.

2.3. Pyrochlore Catalysts

Pyrochlore (A2B2O7) catalysts are promising catalysts for OCM reaction due to their high melting points, thermal stability, tunable M-O bonding, surface alkalinity, and oxygen vacancies [77,78,79,80,81,82]. The crystal structure of A2B2O7 compounds can be tuned with the ionic radius ratio of rA and rB sites cations. In detail, pyrochlore structure exists when rA/rB is between 1.46 and 1.78, where A and B are distributed in order. The disordered defective fluorite phase formed when rA/rB is less than 1.46 and the arrangement of A and B ions are disordered; if rA/rB is higher than 1.78, the monoclinic phase will be formed [77,78]. The crystal structure diagram of pyrochlore and defective fluorite can be seen in Figure 7.
The ideal pyrochlore (A2B2O7) structure belongs to the space group Fd-3m. It can be written as A2B2O6 O’ form, and the formula denotes the existence of two different types of lattice oxygen ions in its crystal structure. Each lattice oxygen ion (O) is coordinated to two A and B sites cations, and the left lattice oxygen ion (O’) is coordinated to four A sites cation. Compared to the ordered cubic fluorite structure, the pyrochlore compounds contain less of a lattice oxygen ion thus forming inherent oxygen vacancies, which increases the oxygen mobility of the pyrochlore catalysts [80,83,84]. Moreover, a typical pyrochlore (A2B2O7) compound containing rare-earth A- and B-site can offer surface alkalinity. In addition, all the characteristics such as thermal stability, surface alkalinity and inherent oxygen vacancies can be altered by replacing or partly substituting of the A and B sites with alkali or alkaline earth metals [85].
Based on these properties, the A2B2O7 type of materials has shown some promising catalytic properties in OCM reaction. Petit et al. [81] reported that the catalytic activity of pyrochlore (A2B2O7; A—rare earth metal, B—Ti, Sn, Zr) on OCM mainly depended on the B-O bond strength. They found that lower B-O binding energy compounds provided higher C2 yield. C. A. Mims et al., [86] prepared Bi-Sn pyrochlore catalysts and then expanded the series of Bi2Sn2−xBixO7−x/2 (0 ≤ x ≤ 0.86) with substitution of Bi cation in the B site of the pyrochlore. These catalysts were used for OCM and have shown improvement in C2 selectivity as the number of Bi cations at B sites increased. In addition, A. C. Roger et al. [87] synthesized Sn-deficient pyrochlore (Sm2Sn2O7) catalysts by using a sol-gel method, and these catalysts improved the C2 yield in the OCM reaction.
Recently, Wang and co-workers have developed different types of pyrochlore catalysts and examined their properties’, such as rA/rB ratios, effect on crystalline structures, surface active oxygen sites, intrinsic oxygen vacancies and surface alkaline sites, being essential for OCM reaction [77,78,85]. A series of Ln2Ce2O7 (Ln = La, Pr, Sm and Y) catalysts have been developed and demonstrated as the defective cubic fluorite phase. Among them, La2Ce2O7 exhibited the highest C2 yield of 16.6% at 750 °C due to the moderated alkaline sites and surface active oxygen species [77]. To examine the relationship between phase structure and reactivity of pyrochlore catalysts for the OCM, three model La2B2O7 (B-Ti, Zr and Ce) pyrochlore compounds with different crystal phases have been prepared. The crystalline phase of La2B2O7 differs from monoclinic layered perovskite (La2Ti2O7) to ordered cubic pyrochlore (La2Zr2O7) and defective cubic fluorite (La2Ce2O7) by declining the rA/rB ratios. These catalytic activities and C2 yields follow the order of La2Ce2O7 > La2Zr2O7 > La2Ti2O7, being consistent with higher surface active oxygen species and moderate surface alkaline sites [78]. Aimed to optimize the La2Ce2O7 catalytic material for OCM, a series of catalysts were fabricated by a sol-gel method with varied La/Ce molar ratios and being doped with Ca, which had a close ionic radius with La and Ce ions. Among them, La2Ce1.5Ca0.5O7 showed better yields (22.5%) at 750 °C due to its enhanced surface alkalinity and oxygen mobility with doping of Ca additive [85]. Table 3 lists the available data in the literature on pyrochlore catalysts for OCM.
La2Ce2O7 pyrochlore catalysts have been synthesized in our group using the sol-gel method and further doped with Sr through optimization on Sr loading [89]. The defective cubic fluorite phase remained after Sr doping. Introduction of Sr in La2Ce2O7, especially allowing strontium to enter into the crystal lattice (La1.5Sr0.5Ce2O7), significantly improved the mobility of lattice oxygen compared to the undoped pyrochlore (La2Ce2O7) and impregnated Sr sample (8% Sr/La2Ce2O7) (Figure 8). The relative contents of O2− lattice oxygen of the three prepared catalysts, i.e., doped, impregnated and undoped, followed the order of La1.5Sr0.5Ce2O7 > 8% Sr/La2Ce2O7 > La2Ce2O7 as shown in XPS results (see Figure 9, Table 4); these obtained results were in line with H2-TPR results and the OCM reaction performance. A sample containing a large amount of lattice oxygen sites showed better reaction performance (see Figure 11). CO2-TPD results (see Figure 10, Table 5) revealed that introduction of Sr enhanced the ratio of strong basic sites, which is usually considered as a beneficial factor for producing C2 products. As a result, the selectivity and yield of C2-based products could reach 57% and 14%, respectively, over La1.5Sr0.5Ce2O7 catalysts at 800 °C (see Figure 11). Moreover, this catalyst showed stable catalytic performance up to 30 h without deactivation (see Figure 12).
In another study, we substitute the B site of La2Ce2O7 by Ca2+ (0.1 nm) and Sr2+ (0.126 nm), respectively, which have an identical atomic radius to Ce4+ (0.097 nm). These catalysts showed better selectivity and yield than host compounds. XRD results (Figure 13) reveal that all doped samples presented well-resolved peaks analogous to their host pyrochlore materials with only a slight shift of the 2θ values, signifying those doped compounds have similar phase structures to their host compounds. The quantitative analysis of H2-TPR results (see Figure 14, Table 6) revealed that the reduction temperature and reducible lattice oxygen species of doped samples were significantly enhanced than the un-doped sample; consequently, this rise in the reduction temperature indicated higher M-O bonding strength. Among these doped catalysts, La2Ce1.5Sr0.5O7 showed a higher reduction temperature and contained a larger amount of lattice oxygen, which was primarily responsible for the higher C2 selectivity in the OCM reaction. Therefore, this catalyst exhibited better yield at 800 °C (see Figure 15).

3. Conclusions and Perspective

Natural gas has attracted considerable attention because of fluctuating fuel prices around the world and continuous depletion of energy reserves of crude oil, and also due to its unique features such as a clean source of fossil energy and as a feedstock of various other chemicals including ethylene. Methane is a major component of natural gas, and its deposits are expected to be higher than crude oil in the coming future. Ethylene is considered as a key building block in the field of industrial chemical production. Therefore, a strong economic interest is developing in the processes that allow the conversion of methane to high value-added products such as numerous intermediate products such as polymers, i.e., polyethylene, polystyrene; and hydrocarbons i.e., ethylene. In this regard, oxidative coupling methane (OCM) is an important reaction process for the catalytic upgrading of methane in natural gas and/or shale gas to ethylene for meeting its industrial production demand. This review article mainly consolidated recent literature that captured and evaluated the modifications during the catalyst development progress made regarding the OCM reaction.
The development of commercially viable catalysts for OCM reaction is still a crucial challenge for researchers. In order to obtain the desired goal, recently, various catalysts have been used based on the metal oxides such as reducible metal oxides, rare earth oxides, alkaline earth oxides, alkali-promoted metal oxides or mixed oxides, and pyrochlore compounds individually or in various combinations. In the current review, we have explored a variety of catalysts such as alkali/alkaline earth metals-based catalysts viz., Li supported catalysts, Mn-NaWO4/SiO2 and pyrochlore catalysts. It has been reviewed that catalyst performance depends on several factors such as their preparation method, structural composition and the basicity of catalyst. From the reports, we have understood that active sites, facile oxygen abundance, surface alkalinity and metal-support interactions are crucial for catalytic performance. However, by using these catalysts, a low yield of C2 products still was achieved (<30%) and could not match the commercial requirement.

4. Perspective

Li-based catalysts have shown higher activity and selectivity in OCM reaction due to strong basicity and M-O bond. However, the catalysts suffer deactivation during the reaction, being related to the loss of Li at high operation temperatures. Therefore, how to stabilize Li and restrain its evaporation at high temperatures by forming strong metal support interactions might be a possible solution for this. Meanwhile, it is noted that pyrochlore (A2B2O7) catalysts have shown some promising properties for OCM reaction due to high melting points, thermal stability, tunable M-O bonding and moderate surface alkalinity. Using pyrochlore as the stable support to load Li might generate strong interactions between Li and pyrochlore, forming new phases of Li compounds, such as Li2SnO3 and LiLaO2, which provides a feasible way to stabilize Li. Some preliminary results have been obtained in our group to show its promising perspectives. In addition, optimization of experimental conditions for the synthesis of pyrochlore catalysts, especially for those with specific morphology, might improve the catalytic performance at lower temperatures. Discrimination of the key factors caused by A/B regulation or substitution and their interactions with the doped Li give the opportunity to further enhance the catalytic activity and selectivity as well as the stability by stabilizing Li with optimized pyrochlore.

Author Contributions

Conceptualization, Methodology, Investigation, Writing Review & Editing, Visualization, P.R.M.; Investigation, Methodology, Review & Editing, Visualization, Y.L.; Reviewing and Editing, G.W.; Reviewing and Editing; Y.D.; Conceptualization, Supervision, Writing—Reviewing and Editing, Funding acquisition, C.S. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by the National Natural Science Foundation of China (Nos. 21872014, 21932002 and 21902018), the National Key R&D Program of China (No. 2017YFA0700103), the Liaoning Revitalization Talent Program (XLYC2008032) and the Fundamental Research Funds for the Central Universities (DUT20ZD205).

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Sattler, J.J.H.B.; Ruiz-Martinez, J.; Santillan-Jimenez, E.; Weckhuysen, B.M. Catalytic dehydrogenation of light alkanes on metals and metal oxides. Chem. Rev. 2014, 114, 10613–10653. [Google Scholar] [CrossRef] [PubMed]
  2. Wood, D.A.; Nwaoha, C.; Towler, B.F. Gas-to-liquids (GTL): A review of an industry offering several routes for monetizing natural gas. J. Nat. Gas Sci. Eng. 2012, 9, 196–208. [Google Scholar] [CrossRef]
  3. Galadima, A.; Muraza, O. Revisiting the oxidative coupling of methane to ethylene in the golden period of shale gas: A review. J. Ind. Eng. Chem. 2016, 37, 1–13. [Google Scholar] [CrossRef]
  4. Guo, B.; Shan, J.; Feng, Y. Productivity of blast-fractured wells in liquid-rich shale gas formations. J. Nat. Gas Sci. Eng. 2014, 18, 360–367. [Google Scholar] [CrossRef]
  5. Seifzadeh Haghighi, S.; Rahimpour, M.R.; Raeissi, S.; Dehghani, O. Investigation of ethylene production in naphtha thermal cracking plant in presence of steam and carbon dioxide. Chem. Eng. J. 2013, 228, 1158–1167. [Google Scholar] [CrossRef]
  6. Le Van Mao, R.; Yan, H.; Muntasar, A.; Al-Yassir, N. Chapter 7—Blending of non-petroleum compounds with current hydrocarbon feeds to use in the thermo-catalytic steam-cracking process for the selective production of light olefins. In New and Future Developments in Catalysis; Suib, S.L., Ed.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 143–173. [Google Scholar]
  7. Keyvanloo, K.; Towfighi, J.; Sadrameli, S.M.; Mohamadalizadeh, A. Investigating the effect of key factors, their interactions and optimization of naphtha steam cracking by statistical design of experiments. J. Anal. Appl. Pyrolysis 2010, 87, 224–230. [Google Scholar] [CrossRef]
  8. Corma, A.; Mengual, J.; Miguel, P.J. IM-5 zeolite for steam catalytic cracking of naphtha to produce propene and ethene. An alternative to ZSM-5 zeolite. Appl. Catal. A Gen. 2013, 460–461, 106–115. [Google Scholar] [CrossRef]
  9. Dry, M.E. The Fischer–Tropsch process: 1950–2000. Catal. Today 2002, 71, 227–241. [Google Scholar] [CrossRef]
  10. Schulz, H. Short history and present trends of Fischer–Tropsch synthesis. Appl. Catal. A Gen. 1999, 186, 3–12. [Google Scholar] [CrossRef]
  11. Tian, P.; Wei, Y.; Ye, M.; Liu, Z. Methanol to olefins (MTO): From fundamentals to commercialization. ACS Catal. 2015, 5, 1922–1938. [Google Scholar] [CrossRef]
  12. Olsbye, U.; Svelle, S.; Bjørgen, M.; Beato, P.; Janssens, T.V.W.; Joensen, F.; Bordiga, S.; Lillerud, K.P. Conversion of methanol to hydrocarbons: How zeolite cavity and pore size controls product selectivity. Angew. Chem. Int. Ed. 2012, 51, 5810–5831. [Google Scholar] [CrossRef]
  13. Guo, X.; Fang, G.; Li, G.; Ma, H.; Fan, H.; Yu, L.; Ma, C.; Wu, X.; Deng, D.; Wei, M.; et al. Direct, nonoxidative conversion of methane to ethylene, aromatics, and hydrogen. Science 2014, 344, 616–619. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, M.; Ke, S.; Wu, H.; Xia, W.; Wan, H. Flower-like Sr-La2O3 microspheres with hierarchically porous structures for oxidative coupling of methane. Ind. Eng. Chem. Res. 2019, 58, 22847–22856. [Google Scholar] [CrossRef]
  15. Lopes, L.B.; Vieira, L.H.; Assaf, J.M.; Assaf, E.M. Effect of Mg substitution on LaTi1−xMgxO3+δ catalysts for improving the C2 selectivity of the oxidative coupling of methane. Catal. Sci. Technol. 2021, 11, 283–296. [Google Scholar] [CrossRef]
  16. Sato, A.; Ogo, S.; Kamata, K.; Takeno, Y.; Yabe, T.; Yamamoto, T.; Matsumura, S.; Hara, M.; Sekine, Y. Ambient-temperature oxidative coupling of methane in an electric field by a cerium phosphate nanorod catalyst. Chem. Commun. 2019, 55, 4019–4022. [Google Scholar] [CrossRef] [PubMed]
  17. Schmack, R.; Friedrich, A.; Kondratenko, E.V.; Polte, J.; Werwatz, A.; Kraehnert, R. A meta-analysis of catalytic literature data reveals property-performance correlations for the OCM reaction. Nat. Commun. 2019, 10, 441. [Google Scholar] [CrossRef]
  18. Pirro, L.; Mendes, P.S.F.; Kemseke, B.; Vandegehuchte, B.D.; Marin, G.B.; Thybaut, J.W. From catalyst to process: Bridging the scales in modeling the OCM reaction. Catal. Today 2021, 365, 35–45. [Google Scholar] [CrossRef]
  19. 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. [Google Scholar] [CrossRef]
  20. Grant, J.T.; Venegas, J.M.; McDermott, W.P.; Hermans, I. Aerobic Oxidations of Light Alkanes over Solid Metal Oxide Catalysts. Chem. Rev. 2018, 118, 2769–2815. [Google Scholar] [CrossRef]
  21. Lomonosov, V.I.; Sinev, M.Y. Oxidative coupling of methane: Mechanism and kinetics. Kinet. Catal. 2016, 57, 647–676. [Google Scholar] [CrossRef]
  22. Keller, G.E.; Bhasin, M.M. Synthesis of ethylene via oxidative coupling of methane: I. Determination of active catalysts. J. Catal. 1982, 73, 9–19. [Google Scholar] [CrossRef]
  23. Ito, T.; Lunsford, J.H. Synthesis of ethylene and ethane by partial oxidation of methane over lithium-doped magnesium oxide. Nature 1985, 314, 721–722. [Google Scholar] [CrossRef]
  24. Jiang, T.; Song, J.; Huo, M.; Yang, N.; Liu, J.; Zhang, J.; Sun, Y.; Zhu, Y. La2O3 catalysts with diverse spatial dimensionality for oxidative coupling of methane to produce ethylene and ethane. RSC Adv. 2016, 6, 34872–34876. [Google Scholar] [CrossRef]
  25. Papa, F.; Luminita, P.; Osiceanu, P.; Birjega, R.; Akane, M.; Balint, I. Acid–base properties of the active sites responsible for C2+ and CO2 formation over MO–Sm2O3 (M = Zn, Mg, Ca and Sr) mixed oxides in OCM reaction. J. Mol. Catal. A Chem. 2011, 346, 46–54. [Google Scholar] [CrossRef]
  26. Wang, P.; Zhao, G.; Wang, Y.; Lu, Y. MnTiO3-driven low-temperature oxidative coupling of methane over TiO2-doped Mn2O3-Na2WO4/SiO2 catalyst. Sci. Adv. 2017, 3, e1603180. [Google Scholar] [CrossRef] [Green Version]
  27. Bai, L.; Polo-Garzon, F.; Bao, Z.; Luo, S.; Moskowitz, B.M.; Tian, H.; Wu, Z. Impact of Surface Composition of SrTiO3 Catalysts for Oxidative Coupling of Methane. ChemCatChem 2019, 11, 2107–2117. [Google Scholar] [CrossRef]
  28. Lee, J.S.; Oyama, S.T. Oxidative Coupling of Methane to Higher Hydrocarbons. Catal. Rev. 1988, 30, 249–280. [Google Scholar] [CrossRef]
  29. Zavyalova, U.; Holena, M.; Schlögl, R.; Baerns, M. Statistical analysis of past catalytic data on oxidative methane coupling for new insights into the composition of high-performance catalysts. ChemCatChem 2011, 3, 1935–1947. [Google Scholar] [CrossRef]
  30. Horn, R.; Schlögl, R. Methane activation by heterogeneous catalysis. Catal. Lett. 2015, 145, 23–39. [Google Scholar] [CrossRef] [Green Version]
  31. Voskresenskaya, E.N.; Roguleva, V.G.; Anshits, A.G. Oxidant activation over structural defects of oxide catalysts in oxidative methane coupling. Catal. Rev. 1995, 37, 101–143. [Google Scholar] [CrossRef]
  32. Ferreira, V.J.; Tavares, P.; Figueiredo, J.L.; Faria, J.L. Effect of Mg, Ca, and Sr on CeO2 based catalysts for the oxidative coupling of methane: Investigation on the oxygen species responsible for catalytic performance. Ind. Eng. Chem. Res. 2012, 51, 10535–10541. [Google Scholar] [CrossRef]
  33. Yıldız, M. Mesoporous TiO2-rutile supported MnxOy-Na2WO4: Preparation, characterization and catalytic performance in the oxidative coupling of methane. J. Ind. Eng. Chem. 2019, 76, 488–499. [Google Scholar] [CrossRef]
  34. Ivanov, D.V.; Isupova, L.A.; Gerasimov, E.Y.; Dovlitova, L.S.; Glazneva, T.S.; Prosvirin, I.P. Oxidative methane coupling over Mg, Al, Ca, Ba, Pb-promoted SrTiO3 and Sr2TiO4: Influence of surface composition and microstructure. Appl. Catal. A Gen. 2014, 485, 10–19. [Google Scholar] [CrossRef]
  35. Peng, L.; Xu, J.; Fang, X.; Liu, W.; Xu, X.; Liu, L.; Li, Z.; Peng, H.; Zheng, R.; Wang, X. SnO2 based catalysts with low-temperature performance for oxidative coupling of methane: Insight into the promotional effects of alkali-metal oxides. Eur. J. Inorg. Chem. 2018, 2018, 1787–1799. [Google Scholar] [CrossRef]
  36. Cong, L.; Zhao, Y.; Li, S.; Sun, Y. Sr-doping effects on La2O3 catalyst for oxidative coupling of methane. Chin. J. Catal. 2017, 38, 899–907. [Google Scholar] [CrossRef]
  37. Kim, I.; Lee, G.; Na, H.B.; Ha, J.-M.; Jung, J.C. Selective oxygen species for the oxidative coupling of methane. Mol. Catal. 2017, 435, 13–23. [Google Scholar] [CrossRef]
  38. Elkins, T.W.; Roberts, S.J.; Hagelin-Weaver, H.E. Effects of alkali and alkaline-earth metal dopants on magnesium oxide supported rare-earth oxide catalysts in the oxidative coupling of methane. Appl. Catal. A Gen. 2016, 528, 175–190. [Google Scholar] [CrossRef]
  39. Senanayake, S.D.; Rodriguez, J.A.; Weaver, J.F. Low temperature activation of methane on metal-oxides and complex interfaces: Insights from surface science. Acc. Chem. Res. 2020, 53, 1488–1497. [Google Scholar] [CrossRef]
  40. Schröder, D.; Roithová, J. Low-temperature activation of methane: It also works without a transition metal. Angew. Chem. Int. Ed. 2006, 45, 5705–5708. [Google Scholar] [CrossRef]
  41. Farsi, A.; Moradi, A.; Ghader, S.; Shadravan, V. Oxidative coupling of methane over Li/MgO: Catalyst and nanocatalyst performance. Chin. J. Chem. Phys. 2011, 24, 70–76. [Google Scholar] [CrossRef]
  42. Qian, K.; You, R.; Guan, Y.; Wen, W.; Tian, Y.; Pan, Y.; Huang, W. Single-Site catalysis of Li-MgO catalysts for oxidative coupling of methane reaction. ACS Catal. 2020, 10, 15142–15148. [Google Scholar] [CrossRef]
  43. Kwapien, K.; Paier, J.; Sauer, J.; Geske, M.; Zavyalova, U.; Horn, R.; Schwach, P.; Trunschke, A.; Schlögl, R. Sites for methane activation on lithium-doped magnesium oxide surfaces. Angew. Chem. Int. Ed. 2014, 53, 8774–8778. [Google Scholar] [CrossRef]
  44. Schwach, P.; Frandsen, W.; Willinger, M.-G.; Schlögl, R.; Trunschke, A. Structure sensitivity of the oxidative activation of methane over MgO model catalysts: I. Kinetic study. J. Catal. 2015, 329, 560–573. [Google Scholar] [CrossRef] [Green Version]
  45. Berger, T.; Schuh, J.; Sterrer, M.; Diwald, O.; Knözinger, E. Lithium ion induced surface reactivity changes on MgO nanoparticles. J. Catal. 2007, 247, 61–67. [Google Scholar] [CrossRef]
  46. Ito, T.; Wang, J.; Lin, C.H.; Lunsford, J.H. Oxidative dimerization of methane over a lithium-promoted magnesium oxide catalyst. J. Am. Chem. Soc. 1985, 107, 5062–5068. [Google Scholar] [CrossRef]
  47. Luo, L.; Jin, Y.; Pan, H.; Zheng, X.; Wu, L.; You, R.; Huang, W. Distribution and role of Li in Li-doped MgO catalysts for oxidative coupling of methane. J. Catal. 2017, 346, 57–61. [Google Scholar] [CrossRef]
  48. Matsumoto, T.; Saito, M.; Ishikawa, S.; Fujii, K.; Yashima, M.; Ueda, W.; Motohashi, T. High catalytic activity of crystalline lithium calcium silicate for oxidative coupling of methane originated from crystallographic joint effects of multiple cations. ChemCatChem 2020, 12, 1968–1972. [Google Scholar] [CrossRef]
  49. Elkins, T.W.; Neumann, B.; Bäumer, M.; Hagelin-Weaver, H.E. Effects of Li Doping on MgO-Supported Sm2O3 and TbOx Catalysts in the Oxidative Coupling of Methane. ACS Catal. 2014, 4, 1972–1990. [Google Scholar] [CrossRef]
  50. Lunsford, J.H. The Catalytic oxidative coupling of methane. Angew. Chem. Int. Ed. 1995, 34, 970–980. [Google Scholar] [CrossRef]
  51. Raouf, F.; Taghizadeh, M.; Yousefi, M. Influence of CaO–ZnO supplementation as a secondary catalytic bed on the oxidative coupling of methane. React. Kinet. Mech. Catal. 2014, 112, 227–240. [Google Scholar] [CrossRef]
  52. Arndt, S.; Otremba, T.; Simon, U.; Yildiz, M.; Schubert, H.; Schomäcker, R. Mn–Na2WO4/SiO2 as catalyst for the oxidative coupling of methane. What is really known? Appl. Catal. A Gen. 2012, 425–426, 53–61. [Google Scholar] [CrossRef]
  53. Kiani, D.; Sourav, S.; Baltrusaitis, J.; Wachs, I.E. Oxidative Coupling of methane (OCM) by SiO2-supported tungsten oxide catalysts promoted with Mn and Na. ACS Catal. 2019, 9, 5912–5928. [Google Scholar] [CrossRef]
  54. Ghose, R.; Hwang, H.T.; Varma, A. Oxidative coupling of methane using catalysts synthesized by solution combustion method: Catalyst optimization and kinetic studies. Appl. Catal. A Gen. 2014, 472, 39–46. [Google Scholar] [CrossRef]
  55. Palermo, A.; Holgado Vazquez, J.P.; Lee, A.F.; Tikhov, M.S.; Lambert, R.M. Critical influence of the amorphous silica-to-cristobalite phase transition on the performance of Mn/Na2WO4/SiO2 catalysts for the oxidative coupling of methane. J. Catal. 1998, 177, 259–266. [Google Scholar] [CrossRef]
  56. Elkins, T.W.; Hagelin-Weaver, H.E. Characterization of Mn–Na2WO4/SiO2 and Mn–Na2WO4/MgO catalysts for the oxidative coupling of methane. Appl. Catal. A Gen. 2015, 497, 96–106. [Google Scholar] [CrossRef]
  57. Wang, J.; Chou, L.; Zhang, B.; Song, H.; Zhao, J.; Yang, J.; Li, S. Comparative study on oxidation of methane to ethane and ethylene over Na2WO4–Mn/SiO2 catalysts prepared by different methods. J. Mol. Catal. A Chem. 2006, 245, 272–277. [Google Scholar] [CrossRef]
  58. Yunarti, R.T.; Gu, S.; Choi, J.-W.; Jae, J.; Suh, D.J.; Ha, J.-M. Oxidative coupling of methane using Mg/Ti-doped SiO2-supported Na2WO4/Mn Catalysts. ACS Sustain. Chem. Eng. 2017, 5, 3667–3674. [Google Scholar] [CrossRef]
  59. Ghose, R.; Hwang, H.T.; Varma, A. Oxidative coupling of methane using catalysts synthesized by solution combustion method. Appl. Catal. A Gen. 2013, 452, 147–154. [Google Scholar] [CrossRef]
  60. Yildiz, M.; Aksu, Y.; Simon, U.; Kailasam, K.; Goerke, O.; Rosowski, F.; Schomäcker, R.; Thomas, A.; Arndt, S. Enhanced catalytic performance of MnxOy–Na2WO4/SiO2 for the oxidative coupling of methane using an ordered mesoporous silica support. Chem. Commun. 2014, 50, 14440–14442. [Google Scholar] [CrossRef] [Green Version]
  61. Li, S.-B. Oxidative Coupling of Methane over W-Mn/SiO2 Catalyst. Chin. J. Chem. 2001, 19, 16–21. [Google Scholar] [CrossRef]
  62. Wang, D.J.; Rosynek, M.P.; Lunsford, J.H. Oxidative Coupling of Methane over Oxide-Supported Sodium-Manganese Catalysts. J. Catal. 1995, 155, 390–402. [Google Scholar] [CrossRef]
  63. Sofranko, J.A.; Leonard, J.J.; Jones, C.A.; Gaffney, A.M.; Withers, H.P. Catalytic oxidative coupling of methane over sodium-promoted Mn/SiO2 and Mn/MgO. Catal. Today 1988, 3, 127–135. [Google Scholar] [CrossRef]
  64. Ortiz-Bravo, C.A.; Figueroa, S.J.A.; Portela, R.; Chagas, C.A.; Bañares, M.A.; Toniolo, F.S. Elucidating the structure of the W and Mn sites on the Mn-Na2WO4/SiO2 catalyst for the oxidative coupling of methane (OCM) at real reaction temperatures. J. Catal. 2021. [Google Scholar] [CrossRef]
  65. Xueping, F.; Shuben, L.; Jingzhi, L.; Yanlai, C. Oxidative coupling of methane on W-Mn catalysts. J. Mol. Catal. China 1992, 6, 427–433. [Google Scholar]
  66. Lee, J.Y.; Jeon, W.; Choi, J.-W.; Suh, Y.-W.; Ha, J.-M.; Suh, D.J.; Park, Y.-K. Scaled-up production of C2 hydrocarbons by the oxidative coupling of methane over pelletized Na2WO4/Mn/SiO2 catalysts: Observing hot spots for the selective process. Fuel 2013, 106, 851–857. [Google Scholar] [CrossRef]
  67. Hayek, N.S.; Khlief, G.J.; Horani, F.; Gazit, O.M. Effect of reaction conditions on the oxidative coupling of methane over doped MnOx-Na2WO4/SiO2 catalyst. J. Catal. 2019, 376, 25–31. [Google Scholar] [CrossRef]
  68. Jeon, W.; Lee, J.Y.; Lee, M.; Choi, J.-W.; Ha, J.-M.; Suh, D.J.; Kim, I.W. Oxidative coupling of methane to C2 hydrocarbons on the Mg–Ti mixed oxide-supported catalysts at the lower reaction temperature: Role of surface oxygen atoms. Appl. Catal. A Gen. 2013, 464–465, 68–77. [Google Scholar] [CrossRef]
  69. Wang, H.; Schmack, R.; Paul, B.; Albrecht, M.; Sokolov, S.; Rümmler, S.; Kondratenko, E.V.; Kraehnert, R. Porous silicon carbide as a support for Mn/Na/W/SiC catalyst in the oxidative coupling of methane. Appl. Catal. A Gen. 2017, 537, 33–39. [Google Scholar] [CrossRef]
  70. Colmenares, M.G.; Simon, U.; Yildiz, M.; Arndt, S.; Schomaecker, R.; Thomas, A.; Rosowski, F.; Gurlo, A.; Goerke, O. Oxidative coupling of methane on the Na2WO4-MnxOy catalyst: COK-12 as an inexpensive alternative to SBA-15. Catal. Commun. 2016, 85, 75–78. [Google Scholar] [CrossRef]
  71. Wu, J.; Zhang, H.; Qin, S.; Hu, C. La-promoted Na2WO4/Mn/SiO2 catalysts for the oxidative conversion of methane simultaneously to ethylene and carbon monoxide. Appl. Catal. A Gen. 2007, 323, 126–134. [Google Scholar] [CrossRef]
  72. Chou, L.; Cai, Y.; Zhang, B.; Niu, J.; Ji, S.; Li, S. Influence of SnO2-doped W-Mn/SiO2 for oxidative conversion of methane to high hydrocarbons at elevated pressure. Appl. Catal. A Gen. 2003, 238, 185–191. [Google Scholar] [CrossRef]
  73. Shahri, S.M.K.; Pour, A.N. Ce-promoted Mn/Na2WO4/SiO2 catalyst for oxidative coupling of methane at atmospheric pressure. J. Nat. Gas Chem. 2010, 19, 47–53. [Google Scholar] [CrossRef]
  74. Sun, W.; Gao, Y.; Zhao, G.; Si, J.; Liu, Y.; Lu, Y. Mn2O3-Na2WO4 doping of CexZr1−xO2 enables increased activity and selectivity for low temperature oxidative coupling of methane. J. Catal. 2021, 400, 372–386. [Google Scholar] [CrossRef]
  75. Kim, G.J.; Ausenbaugh, J.T.; Hwang, H.T. Effect of TiO2 on the Performance of Mn/Na2WO4 Catalysts in Oxidative Coupling of Methane. Ind. Eng. Chem. Res. 2021, 60, 3914–3921. [Google Scholar] [CrossRef]
  76. Ji, S.; Xiao, T.; Li, S.; Chou, L.; Zhang, B.; Xu, C.; Hou, R.; York, A.P.E.; Green, M.L.H. Surface WO4 tetrahedron: The essence of the oxidative coupling of methane over M–W–Mn/SiO2 catalysts. J. Catal. 2003, 220, 47–56. [Google Scholar] [CrossRef]
  77. Xu, J.; Peng, L.; Fang, X.; Fu, Z.; Liu, W.; Xu, X.; Peng, H.; Zheng, R.; Wang, X. Developing reactive catalysts for low temperature oxidative coupling of methane: On the factors deciding the reaction performance of Ln2Ce2O7 with different rare earth A sites. Appl. Catal. A Gen. 2018, 552, 117–128. [Google Scholar] [CrossRef]
  78. Xu, J.; Zhang, Y.; Xu, X.; Fang, X.; Xi, R.; Liu, Y.; Zheng, R.; Wang, X. Constructing La2B2O7 (B = Ti, Zr, Ce) Compounds with Three Typical Crystalline Phases for the Oxidative Coupling of Methane: The Effect of Phase Structures, Superoxide Anions, and Alkalinity on the Reactivity. ACS Catal. 2019, 9, 4030–4045. [Google Scholar] [CrossRef]
  79. Zhang, Y.; Xu, J.; Xu, X.; Xi, R.; Liu, Y.; Fang, X.; Wang, X. Tailoring La2Ce2O7 catalysts for low temperature oxidative coupling of methane by optimizing the preparation methods. Catal. Today 2020, 355, 518–528. [Google Scholar] [CrossRef]
  80. Xu, J.; Xi, R.; Xu, X.; Zhang, Y.; Feng, X.; Fang, X.; Wang, X. A2B2O7 pyrochlore compounds: A category of potential materials for clean energy and environment protection catalysis. J. Rare Earths 2020, 38, 840–849. [Google Scholar] [CrossRef]
  81. Petit, C.; Rehspringer, J.L.; Kaddouri, A.; Libs, S.; Poix, P.; Kiennemann, A. Oxidative coupling of methane by pyrochlore oxide A2B2O7 (A = rare earth, B = Ti, Zr, Sn). Relation between C2 selectivity and B–O bond energy. Catal. Today 1992, 13, 409–416. [Google Scholar] [CrossRef]
  82. Zhang, F.X.; Tracy, C.L.; Lang, M.; Ewing, R.C. Stability of fluorite-type La2Ce2O7 under extreme conditions. J. Alloy Compd. 2016, 674, 168–173. [Google Scholar] [CrossRef] [Green Version]
  83. Wang, Z.; Zhou, G.; Jiang, D.; Wang, S. Recent development of A2B2O7 system transparent ceramics. J. Adv. Ceram. 2018, 7, 289–306. [Google Scholar] [CrossRef] [Green Version]
  84. Modeshia, D.R.; Walton, R.I. Solvothermal synthesis of perovskites and pyrochlores: Crystallisation of functional oxides under mild conditions. Chem. Soc. Rev. 2010, 39, 4303–4325. [Google Scholar] [CrossRef] [PubMed]
  85. Xu, J.; Zhang, Y.; Liu, Y.; Fang, X.; Xu, X.; Liu, W.; Zheng, R.; Wang, X. Optimizing the reaction performance of La2Ce2O7-based catalysts for oxidative coupling of methane (OCM) at lower temperature by lattice doping with Ca cations. Eur. J. Inorg. Chem. 2019, 2019, 183–194. [Google Scholar] [CrossRef]
  86. Mims, C.A.; Jacobson, A.J.; Hall, R.B.; Lewandowski, J.T. Methane oxidative coupling over nonstoichiometric bismuth-tin pyrochlore catalysts. J. Catal. 1995, 153, 197–207. [Google Scholar] [CrossRef]
  87. Roger, A.C.; Petit, C.; Kiennemann, A. Effect of metallo-organic precursors on the synthesis of Sm–Sn pyrochlore catalysts: Application to the oxidative coupling of methane. J. Catal. 1997, 167, 447–459. [Google Scholar] [CrossRef]
  88. Ashcroft, A.T.; Cheetham, A.K.; Green, M.L.H.; Grey, C.P.; Vernon, P.D.F. Oxidative coupling of methane over tin-containing rare-earth pyrochlores. J. Chem. Soc. Chem. Commun. 1989, 1667–1669. [Google Scholar] [CrossRef]
  89. Wu, G.; Murthy, P.R.; Chen, B.; Zhang, X.; Shi, C.; Guo, X.; Song, C. Investigation on Sr doped La2Ce2O7 pyrochlore as effective catalyst for oxidative coupling of methane. Mod. Chem. Ind. 2021. [Google Scholar] [CrossRef]
Figure 1. Global projection of technically recoverable shale gas reserves in trillion cubic feet (TCF). The leading global countries are shown in different colors. Reproduced with permission from [3] Copyright 2016 Elsevier.
Figure 1. Global projection of technically recoverable shale gas reserves in trillion cubic feet (TCF). The leading global countries are shown in different colors. Reproduced with permission from [3] Copyright 2016 Elsevier.
Crystals 11 01011 g001
Figure 2. Diagrammatic representation of oxidative coupling of methane (OCM) reaction on a metal oxide catalyst surface. Reproduced with permission from [19]. Copyright 2018 Elsevier.
Figure 2. Diagrammatic representation of oxidative coupling of methane (OCM) reaction on a metal oxide catalyst surface. Reproduced with permission from [19]. Copyright 2018 Elsevier.
Crystals 11 01011 g002
Figure 3. Various OCM catalysts with C2 yields (≥25%). Reproduced with permission from [29]. Copyright 2011 Wiley-VCH.
Figure 3. Various OCM catalysts with C2 yields (≥25%). Reproduced with permission from [29]. Copyright 2011 Wiley-VCH.
Crystals 11 01011 g003
Figure 4. Comparison of C2 yield in OCM reaction (a) with basic strength of the doped and undoped TbOx/n-MgO catalysts (b). Redrawn from [38]. Copyright 2018 Elsevier.
Figure 4. Comparison of C2 yield in OCM reaction (a) with basic strength of the doped and undoped TbOx/n-MgO catalysts (b). Redrawn from [38]. Copyright 2018 Elsevier.
Crystals 11 01011 g004
Figure 5. Synergistic effect of SiO2-based metal oxides to CH4 conversion and selectivity to 2 (left) and C2 yield (right) for OCM reaction. Reproduced with permission from [53]. Copyright 2019 ACS.
Figure 5. Synergistic effect of SiO2-based metal oxides to CH4 conversion and selectivity to 2 (left) and C2 yield (right) for OCM reaction. Reproduced with permission from [53]. Copyright 2019 ACS.
Crystals 11 01011 g005
Figure 6. Catalytic OCM mechanism on Mn2O3/Na2WO4/SiO2. Reproduced with permission from [53]. Copyright 2019 ACS.
Figure 6. Catalytic OCM mechanism on Mn2O3/Na2WO4/SiO2. Reproduced with permission from [53]. Copyright 2019 ACS.
Crystals 11 01011 g006
Figure 7. (a) Structure of pyrochlore (1/8 unit cell), (b) Structure defective fluorite. Reproduced with permission from [83].
Figure 7. (a) Structure of pyrochlore (1/8 unit cell), (b) Structure defective fluorite. Reproduced with permission from [83].
Crystals 11 01011 g007
Figure 8. H2-TPR profile of the indicated catalysts [89].
Figure 8. H2-TPR profile of the indicated catalysts [89].
Crystals 11 01011 g008
Figure 9. XPS O 1s spectra of the indicated catalysts [89].
Figure 9. XPS O 1s spectra of the indicated catalysts [89].
Crystals 11 01011 g009
Figure 10. CO2-TPD profile of the indicated catalysts [89].
Figure 10. CO2-TPD profile of the indicated catalysts [89].
Crystals 11 01011 g010
Figure 11. OCM reaction over the indicated catalysts: (a) CH4 conversion, (b) C2 selectivity, (c) C2 yield. Reaction condition: CH4:O2:N2 = 4:1:5, WHSV = 72,000 mL·g−1·h−1 [89].
Figure 11. OCM reaction over the indicated catalysts: (a) CH4 conversion, (b) C2 selectivity, (c) C2 yield. Reaction condition: CH4:O2:N2 = 4:1:5, WHSV = 72,000 mL·g−1·h−1 [89].
Crystals 11 01011 g011
Figure 12. Stability test at 800 °C with the La1.5Sr0.5Ce2O7 catalyst. Reaction condition: CH4:O2:N2 = 4:1:5, WHSV = 72,000 mL·g−1·h−1 [89].
Figure 12. Stability test at 800 °C with the La1.5Sr0.5Ce2O7 catalyst. Reaction condition: CH4:O2:N2 = 4:1:5, WHSV = 72,000 mL·g−1·h−1 [89].
Crystals 11 01011 g012
Figure 13. (a) XRD patterns of the indicated catalysts, (b) enlarge images of the main reflection in figure-a.
Figure 13. (a) XRD patterns of the indicated catalysts, (b) enlarge images of the main reflection in figure-a.
Crystals 11 01011 g013
Figure 14. H2-TPR profile of the indicated catalysts.
Figure 14. H2-TPR profile of the indicated catalysts.
Crystals 11 01011 g014
Figure 15. OCM reaction over the La2Ce2O7 and B site substitution catalysts at 800 °C. Reaction condition: CH4:O2:N2 = 4:1:5, WHSV = 72,000 mL·g−1·h−1.
Figure 15. OCM reaction over the La2Ce2O7 and B site substitution catalysts at 800 °C. Reaction condition: CH4:O2:N2 = 4:1:5, WHSV = 72,000 mL·g−1·h−1.
Crystals 11 01011 g015
Table 1. OCM reaction over various Li supported/doped catalyst.
Table 1. OCM reaction over various Li supported/doped catalyst.
YearCatalystReaction ConditionCH4 Conv. (%)C2 Select. (%)C2 Yield (%)Ref.
19857% Li/MgOCH4:O2:N2 = 1.0: 2.0: 25.0, T = 720 °C flow = 0.83 mL s−1, 4 g catalyst37.850.319[46]
1989Li(5 wt%)/MgOCH4/O2/He = 78:8:14, T = 675 °C9.082.07.4[50]
20117%Li/MgO-IWICH4:O2:inert = 3:1:3, T = 800 °C, 0.3 g catalyst26.161.316[41]
20117%Li/MgO-SGCH4:O2:inert = 3:1:3, T = 800 °C, 0.3 g catalyst39.666.426.3[41]
2014Li-TbOx/n-MgOCH4:O2 = 4:1, T = 700 °C, *GHSV = 2400 h−1, 0.4 gcatalysts23.963.815.3[49]
2014Li-Sm2O3/n-MgOCH4:O2 = 4:1, T = 700 °C, GHSV = 2400 h−1, 0.4 g catalysts21.162.513.2[49]
20141% Li/MgOCH4/O2/N2 = 4:2:4, T = 800 °C, GHSV = 4500 h−138.035.013.3[51]
2020Li2CaSiO4CH4:O2:N2 = 1.0:0.25:8.75, T-800 °C, flow rate = 10 mL/min., 1 g catalyst30.871.822.1[48]
2020Li2CaSiO4CH4:O2:N2 = 1.0:0.25:8.75, T-800 °C, flow rate = 10 mL/min., 1-g catalyst45.657.726.3[48]
20201.3 wt.% Li/MgO8% CH4 and 4% O2 balanced with Ar, T = 750 °C, flow rate: 150 mL/min.42.046.019.3[42]
20205.6 wt.% Li/MgO8% CH4 and 4% O2 balanced with Ar, T = 750 °C, flow rate: 150 mL/min.365921.2[42]
*GHSV—gas hourly space velocity.
Table 2. OCM reaction over various Mn-Na2WO4 supported catalysts.
Table 2. OCM reaction over various Mn-Na2WO4 supported catalysts.
YearCatalystReaction ConditionCH4 Conv. (%)C2 Select. (%)C2 Yield (%)Ref.
1989Na/Mn/MgON2/CH4/air, P(CH4) = 0.5 atm; T = 825–925 °C, GHSV = 9600 h−111.3−22.467−70-[63]
19925 wt% Mn/1.9 wt.% Na2WO4/SiO2CH4/O2/N2 = 3/1/2.6, T = 800 °C, *WHSV = 36,000 mL g−1 h−136.864.923.9[65]
2013Na2WO4/Mn/SiO2 (powder)CH4/O2 = 3.5, 60 mol% N2, T = 850 °C, WHSV = 10,000 h−132.045.014.4[66]
201320%TiO2/Na2WO4/Mn/SiO2CH4/O2 = 3.5, 60 mol% N2, T = 850 °C, WHSV = 10,000 h−141.060.024.6[66]
201410% Na2WO4–5% Mn/SiO2CH4:O2 = 32:8, 10% N2, T = 800 °C, 1 g catalyst--24.0[54]
2014MnxOy–Na2WO4/SBA-15CH4:O2:N2 = 4:1:4, flow—60 mL/min, T = 750 °C, 50 mg catalyst14.070.09.8[60]
2014MnxOy–Na2WO4/SiO2CH4:O2:N2 = 4:1:4, flow—60 mL/min, T = 750 °C, 50 mg catalyst7.058.04.06[60]
2015Mn–Na2WO4/n-SiO2CH4:O2 = 4:1, flow—100 mL with N2, T-800 °C, 50 mg catalyst28.573.318.5[56]
2015Mn–Na2WO4/n-MgOCH4:O2 = 4:1, flow —100 mL with N2, T-800 °C, 50 mg catalyst5.464.93.2[56]
2019MnOx-Na2WO4/SiO2CH4:O2:N2:Ar = 4:1:1:4, T = 770 °C, GHSV 60,000 cm3 h−1g−117.670.412.4[67]
2019MnOx-Na2WO4/SiO2CH4:O2:N2:Ar = 4:1:1:4, T = 770 °C, GHSV 60,000 cm3 h−1g−123.270.716.4[67]
*WHSV—Weight hourly space velocity.
Table 3. OCM reaction over various pyrochlore catalysts.
Table 3. OCM reaction over various pyrochlore catalysts.
YearCatalystReaction ConditionCH4 Conv. (%)C2 Select. (%)C2 Yield (%)Ref.
1989La2Sn2O7CH4:O2 = 2:1 and flow rate = 27 mL/min., T = 727 °C29.7154.4[88]
1989Dy2Sn2O7CH4:O2 = 2:1 and flow rate = 27 mL/min., T = 727 °C30.421.96.6[88]
1992Sm2Sn2O7CH4/O2 = 2, T = 750 °C, 67 mg catalyst, flow = 4.51 g h−140.448.819.7[81]
1992Sm2Zr2O7CH4/O2 = 2, T = 750 °C, 67 mg catalyst, flow = 4.51 g h−128.4102.8[81]
1992Sm2Ti2O7CH4/O2 = 2, T = 750 °C, 67 mg catalyst, flow = 4.51 g h−131.922.17.04[81]
2018La2Ce2O7CH4/O2 = 4:1, T = 800 °C, WHSV = 18,000 mL h−1 g−128.358.516.6[77]
2018Pr2Ce2O7CH4/O2 = 4:1, T = 800 °C, WHSV = 18,000 mL h−1 g−11532.64.9[77]
2018Sm2Ce2O7CH4/O2 = 4:1, T = 800 °C, WHSV = 18,000 mL h−1 g−122.65111.5[77]
2018Y2Ce2O7CH4/O2 = 4:1, T = 800 °C, WHSV = 18,000 mL h−1 g−123.352.412.2[77]
2019La2Zr2O7CH4:O2:N2 = 4:1:5, T = 800 °C, WHSV = 18,000 mL h−1 g−124.250.412.2[78]
2019La2Ti2O7CH4:O2:N2 = 4:1:5, T = 800 °C, WHSV = 18,000 mL h−1 g−11832.35.8[78]
2019La2Ce1.5Ca0.5O7CH4:O2:N2 = 4:1:5, T = 800 °C, WHSV = 18,000 mL h−1 g−1327022.4[85]
2019La1.5Ca0.5Ce2O7CH4:O2:N2 = 4:1:5, T = 800 °C, WHSV = 18,000 mL h−1 g−1284713.2[85]
Table 4. H2-TPR and XPS results of the catalysts [89].
Table 4. H2-TPR and XPS results of the catalysts [89].
CatalystsHydrogen Consumption/(mL·g−1)R1/%R2/%O 1s/eVR3/%
>500 °CTotalO2−CO32−O2
La2Ce2O76.317.763.381.4528.5531.2533.137.4
8% Sr/La2Ce2O714.8216.777.188.4528.5531.1533.040.2
La1.5Sr0.5Ce2O718.1120.108.390.1528.6531.2533.243.6
R1 refers to the percentage of actual hydrogen consumption to theoretical hydrogen consumption (calculated by the stoichiometric ratio). R2 refers to the percentage of hydrogen consumption above 500 °C to the total hydrogen consumption. R3 refers to the relative ratio of O2−/(O2− + CO32− + O2).
Table 5. CO2-TPD quantification results of the catalysts [89].
Table 5. CO2-TPD quantification results of the catalysts [89].
CatalystsCO2 Desorption Ratio/%
50–350 °C350–600 °C600–900 °C
La2Ce2O747.818.234.0
8% Sr/La2Ce2O727.08.364.7
La1.5Sr0.5Ce2O720.80.978.3
Table 6. H2-TPR results of the catalysts.
Table 6. H2-TPR results of the catalysts.
CatalystsHydrogen Consumption/(mL·g−1)R1/%
>500 °CTotal
La2Ce2O76.317.7681.4
La2Ce1.5Ca0.5O710.5511.4392.30
La2Ce1.5Sr0.5O711.9912.6494.85
R1 refers to the percentage of hydrogen consumption above 500 °C to the total hydrogen consumption.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Murthy, P.R.; Liu, Y.; Wu, G.; Diao, Y.; Shi, C. Oxidative Coupling of Methane: Perspective for High-Value C2 Chemicals. Crystals 2021, 11, 1011. https://doi.org/10.3390/cryst11091011

AMA Style

Murthy PR, Liu Y, Wu G, Diao Y, Shi C. Oxidative Coupling of Methane: Perspective for High-Value C2 Chemicals. Crystals. 2021; 11(9):1011. https://doi.org/10.3390/cryst11091011

Chicago/Turabian Style

Murthy, Palle Ramana, Yang Liu, Guohao Wu, Yanan Diao, and Chuan Shi. 2021. "Oxidative Coupling of Methane: Perspective for High-Value C2 Chemicals" Crystals 11, no. 9: 1011. https://doi.org/10.3390/cryst11091011

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop