Next Article in Journal
CoMn Catalysts Derived from Hydrotalcite-Like Precursors for Direct Conversion of Syngas to Fuel Range Hydrocarbons
Next Article in Special Issue
Editorial—Special Issue “Catalysis for Energy Production”
Previous Article in Journal
Effect of Calcination Temperature on Structural Properties and Catalytic Performance of Novel Amorphous NiP/Hβ Catalyst for n-Hexane Isomerization
Previous Article in Special Issue
Nickel Phosphide Electrocatalysts for Hydrogen Evolution Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Alkali and Alkaline Earth Metals in the CO2 Methanation Reaction and the Combined Capture and Methanation of CO2

by
Anastasios I. Tsiotsias
1,
Nikolaos D. Charisiou
1,
Ioannis V. Yentekakis
2 and
Maria A. Goula
1,*
1
Laboratory of Alternative Fuels and Environmental Catalysis (LAFEC), Department of Chemical Engineering, University of Western Macedonia, GR-50100 Koila, Greece
2
Laboratory of Physical Chemistry & Chemical Processes, School of Environmental Engineering, Technical University of Crete, GR-73100 Chania, Greece
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(7), 812; https://doi.org/10.3390/catal10070812
Submission received: 7 July 2020 / Revised: 17 July 2020 / Accepted: 20 July 2020 / Published: 21 July 2020
(This article belongs to the Special Issue Catalysis for Energy Production)

Abstract

:
CO2 methanation has great potential for the better utilization of existing carbon resources via the transformation of spent carbon (CO2) to synthetic natural gas (CH4). Alkali and alkaline earth metals can serve both as promoters for methanation catalysts and as adsorbent phases upon the combined capture and methanation of CO2. Their promotion effect during methanation of carbon dioxide mainly relies on their ability to generate new basic sites on the surface of metal oxide supports that favour CO2 chemisorption and activation. However, suppression of methanation activity can also occur under certain conditions. Regarding the combined CO2 capture and methanation process, the development of novel dual-function materials (DFMs) that incorporate both adsorption and methanation functions has opened a new pathway towards the utilization of carbon dioxide emitted from point sources. The sorption and catalytically active phases on these types of materials are crucial parameters influencing their performance and stability and thus, great efforts have been undertaken for their optimization. In this review, we present some of the most recent works on the development of alkali and alkaline earth metal promoted CO2 methanation catalysts, as well as DFMs for the combined capture and methanation of CO2.

Graphical Abstract

1. Introduction

Antropogenic CO2 emissions have skyrocketed in the recent century due to the rapidly expanding energy demands of our society [1]. Since increasing concentrations of greehouse gas emissions have the potential to cause irreversible damage to our planet’s climate, it is clear that this trend eventually has to be reversed [2,3]. Carbon capture and storage (CCS) technologies aim to mitigate this problem by capturing CO2 from emission sources and then depositing it to underground geological formations [4]. However, CO2 utilization after the initial capture step (Carbon capture and utilization (CCU)) through conversion of CO2 into fuels and useful chemicals, such as methane, syngas, methanol and higher hydrocarbons, should be considered as a more prudent methodology to recycle our carbon sources [5]. Among the conversion products, methane or synthetic natural gas (SNG) is often favoured as it is an energy source with a high caloric content and also because the already existing and extensive natural gas grid can be utilized for its distribution [6]. However, the hydrogen used during the hydrogenation step should be produced using renewable energy sources in order to achieve carbon neutrality. Hope is offered by the intense research and development in the field of novel photovoltaics and electrolyzers, aiming to drastically lower the costs for the production of renewable hydrogen [7].
CO2 methanation, also known as the Sabatier reaction, is an exothermic reaction that converts CO2 to methane using hydrogen as the reducing agent [8]:
CO2 + 4H2 ⇌ CH4 + 2H2O, ΔH298K = −164 kJ/mol
This reaction has gained increased traction in recent years, since it can be employed to reduce our carbon footprint by recycling existing carbon sources, and it is also considered vital in the facilitation of long term space exploration missions and manned missions to Mars [8,9]. Other possible products besides CH4 during the hydrogenation of CO2 include CO via the reverse water-gas shift (RWGS) reaction [10], methanol (CH3OH) and C2+ hydrocarbons with two or more carbon atoms through a promotion of the C–C coupling step [11]. CO2 methanation targets the exclusive production of CH4 over the other possible hydrogenation products [6].
According to Vogt et al. [12], three different pathways can be conceived during the hydrogenation of CO2 to methane i.e., the carbide pathway through a CO intermediate, the formate pathway and the carboxyl pathway. The first two pathways have been observed for Ni-based catalysts through operando spectroscopic techniques (in-situ FTIR and ambient-pressure XPS), while the carboxyl pathway, though observed for Cu and Fe, is considered to be inactive for Ni-based catalysts [9,12]. Furthermore, catalyst supports can often play a major role in the methanation mechanism and the overall catalytic activity [13]. For example, Cárdenas-Arenas et al. [14] showed that oxygen species generated via CO2 dissociation on Ni/CeO2 catalysts could migrate throughout the ceria lattice and then be reduced at other sites, while the accumulation of formate intermediates at specific sites on Ni/Al2O3 catalysts delayed the chemisorption and activation of further CO2 molecules. These different reaction routes caused by the oxygen ion mobility and lability of CeO2-based supports could explain the greatly enhanced methanation activity of Ni catalysts supported on CeO2 compared to other metal oxide supports [14,15,16].
The initial capture of CO2 can be achieved by using liquids, such as aqueous amines, or solid materials, such as alkali and alkaline earth metal carbonates and oxides [17]. CaO solids promoted with different stabilizers are some of the most common CO2 sorbents and the cyclical process that integrates CO2 adsorption and desorption through CaO carbonization and calcination is often referred to as calcium looping [18]. Hydrogen from renewable sources can then be used to convert the captured CO2. This whole process (CCU) is more energy efficient, since an exothermic CO2 reduction step can provide the desired thermal energy for the endothermic desorption of captured CO2 [5,19].
The combined CO2 capture and conversion to methane can be realized either via the utilization of two different types of solids, namely one solid sorbent for the CO2 capture and one methanation catalyst, or via the use of a single solid known as a dual-function material (DFM) that incorporates both the sorption and methanation functions [20]. The Farrauto group have been pioneers in this new and dynamic field, being the first to report a DFM consisting of CaO and Ru phases supported on a γ-Al2O3 support. The CaO and Ru phases functioned as CO2 sorption and methanation active sites, respectively [19]. A recent review focusing exclusively on the combined CO2 capture and conversion to methane has been published in early 2019 [20].
A common feature in the design of such DFMs is the incorporation of an alkali or alkaline earth metal oxide/carbonate phase (e.g., CaO, Na2CO3, K2CO3 and MgO) on a high surface area support (e.g., Al2O3 and ZrO2) in order to be combined with adjacent catalytically active metal sites for CO2 methanation [20,21]. Ru has been a favourite active metal due to its easier reducibility and coke-resistant propensity, but Ni is also studied in an attempt to curb the material’s costs [22,23]. Various properties, such as the choice of the right support, the type of the CO2 sorption phase and the appropriate sorbent loading, as well as the choice of the active metal or the right combination of metals, have an important role to play in enhancing the CO2 sorption capacity and CH4 yield of the designed material formulations [23,24,25].
Hereby we aim to summarize the most recent progress made in the field of alkali and alkaline earth metal promotion of the CO2 methanation reaction and provide a comprehensive overview over the development of novel dual-function materials for the integrated capture and methanation of CO2. The latter materials, by functioning as both catalysts and adsorbents towards the combined CO2 capture and methanation, appear capable of improving the process and as a result have attracted intense interest in the last five years.

2. Alkali and Alkaline Earth Metals as Promoters of the CO2 Methanation Reaction

There has been extensive effort in trying to achieve high catalytic activity, especially in the low-temperature region (ca. 100–250 °C), for the CO2 methanation reaction. Rh and Ru are two of the most common precious metal catalysts studied in this reaction, since they achieve a substantial low-temperature activity and exhibit superior stability [26,27]. Ni-based catalysts, on the other hand, remain the favourite due to their low cost and large natural abundance, but they present some serious drawbacks, including inferior low-temperature activity when compared with Rh and Ru-based catalysts, as well as Ni nanoparticle sintering and coke formation [28,29]. Extensive research effort has been devoted in an attempt to enhance the low-temperature activity of Ni-based catalysts or decrease the loading of precious metal CO2 methanation catalysts by introducing some kind of metal or metal oxide promoter [27].
Alkali and especially alkaline earth metal promotion of the CO2 methanation reaction has been a quite popular route in recent years. These types of metals are cheap, abundant and can potentially promote the CO2 methanation reaction by increasing support basicity and by introducing oxygen vacancies, thus favouring the initial CO2 chemisorption [28,30,31], or even by strongly modifying the work function and thus the chemisoptive properties of the metal active phase [32]. Hereby, we categorize the different alkali and alkaline earth metal promoters starting from alkalimetals. Works that perform a screening of different promoters will be presented at first, followed by a categorization of the works focussing on a single metal promoter, depending on the type of metal used.

2.1. Alkalimetals (Li, Na, K and Cs)

2.1.1. Screenings of Various Alkalimetals

One of the few works reporting a promoting effect of various alkalimetals during the CO2 methanation reaction was conducted by Petala et al. [33]. The authors employed quite a small loading of alkalimetal promoters (0.06–0.4 wt.% of Li, Na, K and Cs) by impregnating the corresponding alkali precursor salts on the TiO2 support. A low (0.5 wt.%) and a high (5 wt.%) Ru loading was achieved via a subsequent impregnation step of Ru(NO)(NO3)3. The lower active metal loading resulted in a smaller Ru crystallite size. Interestingly, it was found that alkali addition had a greater impact on the catalyst with the low Ru loading (0.5 wt.%), in comparizon to that with the high Ru loading (5 wt.%), and an optimal Na loading of 0.2 wt.% was obtained for the former (i.e., 0.5 wt.% Ru/0.2 wt.% Na-TiO2). These results agree with the general features of alkali promotion on several catalytic systems, which typically obey a volcano-type behavior with respect to the promoter/metal ratio; low promoter/metal ratios are inefficient to induce succesful promotion, high promoter/metal ratios lead to over-promotion (poisoning) [32]. The catalytic activity and CH4 selectivity were promoted according to the following order with respect to the alkali identity: TiO2 (unpromoted) < Li ≈ K < Cs < Na, as shown in Figure 1 [33]. In a continuation of this work, Panagiotopoulou [34] employed in-situ DRIFTS and transient-mass spectrometry in order to shed light on the reaction mechanism. Alkalimetal addition on the low Ru loading catalyst appeared to favour the formation of Ru-bonded carbonyl species (Rux-CO), which are key intermediates towards the production of CH4.
The impact of alkalimetal addition on a Ru-based catalyst towards the methanation of CO2 was also studied by Cimino et al. [35] in a recent study. In this case, a higher alkali promoter loading was chosen and the impregnation of the alkali precursor salts took place mostly after the impregnation of the Ru active metal. All alkali modified catalysts, except the one modified with LiNO3, exhibited lower CO2 conversions and CH4 selectivities compared to the unmodified 1% Ru/Al2O3 one. The lower catalytic activity was attributed to the coverage of some Ru active sites, while Li could promote the Ru-based catalyst activity by forming Li-aluminate spinel-type phases with the Al2O3 support. The LiNO3 and NaNO3 promoted materials were also tested for cyclic CO2 capture and methanation and the results will be further discussed in the second part of this review.
Liang et al. [36] and Zhang et al. [37] studied the effect of K and Na modification (amongst other metals) on Ni/Al2O3 and Co/Al2O3 catalysts, as well as the alteration of the CO2 methanation mechanism via in-situ DRIFTS. In the first study [36], a total of 16 different metal additives were first impregnated on a γ-Al2O3 support, followed by a subsequent impregnation of Ni. Na and K additives gravely impacted the physicochemical properties of the catalysts, leading to lower activity. The lower catalytic activity of the K-modified catalyst was attributed to the increased population of H2CO* intermediates. In the second study [37], the modification of the alumina support in Co/Al2O3 by K, Na and Mg led to the formation of Al(OH)3 and spinel-type structures. The addition of Na and K followed by the increase in support alkalinity appeared to stir the reaction towards HCOO* and CO32− intermediates, lowering the catalytic activity and CH4 selectivity, as well as favouring coke formation. The promotion of the CO2 chemisorption by alkalimetal addition did not appear to correlate with the catalytic activity.
In regards to the electrocatalytic CO2 methanation, the Vayenas group employed the electrochemical promotion of catalysis (EPOC) concept, an effective in-situ method of catalyst promotion, also known as non-faradaic electrochemical modification of catalytic activity (NEMCA) [32,38] to enhance CO2 methanation performance of a Ru electrocatalyst [39,40]. Taking advandage of the high potential of the method, the authors comparatively investigated a variety of electropositive or electronegative promoters, namely Na+, K+, H+ and O2−. They found that in all cases, increasing catalyst work function (induced by the promoters) enhances the methanation rate and selectivity. In the case of the present interest (alkali promotion), Na+ and K+ supply to Ru surface caused a remarkable promotion of the methanation rate, which was found to pass through a maximum upon increasing alkali loading (volcano-type promotion). This result is in qualitative agreement with the typical features of alkali-promotion [32] and with the aforementioned results of Petala and Panagiotopoulou on CO2 methanation over traditional supported Ru catalysts upon varying Na-loading by an impregnation method [33].

2.1.2. Sodium (Na)

Among the works focusing on the support modification with Na, Le et al. [41] investigated the CO2 methanation reaction over Ni catalysts supported on Na (0.1–1 wt.%) doped SiO2 and CeO2 supports. Na was incorporated (via impregmation) in the previously made 33% Ni/SiO2 and 13% Ni/CeO2 catalysts in a second impregnation step using a NaNO3 precursor. The available active metal surface area was shown to decrease upon Na addition, following H2 chemisorption experiments. Subsequently, CO2 conversion decreased upon Na modification of the initially highly active Ni/CeO2 catalyst, while the increase of the amount of moderately strong basic sites and CO2 chemisorption sites following the incorporation of Na in Ni/SiO2 appeared to offset the negative effect caused by the decrease of the catalytically active surface area.
Moreover, Beierlein et al. [42] employed various preparation methods in order to prepare Ni/Al2O3 catalysts with high Ni loadings (14–88 wt.% Ni). The primary focus of their work was to establish a linear correlation between the surface area of the active metal (Ni) and the CO2 conversion, arguing that CO2 methanation over highly loaded Ni/Al2O3 catalysts is a structure-insensitive reaction. Two of the preparation methods applied, namely deposition precipitation and coprecipitation, required the use of large quantities of NaOH and Na2CO3 reagents, thus introducing a relatively high content of Na on the final catalysts. A high Na content was shown to decrease CH4 selectivity, while boosting CO production. This behaviour was attributed to the lower hydrogen coverage of the active metal and the hindrance of the H-assisted CO dissociation, a key intermediate step (also reported to be the rate determining step [9]) during the conversion of CO2 to CH4.

2.1.3. Pottasium (K)

Pottasium-doping of the catalyst support has been performed by various groups with most of them reporting a negative effect on the CO2 methanation. Büchel et al. [43] studied the CO2 hydrogenation over K- and Ba-modified 1% Rh/Al2O3 catalysts prepared by flame spray pyrolysis. The K-modified catalyst produced only CO over the entire temperature range. This extreme CO selectivity over CH4 was attributed to a weak interaction of CO with the catalyst, as shown during CO chemisorption experiments. This could cause the intermediate CO species to desorb, rather than be further hydrogenated into CH4. Iloy et al. [44] further studied the modification of 15% Co/SiO2 with 1–5 wt.% K. K-dopant was shown to enhance CO2 chemisorption, but favour CO selectivity and C–C coupling along with chain growth reactions. For the production of C2+ hydrocarbons, 1 wt.% K appeared to be optimal, while a higher potassium loading favoured CO as a main product.
Modification of the Al2O3 support with a KOH strong base was performed by Zhang et al. [45] and the reaction intermediates were followed via in-situ DRIFTS. KOH was first dissolved in d-H2O and then introduced into the Al2O3 support, while Ni was impregnated during a second step. KOH addition led to a dramatic decrease of the Al2O3 surface area and after calcination, K2CO3 and K2O phases were also detected. The introduction of the strong KOH base led to the increased presence of adsorbed formate intermediates during CO2 hydrogenation, which was validated via in-situ DRIFTS (Figure 2) [45]. This strong interaction with the formate intermediate species apparently did not allow for their further hydrogenation to CH4 and instead led to a high CO selectivity. The Ni-KOH/Al2O3 catalysts, on the other hand, could be considered as more suitable candidates for the RWGS reaction and the steam reforming of acetic acid, in an antithesis to other works:
He et al. [46] reported a promoting effect of KOH addition towards higher CO2 conversions and CH4 selectivities, when impregnated on a Ni-Al hydrotalcite precursor. This was attributed to the introduction of new strong basic sites in the support. The different behaviour in this case could be likely due to the small loading of the KOH base (0.3 wt.%), similarly to the work of Petala et al. [33]. This also demonstrates the key role of the alkali promoter/active metal ratio in alkali promotion efforts and the nessecity of a thorough screnning for the determination of the optimal alkali promoter loading for each specific catalytic system, alkali indentity and reaction conditions used [47,48].
Another interesting study that includes an in-situ KOH modification of a Ni/SiO2 methanation catalyst has been conducted by Vogt et al. [49]. The authors first created a setup that produced hydrogen through solar driven H2O electrolysis from a KOH-based alkaline electrolyzer. Produced H2 was then transferred downstream to a methanation unit that contained a Ni/SiO2 methanation catalyst. During operation, KOH aerosols were transferred to the methanation reactor and part of them was deposited on the Ni-based catalysts. Interestingly, this in-situ alkali modification promoted the Ni/SiO2 methanation catalyst, whereas a similar ex-situ 0.6 wt.% K addition via a common KOH impregnation resulted to catalyst deactivation.
LaNiO3-based perovskite oxides can serve as catalyst precursors for the preparation of Ni/La2O3 catalysts following a reductive pretreatment. The fine and well-dispersed Ni nanoparticles formed following the perovskite lattice decomposition retain a strong interaction with the La2O3 support and are expected to exhibit a different behaviour compared to the Ni/La2O3 catalysts prepared via a mere impregnation step [50]. Tsounis et al. [51] presented a way to tune the product selectivity of LaNiO3-based perovskites during CO2 hydrogenation, via partially substituting La on the perovskite A-site with K. KNO3 was mixed together with La and Ni nitrates during the perovskite citrate sol-gel synthesis, in order to prepare La1-xKxNiO3 precursors. Following reduction, the catalysts exhibited a drastically different behaviour during CO2 hydrogenation depending on the amount of K-doping. The K-free catalyst had a high CO2 conversion and an almost complete selectivity towards CH4, while with increased amounts of K-substitution CO2 conversion gradually decreased and CO selectivity spiked (Figure 3). A possible explanation given was that easy-to-desorb CO intermediates were formed after K-doping and they in turn stirred the reaction towards the RWGS route.

2.2. Alkaline Earth Metals (Mg, Ca, Sr and Ba)

2.2.1. Screenings of Various Alkaline Earth Metals

In the works discussed previously, some alkaline earth dopants were also used and compared with the effect caused by the doping with alkalimetals. Zhang et al. [37] concluded that the effect of Mg and Ca addition on Co/Al2O3 was less pronounced compared with Na and K and that, although the activity of the modified catalysts was inferior to Co/Al2O3, alkaline earth metal doping still provided a benefit by preventing Co nanoparticle sintering during prolonged exposure under reaction conditions. Liang et al. [36] also mentions that 5% Mg and Ca doping on Ni/Al2O3 led to a slight increase in support alkalinity, but did not appear to benefit the methanation activity.
Guo et al. [52] were among the first to conduct a comparative study on the impact of alkaline earth metal addition on Ni/SiO2 catalysts. During the catalyst preparation, alkaline earth metal nitrates were first impregnated on the SiO2 support, followed by wet impregnation of the Ni active phase. The authors reported that MgO had a negative effect on the catalytic activity, while CaO did not have a pronounced effect. The heavier alkaline earth metal oxides SrO and BaO led to better results, both enhancing the catalytic activity. SrO was shown to be the most suitable additive by providing the highest promotion of the CO2 methanation reaction and by increasing the stability of the modified catalyst.
Liang et al. [53] studied the modification of Ni/Al2O3 catalysts with alkaline earths for the CO2 methanation reaction and reported similar results. The catalysts were also prepared via subsequent impregnation of alkaline earth metal nitrates and then nickel nitrate on the support. The lighter alkaline earth metals Mg and Ca were found to promote the formation of carbonate intermediates and the CO selectivity, while the heavier alkaline earth metals Sr and Ba were the ones being beneficial for the CH4 yield by promoting the formation of *CO and H2CO* intermediates. Another important feature, namely the generation of oxygen vacancies upon Sr-addition, appeared to play a major role via the creation of Lewis basic sites for the adsorption and activation of CO2.
Ni/CeO2 catalysts are among the most active for CO2 methanation and are a main research focus. CeO2 has a high oxygen storage capacity (OSC) that, together with its redox properties due to the Ce4+/Ce3+ redox couple, promotes the stability of the active metal particles and suppresses the formation of coke. Furthermore, CeO2 facilitates a strong interaction with the supported metal phase [54,55]. Liu et al. [30] conducted a comprehensive study, where they doped the CeO2 support of Ni/CeO2 catalysts with various alkaline earth metals, with the aim to further enhance their low-temperature catalytic activity. Alkaline earth metal nitrates were introduced during the citrate sol-gel synthesis of the CeO2 support with a molar ratio of 1/9, compared with Ce(NO3)3·6H2O. Mg2+ and Ca2+ appeared to fully dissolve into the CeO2 lattice, while SrCO3 and BaCO3 secondary phases were detected in the other two samples. The Ca-doped catalyst induced the highest promotion effect, measured via the CO2 conversion and turnover frequency (TOF) increase at low temperatures, with the Sr-doped one being second-to-best. The higher turnover frequency could be linearly correlated with two related factors, namely the increase in the population of oxygen vacancies (Ce3+/(Ce3+ + Ce4+) molar ratio) and moderate alkaline sites (200–450 °C peaks in the CO2-TPD profiles), as shown in Figure 4.
The studies discussed up to this point, which focused on the screening of different alkaline earth metals as dopants, appear to provide conflicting conclusions on their effect, since some of them report enhancement of the CO2 methanation activity, while others a suppression of it (for Mg and Ca). One must take into account that these works differ on the type of support, the alkaline earth metal loading and the preparation procedure. Notably, almost all works focusing on a single alkaline earth metal additive, especially concerning Mg and Ca doping, report a promoting effect for the CO2 methanation. Generally, this promotion is attributed to a better active metal dispersion, as well as to the generation of surface oxygen vacancies and the enhancement of support alkalinity, factors that substantially favour the initial chemisorption and activation of CO2.

2.2.2. Magnesium (Mg)

Mg in the form of MgO has been studied as a type of support during the CO2 methanation reaction. However, Ni/MgO catalysts have been shown to lack the activity of other Ni catalysts supported, for example, on TiO2, Al2O3 and CeO2 [15]. The CO2 methanation mechanism over Ni/MgO has been reported to proceed via a H2COO* intermediate and to follow the formate route, due to the strong metal-support interaction (SMSI) between the Ni nanoparticles and the MgO support [56]. Ni/MgO catalysts encapsulated by a SiO2 shell were prepared by Li et al. [57] via chemical co-precipitation and a modified Stöber method, followed by in-situ reduction of the initial SiO2 encapsulated NiO-MgO solid solution. Interestingly, it was found that the confinement induced by the SiO2 shell helped to control the size of Ni nanoparticles, whereas the MgO phase also contributed to the improved Ni dispersion.
Besides its function as a catalyst support, MgO has also been thoroughly studied as a promoter for the CO2 methanation reaction in an attempt to improve the dispersion of the active metal phase and enhance the support basicity. Park et al. [58,59] reported a bifunctional mechanism of CO2 methanation over a Pd-MgO/SiO2 catalyst. The addition of MgO in Pd/SiO2 led to improved catalytic activity and CH4 selectivity. A mechanism was proposed where MgO and Pd phases functioned as CO2 and H2 sorption active sites, respectively. The adsorbed hydrogen atoms on the Pd surface were then proposed to spillover to Mg carbonates, hydrogenating the CO2-adsorbed phases and forming the CH4 final product. This bifunctional character of CO2 methanation reaction that requires the presence of a catalyst componentent capable of activating the CO2 (Supporting Material: even the commonly considered “inert” Al2O3 acts in this way; support doping is often used for further enhancing this role), and a metallic component typically active in H2 dissociation has been well established in the literature [60,61,62].
Guo et al. [63] prepared Ni/SiO2 catalysts with different MgO loadings and with two different preparation methods, namely co-impregnation and sequential impregnation. Co-impregnation was shown to be superior to sequential impregnation and 1% MgO loading to be sufficient for the reaction promotion, whereas a higher MgO content risked masking some of the Ni active sites. The promotion was reportedly achieved via the enhancement of Ni dispersion and via Ni particle sintering prevention. The beneficial effect of MgO in the Ni nanoparticle dispersion was also reported by Tan et al. [64]. Ni and Mg nitrates were co-impregnated on a ZrO2 support with the simultaneous addition of citric acid and glycol during the impregnation step, applying the citric complexing method. An optimal Mg/Ni molar ratio of 0.25 led to the desired Ni nanoparticle confinement and resistance to sintering and coke formation. Moreover, Xu et al. [65] included Mg in the synthesis of ordered mesoporous Ni/Al2O3 catalysts prepared via evaporation-induced self-assembly (EISA). Mg increased the surface basicity of the catalysts, thereby facilitating the chemisorption and activation of CO2. The optimized Mg content of 5% and the confinement created by the mesoporous alumina structure provided significant advantages for the employment of these materials in the CO2 methanation reaction.
Further studies include the promotion of Ni-based catalysts supported on coconut shell carbon (CSC) [66] and MCM-41 [67] with Mg, as well as the Mg-modification of catalysts comprised of Ni supported on ultrastable Y (USY) and MFI (Mobil-type five) nanosponge zeolites [68,69]. All these studies reported an improvement of the CO2 methanation catalytic performance. Additionaly, a second promoter could also be introduced together with Mg to investigate co-promotion effects [70,71]. La-promoted Ni/MgO-Al2O3 catalysts showed improved low-temperature activity via the increase in the population of moderately strong basic sites [60], while Mg and Mn co-promoted Ni/α-Al2O3 catalysts were shown to be superior compared to the ones prepared with the use of a single promoter [71].
Finally, Mg has been indicated as a suitable promoter for Fe catalysts during CO2 methanation [72]. Fe catalysts, commonly known to be active during the Fischer–Tropsch synthesis, could also be viable candidates for the CO2 hydrogenation into CH4 under high pressure, after the incorporation of a suitable promoter [73]. Baysal et al. [72] incorporated a series of 15 different promoters on a commercial α-Fe2O3 powder via incipient wetness. Improved CO2 adsorption over the Mg-modified Fe catalyst led to highly active Fe carbides, bulk carbon moieties and an overall improved CO2 methanation performance.

2.2.3. Calcium (Ca)

Ca is also one of the favourite promoters for the CO2 methanation reaction, with its main functions being the increase in the population of surface basic sites and the alteration of support defect chemistry by introducing oxygen vacancies. Similarly to the Mg doping of Ni-based catalysts supported on ordered mesoporous alumina [65], Ca was also employed as a basic modifier by Xu et al. [74]. Ca(NO3)2·4H2O was introduced during the evaporation-induced self-assembly (EISA) catalyst synthesis. The surface alkalinity was greatly increased upon Ca-doping and that led to an enhancement of CO2 chemisorption and a lowering of the activation energy for CO2 methanation. The spike in the population of new strong alkaline sites was confirmed by CO2-TPD. The catalytic activity was the highest, when the Ca/Al molar ratio was fixed at 8%, with a roughly 4 times CO2 conversion increase at low temperatures (250 °C).
Liu et al. [75] tried to further optimize these types of catalysts by an additional introduction of 1% Ru. The 10N1R2C-OMA or 1% Ru, 10% Ni, 2% CaO/Al2O3 (ordered mesoporous alumina) catalyst aimed at combining the beneficial effects of: (a) the nanoparticle confinement inside an ordered mesoporous alumina framework, (b) the increase of surface basicity and CO2 chemisorption using a CaO promoter and (c) the introduction of 1% noble metal Ru, forming a bimetallic NiRu phase and creating a synergy between the two catalytically active metals. A step increase in the CO2 conversion and CH4 selectivity was reported following the subsequent introduction of each modifier. Furthermore, the ordered mesoporous structure and the fine dispersion of the active metal phase were retained in the spent catalysts, highlighting their superior stability. The methanation performance of the prepared catalysts, as well as the stability test for the 10N1R2C-OMA one, are shown in Figure 5.
Furthermore, Do et al. [76] reported a promoting effect of CaO addition on a bimetallic PtNi/Al2O3 catalyst during methanation of a CO2-rich gas, while the group of Chu also published a a number of works focussing on the introduction of Ca-dopant as a promoter for the methanation of CO2 [77,78,79]. A 20% CaO addition on 15% Ni/Al2O3 was reported to modify the interaction between Ni and Al2O3 and inhibit the incorporation of Ni into the alumina structure [77], while Ca addition on carbon supported Ni catalysts (carbon nanotubes [78] and activated carbon [79]) also appeared to be beneficial by increasing Ni reducibility and by limiting the growth of the Ni nanoparticles.
Ca2+ cations are soluble into the crystalline lattice of ZrO2 and the corresponding solid solution is commonly known as calcia stabilized zirconia (CSZ). Oxygen vacancies (Vo2) are created in order to compensate for the negatively charged Ca-induced substitutional impurities (CaZr2’), thereby enhancing the material’s ionic conductivity and altering its defect chemistry [80]. Takano et al. [81] impregnated Ni and Ca nitrates on ZrO2 and studied the effect of Ca/(Zr + Ca) molar ratio on the CO2 methanation activity. The authors argued that the presence of medium-strength basic sites caused by the Ca-induced oxygen vacancies could facilitate CO2 chemisorption, which then further reacted with atomic hydrogen adsorbed on Ni adjacent sites. An optimized Ca/(Zr + Ca) molar ratio between 16.7% and 20% was shown to maximize the population of oxygen vacancies, while preventing the formation of undesirable Ca-rich and CaZrO3 phases. More recently, Everett et al. [82] showed that a small amount of Ca2+ substitution (0.7 wt.%) in the lattice of ZrO2 generated pairs of oxygen vacancies and coordinately unsaturated sites (cus). They concluded that these pairs could facilitate the C-O bond scission on adsorbed intermediate CO, reportedly the rate determining step during CO2 methanation.
Finally, perovskite oxide titanates containing Ca have been also used in the CO2 methanation reaction. Jia et al. [83] employed a common perovskite oxide that is also found as a natural mineral, CaTiO3, as a catalyst support and compared it with γ-Al2O3. The 10% Ni/CaTiO3 catalyst exhibited CO and CO2 methanation activity superior to 10% Ni/Al2O3, due to the perovskite oxide’s exceptional stability, increased support basicity and the limited interaction between Ni particles and the CaTiO3 support. Do et al. [84] used Ca as a promoter in order to substitute part of the A-site cations in the NiTiO3 lattice and the whole perovskite was loaded on γ-Al2O3. They performed extensive characterizations on the composite materials and concluded that upon 5% molar substitution of Ni2+ with Ca2+, the catalytic activity could be maximized. Ca-addition promoted the formation of oxygen vacancies in close proximity to Ni active phases and thereby the CO2 chemisorption. The proposed reaction mechanism, as summarized in Figure 6, involved a combination of the CO and formate routes. The initial CO2 adsorption and hydrogenation steps resulted into formate species, which in turn formed adsorbed CO. The carbonyl species were then further reduced to form other hydrogenated intermediates, with CH4 being the final product.

2.2.4. Strontium (Sr) and Barium (Ba)

Despite the fact that the works dealing with the screenings of different alkaline earth metal promoters report a quite beneficial contribution upon Sr and Ba addition [30,52,53], little further research has been conducted on the individual promotion of either of these heavier elements, especially when compared to their lighter alkaline earth metal counterparts. Toemen et al. [85] modified a Ru/Mn/Al2O3 catalyst with Sr and Ce. The Sr-modified catalyst initially adsorbed increased amounts of CO2 as monodentate carbonates, which in turn formed bridged bidentate carbonates before being further hydrogenated. The enhanced catalytic activity of the Sr-doped catalyst was, however, overshadowed by the Ce-doped one, where the methanation reaction followed a different mechanism.
Steiger et al. [86] prepared La1-xSrxFe0.8Ni0.2O3 perovskite oxides with increasing amounts of A-site Sr substitution. Pre-reduction of the initial perovskite caused the lattice Ni to be exsolved on the perovskite surface in the form of fine and well-dispersed Ni nanoparticles that exhibited strong interaction with the perovskite support. The amount of exsolved Ni, as well as the size and dispersion of the Ni nanoparticles, could be tailored by altering the A-site perovskite composition. La substitution on the A-site by Sr caused a destabilization of the perovskite lattice and an increase in the amount and population of exsolved Ni nanoparticles. This in turn promoted the CO2 methanation reaction by increasing the CH4 yield, though a possible modification of the surface chemistry due to the increased presence of Sr could also affect the reaction.
Regarding Ba, it was mentioned in a previously discussed work [43] that K-modified 1% Rh/Al2O3 was highly selective towards CO. The authors also prepared a Ba-modified Rh/Al2O3 catalyst, whose catalytic performance was similar to the unpromoted Rh/Al2O3 one. However, the CH4 yield was lower, especially at lower temperatures, and the enhanced CO2 affinity due to the Ba-addition did not benefit the overall methanation activity. The results appear to contrast the findings of previously discussed works [52,53] that report a promoting role of Ba for the CO2 methanation reaction, probably as a result of different types of catalysts, preparation methods, reaction conditions and Ba loadings used.

2.3. Hydrotalcite-Derived and Zeolite Supported Catalysts

Besides commonly used supports such as Al2O3, ZrO2, SiO2 and TiO2, MgO-Al2O3 mixed oxides derived from Mg-containing hydrotalcites, as well as zeolites, have also become popular CO2 methanation catalyst supports in recent years. Hydrotalcites, or layered double hydroxides, with Ni2+ and Mg2+ as divalent cations can be calcined and reduced at high temperatures in order to produce active catalysts that consist of Ni nanoparticles supported on a MgO-Al2O3 mixed oxide. Mg2+ presence in the initial hydrotalcite facilitates the creation of alkaline sites on the support that favour CO2 chemisorption and activation. Zeolites can also be impregnated with an active metal phase (e.g., Ni and Rh) to be used as methanation catalysts. These types of catalysts have gained increased traction recently and two review articles have been published summarizing the progress made [87,88]. Since an Mg component is mostly present in hydrotalcite-derived catalysts and alkali, as well as alkaline earth metals, serve as compensating cations in zeolites, we will provide here a brief overview on the influence that these two parameters exert on the CO2 methanation activity.

2.3.1. Hydrotalcite-Derived Catalysts

Hydrotalcite materials, also known as layered double hydroxides, have the general formula of [M1-x2+Mx3+(OH)2](An-)x/n·mH2O with M representing divalent and trivalent metals and A being an anion. Mg2+ and Ni2+ can serve as divalent cations, while Al3+ is a common trivalent cation. Upon calcination and reduction, supported Ni catalysts are obtained with high activity and selectivity for specific catalytic reactions [87]. Bette et al. [89] claimed to be the first to present an active CO2 methanation catalyst derived from a hydrotalcite precursor (Ni0.5Mg0.17Al0.33(OH)2(CO3)0.17·mH2O) synthesized via the coprecipitation method. The resulting catalyst consisted of a high Ni loading (59 wt.%) supported on a spinel-type (Mg,Al)Ox mixed oxide and exhibited stable catalytic performance, as well as Ni particle sintering and coking resistance.
The cooperation between the Ni catalytically active phase and the MgO alkaline phase in hydrotalcite-derived catalysts has been studied by Liu et al. [90]. The authors varied the Ni2+/Mg2+ molar ratio during the coprecipitation synthesis, while the ratio between trivalent (Al3+) and divalent (Mg2+ and Ni2+) cations remained constant. After calcination and reduction, the formed Ni nanoparticles were well-dispersed over the MgO-Al2O3 support with an average crystallite size of about 2.5 nm. Two types of basic sites were observed upon CO2-TPD, namely weak basic sites for CO2 adsorbed on Al2O3 phase and moderately strong basic sites for CO2 adsorbed on MgO phase. Regarding the catalytic performance for CO2 methanation, the hydrotalcite-derived catalysts presented far superior low-temperature CO2 conversion compared with other Ni-based catalysts supported on Al2O3, MgO and carbon nanotubes (CNTs) (Figure 7a). Moreover, a linear correlation could be derived between TOFNi values and the amount of moderately strong basic sites on the support (Figure 7c). The reaction mechanism proposed involved a CO2 activation step by the support basic sites to carbonate/hydrocarbonate species and an H2 dissociation step on the Ni metallic sites. The interplay between these two phases facilitated the excellent low-temperature activity towards CO2 methanation.
Further works report on the influence of other parameters, as well as on the incorporation of further promoters and metals. Ren et al. [91] changed some variables during the hydrotalcite coprecipitation synthesis that led to platelet-like and rosette-like support structures and emphasised on the effect of support morphology and metal-support interaction over the CO2 methanation performance. Wierzbicki et al. [92,93] introduced La during the hydrotalcite preparation to further promote the support alkalinity and CO2 chemisorption, as well as the interaction between the Ni phase and the hydrotalcite-derived support. Additionally, bimetallic catalysts were prepared by Martins et al. [94], with Ru being impregnated over the initial Ni-Mg-Al hydrotalcites, while Mebrahtu et al. [95] varied the amount of added Fe during the hydrotalcite synthesis. The intimate mixing of the Ni2+ and Fe3+ metal cations led to the formation of NiFe alloys after reduction and the alloy composition could be correlated with the catalytic performance [95]. We can thereby conclude that hydrotalcite precursors offer a unique possibility to tailor the catalyst composition and morphology, as well as examine the influence of various different factors towards enhancing the CO2 methanation activity and stability.

2.3.2. Zeolite Supported Catalysts

Zeolites are microporous aluminosilicates with a general formula of [AlxSi1-xO2y-]x and consist of a negatively charged network that requires cationic counterions (e.g., Na+, Ca2+) for charge compensation. Through cation exchange and post-synthetic modification, zeolitic physicochemical properties, such as basicity and hydrophilicity, can be finely tuned. Furthermore, a transition metallic phase, such as Ni or Rh, can be introduced via ion-exchange or impregnation to prepare zeolite supported catalysts suitable for CO2 methanation applications [88]. Graça et al. [96] were among the first to introduce Ni on ultrastable Y (USY) zeolites with Na+ counter-ions for application as CO2 methanation catalysts. Ni introduction via impregnation proved to be more suitable compared to introduction via ion-exchange because of the higher reducibility of impregnated NiO species, while a latter Ce addition improved the performance further via its synergy with the Ni metallic phase. Westermann et al. [97] studied the mechanism of CO2 methanation over Ni/USY catalysts and concluded that CO2 was initially weakly adsorbed over the Na+ compensating cation sites on the zeolite. Following the initial chemisorption step, CO2 hydrogenation proceeded via a CO intermediate adsorbed on the Ni particles.
Bacariza et al. [98] then performed an extensive work on the influence that several monovalent and divalent compensating cations exert over the CO2 methanation performance of Ni/USY catalysts. The compensating cations were ion-exchanged in the H-USY zeolite in its acidic form and the Ni phase was impregnated subsequently. Regarding the monovalent cations, a trend was observed where the CO2 methanation performance (CO2 conversion and CH4 selectivity) increased for the heavier elements, following the order Cs+ > Na+ > Li+ > K+ > H+ (with K+ being an exception). These differences were attributed to the increase of the CO2 adsorption capacity through a higher support alkalinity. Regarding the divalent cations, the catalytic activity followed a different trend by decreasing for the heavier elements in the order Mg2+ > Ca2+ > Ba2+. The better performance for ion-exchange with Mg2+ was attributed to the higher Ni dispersion and Ni surface area. Thus, ion-exchange of compensating cations could be a correct pathway to tune specific properties of zeolite supported CO2 methanation catalyst, such as metallic phase dispersion, support alkalinity and water affinity.
Besides the effect of different compensating cations, the optimization of the Si/Al ratio and the introduction of additional promoters (e.g., Mg, La and Ce) via impregnation has been studied. Bacariza et al. [99] modified the USY zeolitic composition in terms of Si/Al ratio and observed that higher Si/Al ratios decreased the water affinity and provided an enhanced methanation activity. Mg [68], La [100] and Ce [101] have also been employed as further promoters and introduced via impregnation over the zeolite support. The introduction of these promoters aimed to improve the dispersion of the metallic phase and promote CO2 chemisorption via enhancing the surface basicity. Bacariza et al. [88] also presented some proposals for the preparation of active zeolite supported CO2 methanation catalysts that incorporate properties such as high metallic phase dispersion, increased zeolite support basicity, low water affinity and sufficient steam resistance.
Lastly, Zhang et al. [102] prepared a novel methanation catalyst by using a Na-modified Y zeolite as a support and Al-source for the subsequent formation of a Ni-Al hydrotalcite precursor phase. After reduction, Ni-based, hydrotalcite-derived catalysts supported on the zeolitic structure were created. The catalysts exhibited a unique mosaic structure and the supported Ni nanoparticles were successfully confined by the AlOx phase, thus improving their resistance to sintering and coke formation. Table 1 summarizes the CO2 methanation performance of some typical catalysts promoted with group I and II metals.

2.4. Other Promoters that Favour CO2 Chemisorption

Besides alkali and alkaline earth metals, lanthanides or rare earths, such as La, Ce and Pr [103], transition metals, such as V, W, Mn, Co and Fe [39,104,105], as well as noble metals, such as Pd, Pt, Rh and Ru [106], are commonly introduced into CO2 methanation catalysts to enhance their activity. A common Ni-based catalyst promotion with metals, such as Ru, Rh, Pt, Co and Fe, mostly aims at the creation of bimetallic and alloy particles in order to improve the properties of the Ni metallic phase, such as reducibility, metal dispersion and resistance to oxidation, agglomeration and coking [8,28]. On the other hand, lanthanides and some transition metals often target the modification of the metal oxide support to improve properties such as surface basicity and oxygen vacancy concentration, leading to enhanced CO2 chemisorption similarly to the promotion with alkali and alkaline earth metals [27,28].
Just like the works dealing with screenings performed using alkali and alkaline earth metals, Xu et al. [103] and Guilera et al. [107] screened several rare earth and transition metals (i.e., La, Ce, Sm, Pr, Y and Zr) as promoters over Ni/Al2O3 catalysts. Among them, Pr, Ce and La exhibited the most pronounced promoting effects. La-promotion specifically, has been reported to be beneficial for the CO2 methanation performance by a number of studies, as it appears to enhance the support alkalinity. La-addition on the support can been performed either via separate impregnation [108,109] or via co-impregnation with Ni that targets the formation of an initial LaNiO3 perovskite precursor phase [110,111]. Lastly, Mn promotion also targets similar improvements, namely enhancing the CO2 adsorption and increasing the metal phase dispersion [104].

3. Alkali and Alkaline Earth Metals as Active Sorbents Phases in the Combined Capture and Methanation of CO2

Carbon capture and utilization (CCU) revolves around the concept of coupling post-combustion CO2 capture to a CO2 utilization step [5]. If this step is an exothermic reduction reaction, then the energy released can be used to mitigate the energy demands arising from the endothermic CO2 desorption step. Additionaly, CH4 selectivity for many catalysts during CO2 hydrogenation is high at temperatures that can coincide with those required for CO2 desorption, thus rendering CH4 as a preferable final product. The combination of these two processes, namely CO2 adsorption and hydrogenation, in a single reactor can be realized via the development of novel dual-function materials (DFMs) that embody both adsorption and catalytic capabilities [19,20,25].
In most DFM materials targeting CH4 production, Ru is chosen as the preferred metal to provide the catalytically active sites, despite it being significantly less abundant and pricier compared to Ni. Since most flue gases are O2-rich, a partial oxidation of the metallic surface takes place during the CO2 capture step under practical operation conditions. The catalyst then needs to be reactivated under a H2-rich gas flow. Ru is most promising in this regard, since the oxidized RuOx layers can be re-reduced under H2 flow at temperatures where CO2 capture takes place (< 350 °C), thereby increasing the process energy efficiency by operating both steps isothermally [22]. Oxidized NiO requires more intense reducing environments to be reactivated, depending also on the support used and the type and loading of the adsorbent phase. A higher temperature activation step or the absence of O2 from the CO2-containing gas are thus necessary prerequisites for the employment of Ni-based DFMs [21,112]. By promoting a Ni catalyst with a second metal though, its reducibility could be enhanced in order for it to meet the requirements to be used under more practical flue-gas applications [23].

3.1. Ru-Based Dual-Function Materials

Duyar et al. [19] introduced the concept of a DFM for the CO2 capture and conversion to CH4 (Figure 8) [113]. The materials were prepared via incipient wetness impregnation, either impregnating Ca(NO3)2 first and then Ru(NO)(NO3)2 on γ-Al2O3, or vice versa. The loadings of the adsorbent (CaO) and the catalyst (Ru) were varied between 1 and 10 wt.%. It was shown that first impregnating the sorbent and then the catalyst was a more suitable route for the material preparation, whereas a high sorbent loading appeared to be beneficial, since it increased the amount of captured CO2 and the extent of CO2 spillover towards the methanation active sites. An optimum composition of 5% Ru, 10% CaO/γ-Al2O3 produced the maximum amount of CH4 per kg of DFM material at 0.5 mol CH4/kg DFM (or 0.5 mmol CH4/g DFM). The authors also showed that the high dispersion of the CaO sorbent phase, that can facilitate a reversible chemisorption and desorption of CO2 at moderate temperatures, in combination with the easily reducible and highly active Ru methanation catalyst, could enable the whole process to proceed isothermally at 320 °C. A follow up parametric study [114] reported an effective DFM preparation method, as well as suitable adsorption and methanation conditions in order for it to be applied to industrial-scale applications.
Following the earlier works, Duyar et al. [113] also used Rh as the active metal with CaO adsorbent phase and they showed that a comparatively miniscule amount of Rh loading in the DFM (0.1 wt.%) could generate comparable amounts of CH4 with a 5% Ru loading. However, the high price of Rh is still a major problem and thus, the less expensive Ru is considered as a more suitable active metal. In the second part of their work, Duyar et al. [113] tested K2CO3, Na2CO3 and MgO as potential CO2 sorbent alternatives. K2CO3 and Na2CO3 exhibited better results, with the 5% Ru, 10% Na2CO3/γ-Al2O3 DFM yielding 1.05 mol CH4 per kg material at 320 °C operation. Thus, it was shown that the optimization of the CO2 sorption phase can be of equal importance to that of the catalytically active metal phase in an attempt to lower the catalyst material costs and increase the process efficiency. A parametric study was also conducted for the new optimized 5% Ru, 10% Na2CO3/γ-Al2O3 DFM [115]. Two parallel reactors, one operating under CO2 capture and the other under CO2 methanation conditions, could lead to an optimal use of available resources, while prolonged operation did not appear to be detrimental for the material and its catalytic properties.
Due to its easy reduction once oxidized, ruthenium’s role is crucial for isothermal DFM applications and it can provide stable performance [22]. From a kinetic point of view, Ru-based DFMs (5% Ru) appearred to achieve faster hydrogenation rates, even compared to their Rh-based analogues (0.5% Rh), when Na2CO3 (or “Na2O”) was used as sorbent. Thus, Ru can be considered a superior DFM catalyst phase based on price and reaction kinetics criteria. Regarding the sorbent’s impact on the methanation kinetics, “Na2O”-type active sorbents appeared far superior and led to an overall higher CH4 yield, despite CaO sorbents initially providing a higher CO2 adsorption capacity [21].
The mechanism behind the two-step CO2 capture and methanation process over the 5% Ru, 10% Na2CO3/γ-Al2O3 (or 5% Ru, 6.1% “Na2O”/γ-Al2O3) DFM was highlighted by Proaño et al. [116] employing in-situ DRIFTS (Figure 9). After calcination and pre-reduction of the materials, Na2CO3 was decomposed and Na-rich surface, as well as Al-O-Na+ sites, were created and denoted as “Na2O”. On Na-free Ru/Al2O3, CO2 was chemisorbed at Ru sites and at –OH groups of the support, whereas on the Na-containing DFM, CO2 was chemisorbed at Al-O-Na+ sites in large quantities. The presence of Na+ on the surface caused CO2 to be primarily adsorbed in the form of chelated carbonato complexes, while the Ru and Al-OH sites could still accommodate adsorbed CO2, as in Ru/Al2O3. The adsorbed bidentate carbonates could then spillover to adjacent Ru catalytically active sites, where they were hydrogenated into CH4 via formate-type intermediates. Enhanced CH4 production for the Na-containing DFM was thus attributed to the larger CO2 capture capacity, due to the additional Al-O-Na+ sites that had a high affinity towards CO2.
Besides the Farrauto group, Bermejo-Lόpez et al. [24] performed a thorough work to describe the mechanism of CO2 capture and methanation, as well as to correlate the type and loading of the adsorbent phase with the material’s performance (Figure 10). Ru metal dispersion was one parameter dependant on the adsorbent type, with CaO seeming to hinder it, while Na2CO3 promoting it. The basicity of the materials increased with higher adsorbent loading and this promoted the CH4 yield. CaO provided strong basic sites, where CO2 was chemisorbed forming stable carbonates (CaCO3) that required higher temperatures to be desorbed. On the other hand, medium-strength basic sites over Na2CO3 led to the formation of less stable carbonates, enabling operation at lower temperatures. The proposed reaction mechanism involved an initial adsorption of CO2 with simultaneous release of H2O from the sorption active phases. During the following methanation step, adsorbed CO2 spilled over to Ru sites and was hydrogenated to methane, while a delay in the H2O signal could be attributed to some H2O-product being adsorbed by the basic sorbent sites and forming Ca(OH)2 and NaOH.
Cimino et al. [35] modified the γ-Al2O3 support with various alkali metals (Li, Na, K and Cs) and 1% Ru was chosen as the active metal phase. Regarding CO2 methanation performance, only Li exhibited a promotion effect and formed Li-aluminate spinel-phases that stabilized the alumina support and its interaction with Ru active sites. During cyclic CO2 capture and methanation at low temperature (230 °C), the Li-promoted material presented faster methanation kinetics and higher CH4 production compared to the Na-promoted one. It was noted that the temperature of the cyclic operation for the Li-promoted 1% Ru DFM (230 °C) was considerably lower compared with that for other DFMs that have higher Ru loadings.
Lastly, Sun et al. [25] employed a DFM consisting of a Ru supported on CeO2 nanorods methanation catalyst and an alkali nitrate promoted MgO sorbent mechanically mixed together (Figure 11). It has been already reported by Wang et al. [117] that the CO2 methanation mechanism over Ru/CeO2 follows a different pathway compared to Ru/Al2O3, with the oxygen vacancy catalyzed formate dissociation being the rate determining step in the former. Sun et al. [25] performed the integrated CO2 capture and methanation process in their DFM at 300 °C. CO2 was initially captured in the form of MgCO3. Subsequently, during the methanation process, CO2 was desorbed and activated by an oxygen vacancy in the CeO2 nanorod support, forming a formate intermediate. The formate was then decomposed into a ruthenium carbonyl and further hydrogenated into CH4 by nearby hydrogen atoms chemisorbed at the Ru surface. It was further noted that, although CH4 yield and CO2 conversion were higher for 10% Ru loading during the 1st cycle, the 5% Ru loaded catalyst showed better performance during the 10th cycle, probably due to the better preservation of the oxygen vacancies in the support.

3.2. Ni-Based Dual-Function Materials

Although Ni would be a more suitable choice as a methanation catalyst in DFMs, being considerably cheaper and more abundant compared to noble metals like Ru and Rh, there are a number of limitations that need to be carefully considered. Arellano-Treviño et al. [21] impregnated Ni on a γ-Al2O3 carrier that had been previously loaded with a Na2CO3 adsorbent and the Ni-based DFM underwent a pre-reduction at 650 °C for the catalyst activation. The authors showed that the Ni-based DFM (10% Ni, 6.1% “Na2O”/γ-Al2O3) operated cyclically at a constant 320 °C temperature could only produce substantial amounts of CH4, if the CO2 capture step was performed under an O2-free gas (7.5% CO2/N2). On the other hand, if the CO2 capture step was performed at 320 °C under a simulated flue-gas (7.5% CO2, 4.5% O2, 15% H2O, balance N2), no CH4 was produced during the subsequent hydrogenation step at the same temperature. This was explained by the inability of NiO to be reactivated under 15% H2/N2 at 320 °C in order for the whole operation to proceed isothermally. Furthermore, Ni-based DFMs retained a certain amount of preadsorbed CO2 that was not further converted.
Following their earlier work, Arellano-Treviño et al. [23] attempted to enhance the Ni-based material’s reducibility by introducing small amounts (0.1 and 1 wt.%) of noble metals Ru, Pt and Pd. Pd-containing samples exhibited comparatively inferior performance, whereas the beneficial effect of Pt and Ru addition rested on the enhancement of NiO reducibility under 13.26% H2/N2 flow at 320 °C. 1 wt.% Pt addition caused 50% of the initially oxidized NiO to be reduced, while this value for the promotion with 1 wt.% Ru reached 70%. The authors speculated that PtOx and RuOx species reduced rapidly and the exposed Pt and Ru active surfaces splitted H2 molecules and helped reduce the adjacent NiO sites. For cyclic operation at 320 °C, the 1% Ru, 10% Ni, 6.1% “Na2O”/γ-Al2O3 material exhibited great cyclic stability and the highest CH4 production, since both Ru and Ni sites are active for CO2 methanation. Compared to the Ni-free DFM with the same Ru loading (1% Ru, 6.1% “Na2O”/γ-Al2O3), the Ni-containing one (1% Ru, 10% Ni, 6.1% “Na2O”/γ-Al2O3) exhibited an increased CO2 adsorption capacity and CH4 yield, at a cost, however, of slower methanation kinetics. As such, the synergy between Ni and Ru metals promoted the material’s performance, in agreement with other studies regarding CO2 methanation [75,118].
Other works typically perform the CO2 capture step with an O2-free gas mixture of CO2 balanced with a carrier gas. Bermejo-Lόpez et al. [112] studied the effects of Ni, as well as CaO and Na2CO3 loadings on the DFM performance (Figure 12). The addition of the adsorbent phase weakened the metal-support interaction and enhanced the reducibility of the Ni metallic phase. Similarly to their previous work [24], the authors showed that CaO adsorbent led to the creation of highly stable carbonates that required higher temperatures to decompose, while less stable carbonates over Na2CO3 decomposed at a lower temperature. Thus, during the following hydrogenation step, CH4 formation was favoured at higher temperatures when CaO was used as the active sorbent, peaking at 520 °C, while maximum CH4 was produced for Na2CO3 at 400 °C. Increasing the Ni loading resulted in increased CO2 sorption capacity and CH4 production. Regarding the proposed mechanism, CO2 was initially adsorbed over NaOH and Ca(OH)2 species, with H2O being simultaneously released. During the subsequent H2/N2 inflow, adsorbed CO2 was reduced after its spillover to Ni sites and detected product H2O was delayed due to its initial adsorption by the sorbent phases. The DFM with CaO sorbent (10NiCa) exhibited a higher CH4 selectivity upon hydrogenation during cyclic operation at 400 °C, while the DFM with Na2CO3 sorbent (10NiNa) produced more CO arising from the RWGS reaction.
Chai et al. [119] also tested the effects of different CO2 adsorbents (MgO, CaO, Na2CO3 and K2CO3) on Ni-based DFMs supported on Al2O3. The catalyst and adsorbent loadings were the same, at 10 wt.%, while the CO2 capture step was performed under an O2-free 9.5% CO2/N2 gas flow. In a similar fashion to the findings of Bermejo-Lόpez et al. [23,102], the authors showed that CaO sorbent led to greater CO2 sorption capacity. However, the release of adsorbed CO2 in the material with CaO sorbent was not fully reversible. The material with Na2CO3 sorbent exhibited a high CO2 capture capacity, along with a high release efficiency. Hence, methanation capacity for the material with Na2CO3 sorbent was superior, a property attributed to the easier release and spillover of adsorbed CO2 towards the Ni catalytically active sites.
Hu et al. [120] prepared Ni-based DFMs supported on ZrO2 and modified with La and K. The use of K and La promoters raised the CO2 sorption capacity of the materials, with the K-modified material increasing its CO2 capture efficiency four-fold, compared with the unmodified one. K-promoted Ni/ZrO2 released CH4 upon hydrogenation much slower, while the La-promoted material exhibited faster methanation kinetics, with the time required for CH4 release roughly matching the time required for CO2 capture. The methanation mechanism was modified upon K-promotion and additional less-reducible intermediate bidentate carbonates were formed, whereas the reaction over unpromoted and La-modified Ni/ZrO2 proceeded over active formate intermediates. It could be concluded, that La-promotion was more suited for operation at lower temperatures (250 °C) by achieving favourable kinetics, while K-modification favoured increased CO2 chemisorption and methanation efficiency at higher temperatures (>350 °C).
Finally, Zhou et al. [121] prepared a 2D-layered Ni/MgO-Al2O3 material to be used during the integrated CO2 capture and methanation (Figure 13). The layered oxides resulted after calcination and reduction of the initial hydrotalcite (or layered double hydroxide) precursors, which were in turn prepared via the coprecipitation method. By increasing the Ni content in the material, the interaction between Ni and the MgO-Al2O3 support weakened and Ni species became more reducible. For the cyclic CO2 capture and methanation process, the authors aimed at optimizing the respective durations, so that CO2 breakthrough was minimal and a maximum utilization of reactant H2 was achieved. Complete capture and methanation could be achieved by a 35 s CO2 sorption time and a 25 s methanation time. Upon conducting in-situ FTIR experiments, monodentate and bidentate carbonates, as well as formate species were detected during the CO2 sorption phase. Following 20% H2/Ar flow, the observed bands suggested a gradual removal of these species via hydrogenation to CH4, with no CO being detected. Further development of materials derived from hydrotalcite precursors containing Mg and Ni could be favourable for DFM applications, since Ni can be highly dispersed and stabilized, while a high Mg content can provide a plethora of surface basic sites for CO2 chemisorption [90]. Table 2 summarizes the performance of DFMs for the combined capture and methanation of CO2.

3.3. Dual-Function Materials for Products Other than Methane

Apart from synthetic natural gas (SNG) obtained during methanation, the concept of DFMs for the integrated CO2 capture and conversion has been extended for the production of CO and syngas, that can be used for the further production of other value-added products. Bobadilla et al. [122] prepared a K-promoted trimetallic FeCrCu catalyst supported on MgO-Al2O3. The K-component was added to increase the CO2 sorption capacity, while the combination of three earth-abundant metals (Fe, Cr and Cu) favoured CO selectivity via RWGS. Hyakutake et al. [123] promoted a Cu/Al2O3 catalyst with K and Ba for CO2 capture and conversion to CO. K-promotion facilitated Cu dispersion and resulted in a very high capacity for CO2 capture as formate intermediates. These intermediates were then selectively reduced to CO via the RWGS route. The Ba-promoted material captured much less CO2 in comparison and thus exhibited an inferior performance as DFM for CO production.
Al-Mamoori et al. [124] introduced a new feature via the use of K and Na-promoted CaO and MgO-based double salts as the CO2 adsorbing components, in an attempt to greatly increase the CO2 capture efficiency and sorption kinetics. The K-Ca, Na-Ca, K-Mg and Na-Mg double salts were supported on γ-Al2O3 via the impregnation of alkalimetal carbonates and alkaline earth metal nitrates, whereas Ni was impregnated subsequently. After the CO2 capture step, ethane was flowed into the reactor in order to produce syngas via the dry reforming of ethane (DER) reaction.
Additionally, Sun et al. [125] employed a common CO2 sorbent, CaO, directly as support material for Ni nanocatalysts with adjacent Ce-species. They argued that the already established calcium-looping process that rests on the carbonization and calcination of CaO for CO2 capture and sequestration (CCS) could be modified through the integration of a CO2 utilization step by converting CO2 to CO. The incorporation of Ni facilitated CO2 reduction, whereas Ce-species provided oxygen vacancies that were active sites for CO2 dissociation and prevented the agglomeration of individual CaO and Ni nanocrystallites.

3.4. Configuartions Using a Seperate Sorbent and Methanation Catalyst

Apart from the use of a single solid that incorporates both sorption and methanation functions, the process of integrated CO2 capture and conversion to methane can be realized via the combination of two different solids that perform these functions individually. Veselovskaya et al. published a number of works that incorporated the use of a K2CO3/Al2O3 sorbent and a Ru/Al2O3 methanation catalyst [126,127,128]. The high CO2 sorption capabilities of the K2CO3-based sorbent enabled it to capture CO2 directly from ambient air (direct air capture or DAC), where its concentration is as low as 400 ppm. The sorbent could be regenerated at the mild temperature of 200 °C under H2 flow and CH4 was produced at a subsequent catalytic reactor that contained the Ru/Al2O3 methanation catalyst. The whole operation was coined as direct air capture and methanation (DACM) process [128].
Finally, Miguel et al. [129] also operated the CO2 capture and conversion to SNG process by using a single reactor that contained alternating layers of the sorbent and catalyst materials. CO2 sorption was performed by a K-promoted hydrotalcite, whereas a commercial Ni-rich material was used as the methanation catalyst. It was shown that the whole amount of captured CO2 could be fully hydrogenated into CH4 and that CH4 production under cyclic operation was quite high, peaking at 2.36 mol CH4 per kilogram material and hour of operation.

4. Conclusions

Given the continuous increase of humanity’s energy needs and the depletion of our fossil fuel deposits, better management of existing carbon resources via the cyclic conversion between carbon’s highest (CO2) and lowest (CH4) oxidation state is of particular importance towards the transition to a renewable energy-based economy.
Alkali and alkaline earth metal addition appears to greatly impact the CO2 methanation activity of supported metal catalysts. Their promoting effect mainly rests on the enhancement of support alkalinity that promotes CO2 chemisorption, the generation of oxygen vacancies on the surface of metal oxide supports and the improvement of active metal dispersion. Being extremely cheap and earth-abundant, group I and II metals seem to be a viable choice to for use as catalyst promoters towards enhancing the low-temperature activity of Ni-based catalysts, or allowing for a lower noble metal loading (e.g., Ru), without compromising the catalytic activity.
It is, however, evident that, under specific conditions, alkali and sometimes alkaline earth metal addition can suppress methanation activity. This can be achieved by the masking of some active metal sites, for example when impregnating the alkaline modifier after the active metal, or by a detrimental change in the reaction pathway, for example, through a strong interaction with formate intermediates or the increased production of easy-to-desorb CO. More detailed studies on the effect of different alkali and alkaline earth metal promoters on already active methanation catalysts (such as benchmark Ni/CeO2 and Ni/Al2O3) should be undertaken to further clarify their promotion function. In general, it appears that alkaline earths constitute reliable promoters, while the existing literature on promotion with heavier group II metals (Sr and Ba) is comparatively scarce.
The development of new dual-function materials (DFMs) functioning as both catalysts and adsorbents towards the combined CO2 capture and methanation appears capable of improving the process and as a result, research efforts towards the preparation of such novel materials have boomed during the last 5 years. The main issues that need to be resolved concern the enhancement of the CO2 sorption capacity and CH4 production (mostly measured in mmol CH4/g DFM), as well as the active metal stability under industrial-like flue-gas applications.
However, since the field is quite new and has great potential, many possible research directions do exist. These could include the comparative study of different support materials (e.g., Al2O3, SiO2, ZrO2, CeO2 and mesoporous silicas), the employment of additional catalyst synthesis methods (e.g., hydrotalcite-derived catalysts by coprecipitation), the further use of other promoter elements that favour CO2 adsorption (e.g., La, Sr) and the fabrication of DFMs with stable bimetallic nickel-based catalysts with other transition metals (e.g., NiFe and NiCo alloys).

Author Contributions

Conceptualization, A.I.T.; Data curation, A.I.T.; Formal analysis, A.I.T.; Funding acquisition, I.V.Y., M.A.G.; Investigation, Methodology, M.A.G.; Project administration, M.A.G., I.V.Y.; Project coordination, I.V.Y.; Resources, M.A.G.; Supervision, N.D.C.; Writing—original draft, A.I.T., N.D.C.; Writing—review & editing, N.D.C., I.V.Y. and M.A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been co-financed by the European Union and Greek national funds through the operational program ‘Regional Excellence’ and the operational program Competitiveness, Entrepreneurship and Innovation, under the call Research - Create - Innovate (Project code: T1EDK-00782). The authors are also grateful to the Research Committee of UOWM for financial support through grant no. 70304.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mardani, A.; Streimikiene, D.; Cavallaro, F.; Loganathan, N.; Khoshnoudi, M. Carbon dioxide (CO2) emissions and economic growth: A systematic review of two decades of research from 1995 to 2017. Sci. Total Environ. 2019, 649, 31–49. [Google Scholar] [CrossRef]
  2. Szulejko, J.E.; Kumar, P.; Deep, A.; Kim, K.H. Global warming projections to 2100 using simple CO2 greenhouse gas modeling and comments on CO2 climate sensitivity factor. Atmos. Pollut. Res. 2017, 8, 136–140. [Google Scholar] [CrossRef]
  3. Yentekakis, I.V.; Dong, F. Grand challenges for Catalytic Remediation in environmental and energy applications towards a cleaner and sustainable future. Front. Environ. Chem. 2020, 1, 5. [Google Scholar] [CrossRef]
  4. L’Orange Seigo, S.; Dohle, S.; Siegrist, M. Public perception of carbon capture and storage (CCS): A review. Renew. Sustain. Energy Rev. 2014, 38, 848–863. [Google Scholar] [CrossRef]
  5. Al-Mamoori, A.; Krishnamurthy, A.; Rownaghi, A.A.; Rezaei, F. Carbon capture and utilization update. Energy Technol. 2017, 5, 834–849. [Google Scholar] [CrossRef] [Green Version]
  6. Li, W.; Wang, H.; Jiang, X.; Zhu, J.; Liu, Z.; Guo, X.; Song, C. A short review of recent advances in CO2 hydrogenation to hydrocarbons over heterogeneous catalysts. RSC Adv. 2018, 8, 7651–7669. [Google Scholar] [CrossRef] [Green Version]
  7. Thema, M.; Bauer, F.; Sterner, M. Power-to-Gas: Electrolysis and methanation status review. Renew. Sustain. Energy Rev. 2019, 112, 775–787. [Google Scholar] [CrossRef]
  8. Frontera, P.; Macario, A.; Ferraro, M.; Antonucci, P.L. Supported catalysts for CO2 methanation: A review. Catalysts 2017, 7, 59. [Google Scholar] [CrossRef]
  9. Vogt, C.; Monai, M.; Kramer, G.J.; Weckhuysen, B.M. The renaissance of the Sabatier reaction and its applications on Earth and in space. Nat. Catal. 2019, 2, 188–197. [Google Scholar] [CrossRef]
  10. Daza, Y.A.; Kuhn, J.N. CO2 conversion by reverse water gas shift catalysis: Comparison of catalysts, mechanisms and their consequences for CO2 conversion to liquid fuels. RSC Adv. 2016, 6, 49675–49691. [Google Scholar] [CrossRef]
  11. Ye, R.P.; Ding, J.; Gong, W.; Argyle, M.D.; Zhong, Q.; Wang, Y.; Russell, C.K.; Xu, Z.; Russell, A.G.; Li, Q.; et al. CO2 hydrogenation to high-value products via heterogeneous catalysis. Nat. Commun. 2019, 10, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Vogt, C.; Monai, M.; Sterk, E.B.; Palle, J.; Melcherts, A.E.M.; Zijlstra, B.; Groeneveld, E.; Berben, P.H.; Boereboom, J.M.; Hensen, E.J.M.; et al. Understanding carbon dioxide activation and carbon–carbon coupling over nickel. Nat. Commun. 2019, 10, 1–10. [Google Scholar] [CrossRef] [PubMed]
  13. Miao, B.; Ma, S.S.K.; Wang, X.; Su, H.; Chan, S.H. Catalysis mechanisms of CO2 and CO methanation. Catal. Sci. Technol. 2016, 6, 4048–4058. [Google Scholar] [CrossRef]
  14. Cárdenas-Arenas, A.; Quindimil, A.; Davó-Quiñonero, A.; Bailón-García, E.; Lozano-Castelló, D.; De-La-Torre, U.; Pereda-Ayo, B.; González-Marcos, J.A.; González-Velasco, J.R.; Bueno-López, A. Isotopic and in situ DRIFTS study of the CO2 methanation mechanism using Ni/CeO2 and Ni/Al2O3 catalysts. Appl. Catal. B Environ. 2020, 265, 118538. [Google Scholar] [CrossRef]
  15. Tada, S.; Shimizu, T.; Kameyama, H.; Haneda, T.; Kikuchi, R. Ni/CeO2 catalysts with high CO2 methanation activity and high CH4 selectivity at low temperatures. Int. J. Hydrog. Energy 2012, 37, 5527–5531. [Google Scholar] [CrossRef]
  16. Boaro, M.; Colussi, S.; Trovarelli, A. Ceria-based materials in hydrogenation and reforming reactions for CO2 valorization. Front. Chem. 2019, 7, 28. [Google Scholar] [CrossRef]
  17. Choi, S.; Drese, J.H.; Jones, C.W. Adsorbent materials for carbon dioxide capture from large anthropogenic point sources. ChemSusChem 2009, 2, 796–854. [Google Scholar] [CrossRef]
  18. Kierzkowska, A.M.; Pacciani, R.; Müller, C.R. CaO-based CO2 sorbents: From fundamentals to the development of new, highly effective materials. ChemSusChem 2013, 6, 1130–1148. [Google Scholar] [CrossRef]
  19. Duyar, M.S.; Arellano-Treviño, M.A.; Farrauto, R.J. Dual function materials for CO2 capture and conversion using renewable H2. Appl. Catal. B Environ. 2015, 168–169, 370–376. [Google Scholar] [CrossRef]
  20. Melo Bravo, P.; Debecker, D.P. Combining CO2 capture and catalytic conversion to methane. Waste Dispos. Sustain. Energy 2019, 1, 53–65. [Google Scholar] [CrossRef] [Green Version]
  21. Arellano-Treviño, M.A.; He, Z.; Libby, M.C.; Farrauto, R.J. Catalysts and adsorbents for CO2 capture and conversion with dual function materials: Limitations of Ni-containing DFMs for flue gas applications. J. CO2 Util. 2019, 31, 143–151. [Google Scholar] [CrossRef]
  22. Wang, S.; Schrunk, E.T.; Mahajan, H.; Farrauto, R.J. The role of ruthenium in CO2 capture and catalytic conversion to fuel by dual function materials (DFM). Catalysts 2017, 7, 88. [Google Scholar] [CrossRef] [Green Version]
  23. Arellano-Treviño, M.A.; Kanani, N.; Jeong-Potter, C.W.; Farrauto, R.J. Bimetallic catalysts for CO2 capture and hydrogenation at simulated flue gas conditions. Chem. Eng. J. 2019, 375, 121953. [Google Scholar] [CrossRef]
  24. Bermejo-López, A.; Pereda-Ayo, B.; González-Marcos, J.A.; González-Velasco, J.R. Mechanism of the CO2 storage and in situ hydrogenation to CH4. Temperature and adsorbent loading effects over Ru-CaO/Al2O3 and Ru-Na2CO3/Al2O3 catalysts. Appl. Catal. B Environ. 2019, 256, 117845. [Google Scholar] [CrossRef]
  25. Sun, H.; Zhang, Y.; Guan, S.; Huang, J.; Wu, C. Direct and highly selective conversion of captured CO2 into methane through integrated carbon capture and utilization over dual functional materials. J. CO2 Util. 2020, 38, 262–272. [Google Scholar] [CrossRef]
  26. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Ahmad, A. CO2 methanation over heterogeneous catalysts: Recent progress and future prospects. Green Chem. 2015, 17, 2647–2663. [Google Scholar] [CrossRef]
  27. Lee, W.J.; Li, C.; Prajitno, H.; Yoo, J.; Patel, J.; Yang, Y.; Lim, S. Recent trend in thermal catalytic low temperature CO2 methanation: A critical review. Catal. Today 2020, in press. [Google Scholar] [CrossRef]
  28. Lv, C.; Xu, L.; Chen, M.; Cui, Y.; Wen, X.; Li, Y.; Wu, C.E.; Yang, B.; Miao, Z.; Hu, X.; et al. Recent Progresses in Constructing the Highly Efficient Ni Based Catalysts With Advanced Low-Temperature Activity Toward CO2 Methanation. Front. Chem. 2020, 8, 1–32. [Google Scholar] [CrossRef]
  29. Charisiou, N.D.; Siakavelas, G.; Tzounis, L.; Sebastian, V.; Monzon, A.; Baker, M.A.; Hinder, S.J.; Polychronopoulou, K.; Yentekakis, I.V.; Goula, M.A. An in depth investigation of deactivation through carbon formation during the biogas dry reforming reaction for Ni supported on modified with CeO2 and La2O3 zirconia catalysts. Int. J. Hydrog. Energy 2018, 43, 18955–18976. [Google Scholar] [CrossRef]
  30. Liu, K.; Xu, X.; Xu, J.; Fang, X.; Liu, L.; Wang, X. The distributions of alkaline earth metal oxides and their promotional effects on Ni/CeO2 for CO2 methanation. J. CO2 Util. 2020, 38, 113–124. [Google Scholar] [CrossRef]
  31. Charisiou, N.D.; Baklavaridis, A.; Papadakis, V.G.; Goula, M.A. Synthesis Gas Production via the Biogas Reforming Reaction Over Ni/MgO–Al2O3 and Ni/CaO–Al2O3 Catalysts. Waste Biomass Valorization 2016, 7, 725–736. [Google Scholar] [CrossRef]
  32. Yentekakis, I.V.; Vernoux, P.; Goula, G.; Caravaca, A. Electropositive promotion by alkalis or alkaline earths of Pt-group metals in emissions control catalysis: A status report. Catalysts 2019, 9, 157. [Google Scholar] [CrossRef] [Green Version]
  33. Petala, A.; Panagiotopoulou, P. Methanation of CO2 over alkali-promoted Ru/TiO2 catalysts: I. Effect of alkali additives on catalytic activity and selectivity. Appl. Catal. B Environ. 2018, 224, 919–927. [Google Scholar] [CrossRef]
  34. Panagiotopoulou, P. Methanation of CO2 over alkali-promoted Ru/TiO2 catalysts: II. Effect of alkali additives on the reaction pathway. Appl. Catal. B Environ. 2018, 236, 162–170. [Google Scholar] [CrossRef]
  35. Cimino, S.; Boccia, F.; Lisi, L. Effect of alkali promoters (Li, Na, K) on the performance of Ru/Al2O3 catalysts for CO2 capture and hydrogenation to methane. J. CO2 Util. 2020, 37, 195–203. [Google Scholar] [CrossRef]
  36. Liang, C.; Ye, Z.; Dong, D.; Zhang, S.; Liu, Q.; Chen, G.; Li, C.; Wang, Y.; Hu, X. Methanation of CO2: Impacts of modifying nickel catalysts with variable-valence additives on reaction mechanism. Fuel 2019, 254, 115654. [Google Scholar] [CrossRef]
  37. Zhang, Z.; Zhang, X.; Zhang, L.; Gao, J.; Shao, Y.; Dong, D.; Zhang, S.; Liu, Q.; Xu, L.; Hu, X. Impacts of alkali or alkaline earth metals addition on reaction intermediates formed in methanation of CO2 over cobalt catalysts. J. Energy Inst. 2020, 93, 1581–1596. [Google Scholar] [CrossRef]
  38. Vayenas, C.G.; Bebelis, S.; Yentekakis, I.V.; Lintz, H.G. Non-faradaic electrochemical modification of catalytic activity: A status report. Catal. Today 1992, 11, 303–438. [Google Scholar] [CrossRef]
  39. Klaitzidou, I.; Makri, M.; Theleritis, D.; Katsaounis, A.; Vayenas, C.G. Comparative study of the electrochemical promotion of CO2 hydrogenation on Ru using Na+, K+, H+ and O2- conducting solid electrolytes. Surf. Sci. 2016, 646, 194–203. [Google Scholar] [CrossRef]
  40. Theleritis, D.; Makri, M.; Souentie, S.; Caravaca, A.; Κatsaounis, A.; Vayenas, C.G. Comparative study of the electrochemical promotion of CO2 hydrogenation over Ru-supported catalysts using electronegative and electropositive promoters. ChemElectroChem 2014, 1, 254–262. [Google Scholar] [CrossRef]
  41. Le, T.A.; Kim, T.W.; Lee, S.H.; Park, E.D. Effects of Na content in Na/Ni/SiO2 and Na/Ni/CeO2 catalysts for CO and CO2 methanation. Catal. Today 2018, 303, 159–167. [Google Scholar] [CrossRef]
  42. Beierlein, D.; Häussermann, D.; Pfeifer, M.; Schwarz, T.; Stöwe, K.; Traa, Y.; Klemm, E. Is the CO2 methanation on highly loaded Ni-Al2O3 catalysts really structure-sensitive? Appl. Catal. B Environ. 2019, 247, 200–219. [Google Scholar] [CrossRef]
  43. Büchel, R.; Baiker, A.; Pratsinis, S.E. Effect of Ba and K addition and controlled spatial deposition of Rh in Rh/Al2O3 catalysts for CO2 hydrogenation. Appl. Catal. A Gen. 2014, 477, 93–101. [Google Scholar] [CrossRef]
  44. Iloy, R.A.; Jalama, K. Effect of Operating Temperature, Pressure and Pottasium Loading on the Performance of Silica-Supported Cobalt Catalyst in CO2 Hydrogenation to Hydrocarbon Fuel. Catalysts 2019, 9, 807. [Google Scholar] [CrossRef] [Green Version]
  45. Zhang, Z.; Hu, X.; Wang, Y.; Hu, S.; Xiang, J.; Li, C.; Chen, G.; Liu, Q.; Wei, T.; Dong, D. Regulation the reaction intermediates in methanation reactions via modification of nickel catalysts with strong base. Fuel 2019, 237, 566–579. [Google Scholar] [CrossRef]
  46. He, L.; Lin, Q.; Liu, Y.; Huang, Y. Unique catalysis of Ni-Al hydrotalcite derived catalyst in CO2 methanation: Cooperative effect between Ni nanoparticles and a basic support. J. Energy Chem. 2014, 23, 587–592. [Google Scholar] [CrossRef]
  47. Konsolakis, M.; Yentekakis, I.V. Strong promotional effects of Li, K, Rb and Cs on the Pt-catalysed reduction of NO by propene. Appl. Catal. B 2001, 29, 103–113. [Google Scholar] [CrossRef] [Green Version]
  48. Yentekakis, I.V.; Konsolakis, M.; Lambert, R.M.; Palermo, A.; Tikhov, M. Successful application of the electrochemical promotion to the design of effective conventional catalyst formulations. Solid State Ion. 2000, 136, 783–790. [Google Scholar] [CrossRef]
  49. Vogt, C.; Wijten, J.; Madeira, C.L.; Kerkenaar, O.; Xu, K.; Holzinger, R.; Monai, M.; Weckhuysen, B.M. Alkali Promotion in the Formation of CH4 from CO2 and Renewably Produced H2 over Supported Ni Catalysts. ChemCatChem 2020, 12, 2792–2800. [Google Scholar] [CrossRef]
  50. Yang, Q.; Liu, G.; Liu, Y. Perovskite-Type Oxides as the Catalyst Precursors for Preparing Supported Metallic Nanocatalysts: A Review. Ind. Eng. Chem. Res. 2018, 57, 1–17. [Google Scholar] [CrossRef]
  51. Tsounis, C.; Wang, Y.; Arandiyan, H.; Wong, R.J.; Toe, C.Y.; Amal, R.; Scott, J. Tuning the Selectivity of LaNiO3 Perovskites for CO2 hydrogenation through potassium substitution. Catalysts 2020, 10, 409. [Google Scholar] [CrossRef] [Green Version]
  52. Guo, M.; Lu, G. The difference of roles of alkaline-earth metal oxides on silica-supported nickel catalysts for CO2 methanation. RSC Adv. 2014, 4, 58171–58177. [Google Scholar] [CrossRef]
  53. Liang, C.; Hu, X.; Wei, T.; Jia, P.; Zhang, Z.; Dong, D.; Zhang, S.; Liu, Q.; Hu, G. Methanation of CO2 over Ni/Al2O3 modified with alkaline earth metals: Impacts of oxygen vacancies on catalytic activity. Int. J. Hydrog. Energy 2019, 44, 8197–8213. [Google Scholar] [CrossRef]
  54. Yentekakis, I.V.; Goula, G.; Hatzisymeon, M.; Betsi-Argyropoulou, I.; Botzolaki, G.; Kousi, K.; Kondarides, D.I.; Taylor, M.J.; Parlett, C.M.A.; 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. [Google Scholar] [CrossRef] [Green Version]
  55. Goula, M.A.; Charisiou, N.D.; Siakavelas, G.; Tzounis, L.; Tsiaoussis, I.; Panagiotopoulou, P.; Goula, G.; Yentekakis, I.V. Syngas production via the biogas dry reforming reaction over Ni supported on zirconia modified with CeO2 or La2O3 catalysts. Int. J. Hydrog. Energy 2017, 42, 13724–13740. [Google Scholar] [CrossRef]
  56. Huang, J.; Li, X.; Wang, X.; Fang, X.; Wang, H.; Xu, X. New insights into CO2 methanation mechanisms on Ni/MgO catalysts by DFT calculations: Elucidating Ni and MgO roles and support effects. J. CO2 Util. 2019, 33, 55–63. [Google Scholar] [CrossRef]
  57. Li, Y.; Lu, G.; Ma, J. Highly active and stable nano NiO-MgO catalyst encapsulated by silica with a core-shell structure for CO2 methanation. RSC Adv. 2014, 4, 17420–17428. [Google Scholar] [CrossRef]
  58. Park, J.N.; McFarland, E.W. A highly dispersed Pd-Mg/SiO2 catalyst active for methanation of CO2. J. Catal. 2009, 266, 92–97. [Google Scholar] [CrossRef]
  59. Kim, H.Y.; Lee, H.M.; Park, J.N. Bifunctional mechanism of CO2 methanation on Pd-MgO/SiO2 catalyst: Independent roles of MgO and Pd on CO2 methanation. J. Phys. Chem. C 2010, 114, 7128–7131. [Google Scholar] [CrossRef]
  60. Kwak, J.H.; Kovarik, L.; Szanyi, J. Heterogeneous catalysis on atomically dispersed supported metals: CO2 reduction on multifunctional Pd catalysts. ACS Catal. 2013, 3, 2094–2100. [Google Scholar] [CrossRef]
  61. Alcalde-Santiago, V.; Davó-Quiñonero, A.; Lozano-Castelló, D.; Quindimil, A.; De-La-Torre, U.; Pereda-Ayo, B.; González-Marcos, J.A.; González-Velasco, J.R.; Bueno-López, A. Ni/LnOx catalysts (Ln = La, Ce or Pr) for CO2 methanation. ChemCatChem 2019, 11, 810–819. [Google Scholar]
  62. Cárdenas-Arenas, A.; Quindimil, A.; Davó-Quinonero, A.; Bailón-García, E.; Lozano-Castelló, D.; De-La-Torre, U.; Pereda-Ayo, B.; González-Marcos, J.A.; González-Velasco, J.R.; Bueno-López, A. Design of active sites in Ni/CeO2 catalysts for the methanation of CO2: Tailoring the Ni-CeO2 contact. Appl. Mater. Today 2020, 19, 100591. [Google Scholar] [CrossRef]
  63. Guo, M.; Lu, G. The effect of impregnation strategy on structural characters and CO2 methanation properties over MgO modified Ni/SiO2 catalysts. Catal. Commun. 2014, 54, 55–60. [Google Scholar] [CrossRef]
  64. Tan, J.; Wang, J.; Zhang, Z.; Ma, Z.; Wang, L.; Liu, Y. Highly dispersed and stable Ni nanoparticles confined by MgO on ZrO2 for CO2 methanation. Appl. Surf. Sci. 2019, 481, 1538–1548. [Google Scholar] [CrossRef]
  65. Xu, L.; Wang, F.; Chen, M.; Yang, H.; Nie, D.; Qi, L.; Lian, X. Alkaline-promoted Ni based ordered mesoporous catalysts with enhanced low-temperature catalytic activity toward CO2 methanation. RSC Adv. 2017, 7, 18199–18210. [Google Scholar] [CrossRef] [Green Version]
  66. Li, X.; Wang, Y.; Zhang, G.; Sun, W.; Bai, Y.; Zheng, L.; Han, X.; Wu, L. Influence of Mg-promoted Ni-based Catalyst Supported on Coconut Shell Carbon for CO2 Methanation. ChemistrySelect 2019, 4, 838–845. [Google Scholar] [CrossRef]
  67. Xiao-liu, W.; Meng, Y.; Ling-jun, Z.; Xiao-nan, Z.; Shu-rong, W. CO2 methanation over Ni /Mg@MCM-41 prepared by in-situ synthesis method. J. Fuel Chem. Technol. 2020, 48, 456–465. [Google Scholar]
  68. Bacariza, M.C.; Graça, I.; Bebiano, S.S.; Lopes, J.M.; Henriques, C. Magnesium as Promoter of CO2 Methanation on Ni-Based USY Zeolites. Energy Fuels 2017, 31, 9776–9789. [Google Scholar] [CrossRef] [Green Version]
  69. Yang, H.; Xu, L.; Chen, M.; Lv, C.; Cui, Y.; Wen, X.; Wu, C.E.; Yang, B.; Miao, Z.; Hu, X.; et al. Facilely fabricating highly dispersed Ni-based catalysts supported on mesoporous MFI nanosponge for CO2 methanation. Microporous Mesoporous Mater. 2020, 302, 110250. [Google Scholar] [CrossRef]
  70. Zhang, L.; Bian, L.; Zhu, Z.; Li, Z. La-promoted Ni/Mg-Al catalysts with highly enhanced low-temperature CO2 methanation performance. Int. J. Hydrog. Energy 2018, 43, 2197–2206. [Google Scholar] [CrossRef]
  71. Jingyu, Z.; Hongfang, M.; Fangyu, J.; Haitao, Z.; Weiyong, Y. Mn and Mg dual promoters modified Ni/α-Al2O3 catalysts for high temperature syngas methanation. Fuel Process. Technol. 2018, 172, 225–232. [Google Scholar]
  72. Baysal, Z.; Kureti, S. CO2 methanation on Mg-promoted Fe catalysts. Appl. Catal. B Environ. 2020, 262, 118300. [Google Scholar] [CrossRef]
  73. Kirchner, J.; Baysal, Z.; Kureti, S. Activity and Structural Changes of Fe-based Catalysts during CO2 Hydrogenation towards CH4—A Mini Review. ChemCatChem 2020, 12, 981–988. [Google Scholar] [CrossRef] [Green Version]
  74. Xu, L.; Yang, H.; Chen, M.; Wang, F.; Nie, D.; Qi, L.; Lian, X.; Chen, H.; Wu, M. CO2 methanation over Ca doped ordered mesoporous Ni-Al composite oxide catalysts: The promoting effect of basic modifier. J. CO2 Util. 2017, 21, 200–210. [Google Scholar] [CrossRef]
  75. Liu, Q.; Wang, S.; Zhao, G.; Yang, H.; Yuan, M.; An, X.; Zhou, H.; Qiao, Y.; Tian, Y. CO2 methanation over ordered mesoporous NiRu-doped CaO-Al2O3 nanocomposites with enhanced catalytic performance. Int. J. Hydrog. Energy 2018, 43, 239–250. [Google Scholar] [CrossRef]
  76. Do, P.N.U.; Luu, C.L.; Nguyen, P.A.; Van Nguyen, T.T.; Hoang, T.C.; Nguyen, T. Promotion effect of CO on methanation of CO2-rich gas over nanostructured modified NiO/γ-Al2O3 catalysts. Adv. Nat. Sci. Nanosci. Nanotechnol. 2019, 10, 035004. [Google Scholar] [CrossRef]
  77. Yang, W.; Feng, Y.; Chu, W. Promotion effect of CaO modification on mesoporous Al2O3-supported Ni catalysts for CO2 methanation. Int. J. Chem. Eng. 2016, 2016. 2041821. [Google Scholar] [CrossRef] [Green Version]
  78. Hu, X.F.; Yang, W.; Wang, N.; Luo, S.Z.; Chu, W. Catalytic properties of Ni/CNTs and Ca-Promoted Ni/CNTs for methanation reaction of carbon dioxide. Adv. Mater. Res. 2014, 924, 217–226. [Google Scholar] [CrossRef]
  79. Feng, Y.; Yang, W.; Chu, W. Effect of Ca modification on the catalytic performance of Ni/AC for CO2 methanation. Integr. Ferroelectr. 2016, 172, 40–48. [Google Scholar] [CrossRef]
  80. Zhou, M.; Ahmad, A. Synthesis, processing and characterization of calcia-stabilized zirconia solid electrolytes for oxygen sensing applications. Mater. Res. Bull. 2006, 41, 690–696. [Google Scholar] [CrossRef]
  81. Takano, H.; Shinomiya, H.; Izumiya, K.; Kumagai, N.; Habazaki, H.; Hashimoto, K. CO2 methanation of Ni catalysts supported on tetragonal ZrO2 doped with Ca2+ and Ni2+ ions. Int. J. Hydrog. Energy 2015, 40, 8347–8355. [Google Scholar] [CrossRef]
  82. Everett, O.E.; Zonetti, P.C.; Alves, O.C.; de Avillez, R.R.; Appel, L.G. The role of oxygen vacancies in the CO2 methanation employing Ni/ZrO2 doped with Ca. Int. J. Hydrog. Energy 2020, 45, 6352–6359. [Google Scholar] [CrossRef]
  83. Jia, C.; Gao, J.; Li, J.; Gu, F.; Xu, G.; Zhong, Z.; Su, F. Nickel catalysts supported on calcium titanate for enhanced CO methanation. Catal. Sci. Technol. 2013, 3, 490–499. [Google Scholar] [CrossRef]
  84. Do, J.Y.; Park, N.K.; Seo, M.W.; Lee, D.; Ryu, H.J.; Kang, M. Effective thermocatalytic carbon dioxide methanation on Ca-inserted NiTiO3 perovskite. Fuel 2020, 271, 117624. [Google Scholar] [CrossRef]
  85. Toemen, S.; Abu Bakar, W.A.W.; Ali, R. Effect of ceria and strontia over Ru/Mn/Al2O3 catalyst: Catalytic methanation, physicochemical and mechanistic studies. J. CO2 Util. 2016, 13, 38–49. [Google Scholar] [CrossRef]
  86. Steiger, P.; Kröcher, O.; Ferri, D. Increased nickel exsolution from LaFe0.8Ni0.2O3 perovskite-derived CO2 methanation catalysts through strontium doping. Appl. Catal. A Gen. 2020, 590, 117328. [Google Scholar] [CrossRef]
  87. Huynh, H.L.; Yu, Z. CO2 Methanation on Hydrotalcite-Derived Catalysts and Structured Reactors: A Review. Energy Technol. 2020, 8, 1901475. [Google Scholar] [CrossRef] [Green Version]
  88. Bacariza, M.C.; Graça, I.; Lopes, J.M.; Henriques, C. Tuning Zeolite Properties towards CO2 Methanation: An Overview. ChemCatChem 2019, 11, 2388–2400. [Google Scholar] [CrossRef]
  89. Bette, N.; Thielemann, J.; Schreiner, M.; Mertens, F. Methanation of CO2 over a (Mg,Al)Ox Supported Nickel Catalyst Derived from a (Ni,Mg,Al)-Hydrotalcite-like Precursor. ChemCatChem 2016, 8, 2903–2906. [Google Scholar] [CrossRef]
  90. Liu, J.; Bing, W.; Xue, X.; Wang, F.; Wang, B.; He, S.; Zhang, Y.; Wei, M. Alkaline-assisted Ni nanocatalysts with largely enhanced low-temperature activity toward CO2 methanation. Catal. Sci. Technol. 2016, 6, 3976–3983. [Google Scholar] [CrossRef]
  91. Ren, J.; Mebrahtu, C.; Palkovits, R. Ni-based catalysts supported on Mg-Al hydrotalcites with different morphologies for CO2 methanation: Exploring the effect of metal-support interaction. Catal. Sci. Technol. 2020, 10, 1902–1913. [Google Scholar] [CrossRef]
  92. Wierzbicki, D.; Debek, R.; Motak, M.; Grzybek, T.; Gálvez, M.E.; Da Costa, P. Novel Ni-La-hydrotalcite derived catalysts for CO2 methanation. Catal. Commun. 2016, 83, 5–8. [Google Scholar] [CrossRef] [Green Version]
  93. Wierzbicki, D.; Motak, M.; Grzybek, T.; Gálvez, M.E.; Da Costa, P. The influence of lanthanum incorporation method on the performance of nickel-containing hydrotalcite-derived catalysts in CO2 methanation reaction. Catal. Today 2018, 307, 205–211. [Google Scholar] [CrossRef]
  94. Martins, J.A.; Faria, A.C.; Soria, M.A.; Miguel, C.V.; Rodrigues, A.E.; Madeira, L.M. CO2 methanation over hydrotalcite-derived nickel/ruthenium and supported ruthenium catalysts. Catalysts 2019, 9, 1008. [Google Scholar] [CrossRef] [Green Version]
  95. Mebrahtu, C.; Krebs, F.; Perathoner, S.; Abate, S.; Centi, G.; Palkovits, R. Hydrotalcite based Ni-Fe/(Mg, Al)Ox catalysts for CO2 methanation-tailoring Fe content for improved CO dissociation, basicity, and particle size. Catal. Sci. Technol. 2018, 8, 1016–1027. [Google Scholar] [CrossRef]
  96. Graça, I.; González, L.V.; Bacariza, M.C.; Fernandes, A.; Henriques, C.; Lopes, J.M.; Ribeiro, M.F. CO2 hydrogenation into CH4 on NiHNaUSY zeolites. Appl. Catal. B Environ. 2014, 147, 101–110. [Google Scholar] [CrossRef]
  97. Westermann, A.; Azambre, B.; Bacariza, M.C.; Graça, I.; Ribeiro, M.F.; Lopes, J.M.; Henriques, C. Insight into CO2 methanation mechanism over NiUSY zeolites: An operando IR study. Appl. Catal. B Environ. 2015, 174, 120–125. [Google Scholar] [CrossRef]
  98. Bacariza, M.C.; Bértolo, R.; Graça, I.; Lopes, J.M.; Henriques, C. The effect of the compensating cation on the catalytic performances of Ni/USY zeolites towards CO2 methanation. J. CO2 Util. 2017, 21, 280–291. [Google Scholar] [CrossRef]
  99. Bacariza, M.C.; Graça, I.; Lopes, J.M.; Henriques, C. Enhanced activity of CO2 hydrogenation to CH4 over Ni based zeolites through the optimization of the Si/Al ratio. Microporous Mesoporous Mater. 2018, 267, 9–19. [Google Scholar] [CrossRef]
  100. Quindimil, A.; De-La-Torre, U.; Pereda-Ayo, B.; González-Marcos, J.A.; González-Velasco, J.R. Ni catalysts with La as promoter supported over Y- and BETA- zeolites for CO2 methanation. Appl. Catal. B Environ. 2018, 238, 393–403. [Google Scholar] [CrossRef]
  101. Bacariza, M.C.; Graça, I.; Lopes, J.M.; Henriques, C. Ni-Ce/Zeolites for CO2 Hydrogenation to CH4: Effect of the Metal Incorporation Order. ChemCatChem 2018, 10, 2773–2781. [Google Scholar] [CrossRef]
  102. Zhang, F.; Lu, B.; Sun, P. Highly stable Ni-based catalysts derived from LDHs supported on zeolite for CO2 methanation. Int. J. Hydrog. Energy 2020, 45, 16183–16192. [Google Scholar] [CrossRef]
  103. Xu, L.; Wang, F.; Chen, M.; Nie, D.; Lian, X.; Lu, Z.; Chen, H.; Zhang, K.; Ge, P. CO2 methanation over rare earth doped Ni based mesoporous catalysts with intensified low-temperature activity. Int. J. Hydrog. Energy 2017, 42, 15523–15539. [Google Scholar] [CrossRef]
  104. Zhao, K.; Li, Z.; Bian, L. CO2 methanation and co-methanation of CO and CO2 over Mn-promoted Ni/Al2O3 catalysts. Front. Chem. Sci. Eng. 2016, 10, 273–280. [Google Scholar] [CrossRef]
  105. Liu, Q.; Bian, B.; Fan, J.; Yang, J. Cobalt doped Ni based ordered mesoporous catalysts for CO2 methanation with enhanced catalytic performance. Int. J. Hydrog. Energy 2018, 43, 4893–4901. [Google Scholar] [CrossRef]
  106. Mihet, M.; Lazar, M.D. Methanation of CO2 on Ni/γ-Al2O3: Influence of Pt, Pd or Rh promotion. Catal. Today 2018, 306, 294–299. [Google Scholar] [CrossRef]
  107. Guilera, J.; Del Valle, J.; Alarcón, A.; Díaz, J.A.; Andreu, T. Metal-oxide promoted Ni/Al2O3 as CO2 methanation micro-size catalysts. J. CO2 Util. 2019, 30, 11–17. [Google Scholar] [CrossRef]
  108. Garbarino, G.; Wang, C.; Cavattoni, T.; Finocchio, E.; Riani, P.; Flytzani-Stephanopoulos, M.; Busca, G. A study of Ni/La-Al2O3 catalysts: A competitive system for CO2 methanation. Appl. Catal. B Environ. 2019, 248, 286–297. [Google Scholar] [CrossRef]
  109. Liang, C.; Wei, T.; Wang, H.; Yu, Z.; Dong, D.; Zhang, S.; Liu, Q.; Hu, G.; Hu, X. Impacts of La addition on formation of the reaction intermediates over alumina and silica supported nickel catalysts in methanation of CO2. J. Energy Inst. 2020, 93, 723–738. [Google Scholar] [CrossRef]
  110. Wang, X.; Zhu, L.; Zhuo, Y.; Zhu, Y.; Wang, S. Enhancement of CO2 Methanation over La-Modified Ni/SBA-15 Catalysts Prepared by Different Doping Methods. ACS Sustain. Chem. Eng. 2019, 7, 14647–14660. [Google Scholar] [CrossRef]
  111. Zhang, T.; Liu, Q. Perovskite LaNiO3 Nanocrystals inside Mesostructured Cellular Foam Silica: High Catalytic Activity and Stability for CO2 Methanation. Energy Technol. 2020, 8, 1–12. [Google Scholar] [CrossRef]
  112. Bermejo-López, A.; Pereda-Ayo, B.; González-Marcos, J.A.; González-Velasco, J.R. Ni loading effects on dual function materials for capture and in-situ conversion of CO2 to CH4 using CaO or Na2CO3. J. CO2 Util. 2019, 34, 576–587. [Google Scholar] [CrossRef]
  113. Duyar, M.S.; Wang, S.; Arellano-Treviño, M.A.; Farrauto, R.J. CO2 utilization with a novel dual function material (DFM) for capture and catalytic conversion to synthetic natural gas: An update. J. CO2 Util. 2016, 15, 65–71. [Google Scholar] [CrossRef]
  114. Zheng, Q.; Farrauto, R.; Chau Nguyen, A. Adsorption and Methanation of Flue Gas CO2 with Dual Functional Catalytic Materials: A Parametric Study. Ind. Eng. Chem. Res. 2016, 55, 6768–6776. [Google Scholar] [CrossRef]
  115. Wang, S.; Farrauto, R.J.; Karp, S.; Jeon, J.H.; Schrunk, E.T. Parametric, cyclic aging and characterization studies for CO2 capture from flue gas and catalytic conversion to synthetic natural gas using a dual functional material (DFM). J. CO2 Util. 2018, 27, 390–397. [Google Scholar] [CrossRef]
  116. Proaño, L.; Tello, E.; Arellano-Trevino, M.A.; Wang, S.; Farrauto, R.J.; Cobo, M. In-situ DRIFTS study of two-step CO2 capture and catalytic methanation over Ru,“Na2O”/Al2O3 Dual Functional Material. Appl. Surf. Sci. 2019, 479, 25–30. [Google Scholar] [CrossRef]
  117. Wang, F.; He, S.; Chen, H.; Wang, B.; Zheng, L.; Wei, M.; Evans, D.G.; Duan, X. Active Site Dependent Reaction Mechanism over Ru/CeO2 Catalyst toward CO2 Methanation. J. Am. Chem. Soc. 2016, 138, 6298–6305. [Google Scholar] [CrossRef]
  118. Zhen, W.; Li, B.; Lu, G.; Ma, J. Enhancing catalytic activity and stability for CO2 methanation on Ni–Ru/γ-Al2O3 via modulating impregnation sequence and controlling surface active species. RSC Adv. 2014, 4, 16472–16479. [Google Scholar] [CrossRef]
  119. Chai, K.H.; Leong, L.K.; Wong, D.S.H.; Tsai, D.H.; Sethupathi, S. Effect of CO2 adsorbents on the Ni-based dual-function materials for CO2 capturing and in situ methanation. J. Chin. Chem. Soc. 2020, 67, 1–11. [Google Scholar] [CrossRef]
  120. Hu, L.; Urakawa, A. Continuous CO2 capture and reduction in one process: CO2 methanation over unpromoted and promoted Ni/ZrO2. J. CO2 Util. 2018, 25, 323–329. [Google Scholar] [CrossRef] [Green Version]
  121. Zhou, Z.; Sun, N.; Wang, B.; Han, Z.; Cao, S.; Hu, D.; Zhu, T.; Shen, Q.; Wei, W. 2D-Layered Ni–MgO–Al2O3 Nanosheets for Integrated Capture and Methanation of CO2. ChemSusChem 2020, 13, 360–368. [Google Scholar] [CrossRef]
  122. Bobadilla, L.F.; Riesco-García, J.M.; Penelás-Pérez, G.; Urakawa, A. Enabling continuous capture and catalytic conversion of flue gas CO2 to syngas in one process. J. CO2 Util. 2016, 14, 106–111. [Google Scholar] [CrossRef] [Green Version]
  123. Hyakutake, T.; Van Beek, W.; Urakawa, A. Unravelling the nature, evolution and spatial gradients of active species and active sites in the catalyst bed of unpromoted and K/Ba-promoted Cu/Al2O3 during CO2 capture-reduction. J. Mater. Chem. A 2016, 4, 6878–6885. [Google Scholar] [CrossRef]
  124. Al-Mamoori, A.; Rownaghi, A.A.; Rezaei, F. Combined Capture and Utilization of CO2 for Syngas Production over Dual-Function Materials. ACS Sustain. Chem. Eng. 2018, 6, 13551–13561. [Google Scholar] [CrossRef]
  125. Sun, H.; Wang, J.; Zhao, J.; Shen, B.; Shi, J.; Huang, J.; Wu, C. Dual functional catalytic materials of Ni over Ce-modified CaO sorbents for integrated CO2 capture and conversion. Appl. Catal. B Environ. 2019, 244, 63–75. [Google Scholar] [CrossRef] [Green Version]
  126. Veselovskaya, J.V.; Parunin, P.D.; Netskina, O.V.; Kibis, L.S.; Lysikov, A.I.; Okunev, A.G. Catalytic methanation of carbon dioxide captured from ambient air. Energy 2018, 159, 766–773. [Google Scholar] [CrossRef]
  127. Veselovskaya, J.V.; Parunin, P.D.; Netskina, O.V.; Okunev, A.G. A Novel Process for Renewable Methane Production: Combining Direct Air Capture by K2CO3/Alumina Sorbent with CO2 Methanation over Ru/Alumina Catalyst. Top. Catal. 2018, 61, 1528–1536. [Google Scholar] [CrossRef]
  128. Veselovskaya, J.V.; Lysikov, A.I.; Netskina, O.V.; Kuleshov, D.V.; Okunev, A.G. K2CO3-Containing Composite Sorbents Based on Thermally Modified Alumina: Synthesis, Properties, and Potential Application in a Direct Air Capture/Methanation Process. Ind. Eng. Chem. Res. 2020, 59, 7130–7139. [Google Scholar] [CrossRef]
  129. Miguel, C.V.; Soria, M.A.; Mendes, A.; Madeira, L.M. A sorptive reactor for CO2 capture and conversion to renewable methane. Chem. Eng. J. 2017, 322, 590–602. [Google Scholar] [CrossRef]
Figure 1. (a) Conversions of CO2 and (b) selectivity toward CH4 and CO as a function of reaction Table 0. Ru/X-TiO2 catalysts of the same alkali loading (0.2 wt.%). Experimental conditions: mass of catalyst: 100 mg; particle diameter: 0.18 < dp < 0.25 mm; feed composition: 5% CO2, 20% H2 (balance He); total flow rate: 150 cm3 min−1. Reproduced with permission from ref. [33]. Copyright 2018 Elsevier.
Figure 1. (a) Conversions of CO2 and (b) selectivity toward CH4 and CO as a function of reaction Table 0. Ru/X-TiO2 catalysts of the same alkali loading (0.2 wt.%). Experimental conditions: mass of catalyst: 100 mg; particle diameter: 0.18 < dp < 0.25 mm; feed composition: 5% CO2, 20% H2 (balance He); total flow rate: 150 cm3 min−1. Reproduced with permission from ref. [33]. Copyright 2018 Elsevier.
Catalysts 10 00812 g001
Figure 2. In-situ DRIFTS characterization for the CO2 methanation over Ni/Al2O3 and Ni-KOH/Al2O3 catalysts. (a): Ni/Al2O3, (b): Ni-KOH/Al2O3. Reaction conditions: CO2/H2 = 1/4; (CO2 = 10 mL/min; H2 = 40 mL/min; N2 = 50 mL/min). Reproduced with permission from ref. [45]. Copyright 2018 Elsevier.
Figure 2. In-situ DRIFTS characterization for the CO2 methanation over Ni/Al2O3 and Ni-KOH/Al2O3 catalysts. (a): Ni/Al2O3, (b): Ni-KOH/Al2O3. Reaction conditions: CO2/H2 = 1/4; (CO2 = 10 mL/min; H2 = 40 mL/min; N2 = 50 mL/min). Reproduced with permission from ref. [45]. Copyright 2018 Elsevier.
Catalysts 10 00812 g002
Figure 3. (a) CO2 hydrogenation product selectivity for the catalysts derived from LaNiO3 and K-substituted (La1-xKxNiO3) perovskites; (b) rising CO selectivity towards the RWGS at T50% for the catalysts containing increased amounts of K. Reproduced with permission from ref. [51]. Copyright 2020 MDPI.
Figure 3. (a) CO2 hydrogenation product selectivity for the catalysts derived from LaNiO3 and K-substituted (La1-xKxNiO3) perovskites; (b) rising CO selectivity towards the RWGS at T50% for the catalysts containing increased amounts of K. Reproduced with permission from ref. [51]. Copyright 2020 MDPI.
Catalysts 10 00812 g003
Figure 4. CO2 conversion (a) and CH4 selectivity (b) of the Ni/M0.1Ce0.9Ox catalysts; WHSV = 36,000 mL h−1 g−1. (c) CO2 desorption profiles of the Ni/M0.1Ce0.9Ox catalysts and (d) Linear relationship between the TOF values at 250 °C and the amount of surface moderate alkaline sites on the Ni/M0.1Ce0.9Ox catalysts. a. Ni/CeO2; b. Ni/Mg0.1Ce0.9Ox; c. Ni/Ca0.1Ce0.9Ox; d. Ni/Sr0.1Ce0.9Ox; e. Ni/Ba0.1Ce0.9Ox; f. Ni/Ca0.1Ce0.9Ox-IMP. Reproduced with permission from ref. [30]. Copyright 2020 Elsevier.
Figure 4. CO2 conversion (a) and CH4 selectivity (b) of the Ni/M0.1Ce0.9Ox catalysts; WHSV = 36,000 mL h−1 g−1. (c) CO2 desorption profiles of the Ni/M0.1Ce0.9Ox catalysts and (d) Linear relationship between the TOF values at 250 °C and the amount of surface moderate alkaline sites on the Ni/M0.1Ce0.9Ox catalysts. a. Ni/CeO2; b. Ni/Mg0.1Ce0.9Ox; c. Ni/Ca0.1Ce0.9Ox; d. Ni/Sr0.1Ce0.9Ox; e. Ni/Ba0.1Ce0.9Ox; f. Ni/Ca0.1Ce0.9Ox-IMP. Reproduced with permission from ref. [30]. Copyright 2020 Elsevier.
Catalysts 10 00812 g004
Figure 5. Catalytic properties of the prepared catalysts: (a) CO2 conversion and (b) CH4 selectivity; (c) lifetime test of the 10N1R2C-OMA catalyst. Reproduced with permission from ref. [75]. Copyright 2018 Elsevier.
Figure 5. Catalytic properties of the prepared catalysts: (a) CO2 conversion and (b) CH4 selectivity; (c) lifetime test of the 10N1R2C-OMA catalyst. Reproduced with permission from ref. [75]. Copyright 2018 Elsevier.
Catalysts 10 00812 g005aCatalysts 10 00812 g005b
Figure 6. Plausible mechanism for CO2 methanation over the Ca-NiTiO3/γ-Al2O3 catalysts. Reproduced with permission from ref. [84]. Copyright 2020 Elsevier.
Figure 6. Plausible mechanism for CO2 methanation over the Ca-NiTiO3/γ-Al2O3 catalysts. Reproduced with permission from ref. [84]. Copyright 2020 Elsevier.
Catalysts 10 00812 g006
Figure 7. (a) CO2 methanation at steady state as a function of reaction temperature over the as-prepared Ni/MgAl-MMO, Ni/CNT, Ni/Al2O3 and Ni/MgO catalysts (GHSV: 2400 h−1). (b) The corresponding selectivity toward CH4 as a function of reaction temperature. (c) Relationship between the TOFNi values and the concentration of alkaline sites in the Nix/Mg2−xAl-MMO catalyst and (d) CO2 methanation over the hydrotalcite-derived catalysts that consist of Ni nanoparticles finely dispersed over a MgO-Al2O3 mixed oxide support (Ni/MgAl-MMO). Reproduced with permission from ref. [90]. Copyright 2016 Royal Society of Chemistry.
Figure 7. (a) CO2 methanation at steady state as a function of reaction temperature over the as-prepared Ni/MgAl-MMO, Ni/CNT, Ni/Al2O3 and Ni/MgO catalysts (GHSV: 2400 h−1). (b) The corresponding selectivity toward CH4 as a function of reaction temperature. (c) Relationship between the TOFNi values and the concentration of alkaline sites in the Nix/Mg2−xAl-MMO catalyst and (d) CO2 methanation over the hydrotalcite-derived catalysts that consist of Ni nanoparticles finely dispersed over a MgO-Al2O3 mixed oxide support (Ni/MgAl-MMO). Reproduced with permission from ref. [90]. Copyright 2016 Royal Society of Chemistry.
Catalysts 10 00812 g007aCatalysts 10 00812 g007b
Figure 8. Schematic summarizing the dual-function materials (DFM) concept (adopted from [113]).
Figure 8. Schematic summarizing the dual-function materials (DFM) concept (adopted from [113]).
Catalysts 10 00812 g008
Figure 9. Proposed CO2 methanation mechanism over 5%Ru/Al2O3 (down) and extended (up and down) for over 5%Ru, 6.1%“Na2O”/Al2O3 DFM. Reproduced with permission from ref. [116]. Copyright 2019 Elsevier.
Figure 9. Proposed CO2 methanation mechanism over 5%Ru/Al2O3 (down) and extended (up and down) for over 5%Ru, 6.1%“Na2O”/Al2O3 DFM. Reproduced with permission from ref. [116]. Copyright 2019 Elsevier.
Catalysts 10 00812 g009
Figure 10. CO2, H2O, CH4 and CO concentration profiles during one CO2 adsorption and hydrogenation cycle at 370 °C with (a) Ru10CaO and (b) Ru10Na2CO3 samples. Reproduced with permission from ref. [24]. Copyright 2019 Elsevier.
Figure 10. CO2, H2O, CH4 and CO concentration profiles during one CO2 adsorption and hydrogenation cycle at 370 °C with (a) Ru10CaO and (b) Ru10Na2CO3 samples. Reproduced with permission from ref. [24]. Copyright 2019 Elsevier.
Catalysts 10 00812 g010
Figure 11. Schematic illustration for the CO2 capture from the MgO sorbent phase and the conversion of captured CO2 into CH4 over the Ru/CeO2 catalyst phase. Reproduced with permission from ref. [25]. Copyright 2020 Elsevier.
Figure 11. Schematic illustration for the CO2 capture from the MgO sorbent phase and the conversion of captured CO2 into CH4 over the Ru/CeO2 catalyst phase. Reproduced with permission from ref. [25]. Copyright 2020 Elsevier.
Catalysts 10 00812 g011
Figure 12. CO2, H2O, CH4 and CO concentration profiles during one CO2 adsorption and hydrogenation cycle at 400 °C with (a) 10NiCa and (b) 10NiNa catalysts. Reproduced with permission from ref. [112]. Copyright 2019 Elsevier.
Figure 12. CO2, H2O, CH4 and CO concentration profiles during one CO2 adsorption and hydrogenation cycle at 400 °C with (a) 10NiCa and (b) 10NiNa catalysts. Reproduced with permission from ref. [112]. Copyright 2019 Elsevier.
Catalysts 10 00812 g012
Figure 13. (a) TEM images and (b) EDX Mapping of 2.0NiMgAl-LDO. Continuous CO2 capture and methanation over 2.0NiMgAl-LDO-re at 250 °C (Adsorption: 15 vol % CO2/N2 (70 mL min−1), 1 bar; methanation: 100 % H2 (70 mL min−1), 1 bar); (c) tadsorption = 30 s, tmethanation = 30 s; and (d) tadsorption = 35 s, tmethanation = 25 s. Reproduced with permission from ref. [121]. Copyright 2019 Wiley.
Figure 13. (a) TEM images and (b) EDX Mapping of 2.0NiMgAl-LDO. Continuous CO2 capture and methanation over 2.0NiMgAl-LDO-re at 250 °C (Adsorption: 15 vol % CO2/N2 (70 mL min−1), 1 bar; methanation: 100 % H2 (70 mL min−1), 1 bar); (c) tadsorption = 30 s, tmethanation = 30 s; and (d) tadsorption = 35 s, tmethanation = 25 s. Reproduced with permission from ref. [121]. Copyright 2019 Wiley.
Catalysts 10 00812 g013
Table 1. Summary of some typical alkali and alkaline earth promoted catalysts on the methanation of CO2.
Table 1. Summary of some typical alkali and alkaline earth promoted catalysts on the methanation of CO2.
PromoterCatalystSynthesis MethodConditionsPerformanceCommentsRef.
0.06–0.4% Li, Na, K and Cs0.5% and 5% Ru/TiO2Wet impregnationGHSV = 56,000 h−1
H2/CO2 = 4/1
XCO2 = 60%,
SCH4 = 95%
at 400 °C
(0.2% Na on 0.5% Ru/TiO2)
Catalytic activity improved after addition of small amounts of alkalimetals on 0.5% Ru/TiO2. Less pronounced effect for 5% Ru/TiO2[33]
10% LiNO3, NaNO3, Na2CO3 and K2CO31% Ru/Al2O3Incipient wetness impregnationGHSV = 25,000 h−1
H2/CO2 = 4/1
XCO2 = 60%,
SCH4 = 99%
at 350 °C
(10% LiNO3 on 1% Ru/Al2O3)
Only LiNO3 significantly promoted activity. Na2CO3 and K2CO3 masked Ru particles and suppressed methanation activity[35]
5% Na, K, Mg and Ca20% Co/Al2O3Incipient wetness impregnationH2/CO2 = 4/1XCO2 = 52%,
SCH4 = 67%
at 400 °C
(5% Na on 10% Co/Al2O3)
Na and K increased the population of basic sites but promoted CO selectivity. Mg and Ca did not favour methanation activity either, but prevented Co nanoparticle sintering[37]
0.1–1% Na33% Ni/SiO2 and 13% Ni/CeO2Wet impregnationH2/CO2 = 50/1XCO2 = 85%,
SCH4 = 85%
at 250 °C
(0.1% Na on 33% Ni/SiO2)
Ni active surface area decreased upon Na addition. Negative impact on activity for Ni/CeO2, but slightly positive impact for Ni/SiO2[41]
16.5% K and Ba1% Rh/Al2O3Flame spray pyrolysisGHSV = 6000 h−1
H2/CO2 = 4/1
XCO2 = 25%,
SCH4 = 0%
at 375 °C
(16.5 % K on 1% Rh/Al2O3)
K-addition led to pure CO production over the entire temperature range. Ba-modified catalyst slightly inferior to Rh/Al2O3 for CO2 methanation[43]
0.5–20% KOH0.5–20% Ni/Al2O3Incipient wetness impregnationH2/CO2 = 4/1XCO2 = 40%,
SCH4 = 45%
at 400 °C
(10% K on 12% Ni/Al2O3)
Strongly adsorbed formate intermediates over KOH-modified catalysts favoured CO selectivity[45]
1.7–7.6% K
(x = 0.1–0.4 for La1-xKxNiO3)
24–28.5% Ni/La2O3Citrate sol-gel La1-xKxNiO3 synthesis and reductionGHSV = 48,000 h−1
H2/CO2 = 12.5/1
XCO2 = 50%,
SCH4 = 45%
at 306 °C
(for x= 0.3 => La0.7K0.3NiO3)
Easy-to-desorb CO species formed after K-doping stirred the reaction towards RWGS and CO selectivity[51]
4% MgO, CaO, SrO and BaO10% Ni/SiO2Wet impregnationGHSV = 15,000 h−1
H2/CO2 = 4/1
XCO2 = 71%,
SCH4 = 99%
at 300 °C
(4% SrO on 10% Ni/SiO2)
MgO inhibited activity, CaO had little effect, SrO and BaO enhanced activity, while SrO also promoted the catalyst stability[52]
1–7.5% Mg, Ca, Sr and Ba20% Ni/Al2O3Incipient wetness impregnationGHSV = 16,000 h−1
H2/CO2 = 4/1
XCO2 = 80%,
SCH4 = 100%
at 300 °C
(5% Ba on 20% Ni/Al2O3)
Lighter elements Mg and Ca promoted CO selectivity, while heavier elements Sr and Ba favoured methanation[53]
1.4% Mg, 2.2% Ca, 4.7% Sr and 7.2% Ba (Ce0.9M0.1Ox support)10% Ni/Ce0.9M0.1OxCitrate sol-gel modified CeO2 synthesis and Ni incipient wetness impregnationWHSV = 36,000 mL h−1 g−1
H2/CO2 = 4/1
XCO2 = 70%,
SCH4 = 95%
at 270 °C
(10% Ni/Ce0.9Ca0.1Ox)
Ca-doping led to the highest methanation activity (with Sr being second-to-best). TOFNi could be linearly correlated to the amount of oxygen vacancies and moderately strong basic sites on the support surface[30]
1–4% MgO10% Ni/SiO2Sequential wet impregnations and co-impregnationWHSV = 36,000 mL h−1 g−1
H2/CO2 = 4/1
XCO2 = 65%,
SCH4 = 98%
at 400 °C
(1% Mg on 10% Ni/SiO2)
1% MgO introduced by co-impregantion led to the highest catalytic activity and stability. Higher MgO contents blocked some Ni active sites[63]
0.4–2.5% Mg (Change in Ni/Mg molar ratio)6% Ni/ZrO2Citrate complexing methodWHSV = 15,000 mL h−1 g−1,
p = 0,1 MPa,
H2/CO2 = 4/1
XCO2 = 90%,
SCH4 = 100%
at 250 °C
(1% Mg on 6% Ni/ZrO2)
Mg-addition led to Ni nanoparticle confinement, as well as resistance to sintering and coke formation[64]
0.5–5% Mg12% Ni/Al2O3 (OMA)Evaporation-induced self-assembly (EISA)GHSV = 15,000 h−1
H2/CO2 = 4/1
XCO2 = 65%,
SCH4 = 96%
at 300 °C
(2.5% Mg on 12% Ni/Al2O3)
Mg-dopant increased the surface basicity of the catalysts and facilitated the chemisorption and activation of CO2[65]
1% Mgα-Fe2O3Incipient wetness impregnationGHSV = 150,000 h−1,
p = 8 bar,
H2/CO2 = 4/1
XCO2 = 22%,
SCH4 = 9%
at 400 °C
after 18 h
(2.% Mg on α-Fe2O3)
Improved CO2 adsorption over the Mg-modified Fe catalyst led to highly active Fe carbides and better methanation performance[72]
0.8–8% Ca12% Ni/Al2O3 (OMA)Evaporation-induced self-assembly (EISA)GHSV = 15,000 h−1
H2/CO2 = 4/1
XCO2 = 70%,
SCH4 = 98%
at 300 °C
(6.4% Ca on 12% Ni/Al2O3)
Increased surface alkalinity of the support led to enhanced CO2 chemisorption and lower activation energy for CO2 methanation[74]
2% CaO1% Ru, 10% Ni/Al2O3 (OMA)Evaporation-induced self-assembly (EISA)WHSV = 30,000 mL h−1 g−1
H2/CO2 = 4/1
XCO2 = 80%,
SCH4 = 98%
at 360 °C
(2% CaO on 0.8% Ru, 8% Ni/Al2O3)
Step increase in the CO2 conversion and CH4 selectivity after adding Ru second metal and CaO promoter[75]
0.7% Ca5% Ni/CaxZr1-xO2-δWet impregnationH2/CO2 = 4/1XCO2 = 72%,
SCH4 = 100%
at 350 °C
(0.7% Ca on 5% Ni/CaxZr1-xO2-δ)
Oxygen vacancies-coordinately unsaturated sites (cus) pairs facilitated the C-O bond scission, a crucial step during CO2 methanation[82]
0.1–1% Ca
(x = 0.01–0.1 for CaxNi1-xTiO3-δ)
30% Ca-NiTiO3/Al2O3 (11.5% Ni)Coprecipitation Ca-NiTiO3 synthesis and wet impregnationGHSV = 5000 h−1
H2/CO2 = 4/1
XCO2 = 55%,
SCH4 = 99.7%
at 350 °C
(0.5% Ca on 30% Ca-NiTiO3/Al2O3)
Ca-doping on NiTiO3 A-site promoted the formation of oxygen vacancies in close proximity to active Ni and therefore the chemisorption of CO2[84]
14% Mg59% Ni/MgO-
Al2O3
Coprecipitation hydrotalcite synthesis and reductionWHSV = 1100 min-1
g−1
H2/CO2 = 4/1
XCO2 = 70%,
SCH4 = 90%
at 315 °C(14% Mg on 59% Ni/MgO-
Al2O3)
Stable performance, as well as sintering and coking resistance for the hydrotalcite-derived catalysts[89]
6.1–28.5% Mg17.2–57.8% Ni/MgO-
Al2O3
Coprecipitation hydrotalcite synthesis and reductionGHSV = 2400 h−1
H2/CO2 = 4/1
XCO2 = 97.9%, SCH4 = 97.5%
at 250 °C
(17.3% Mg on 37.5% Ni/MgO-Al2O3
Far superior low-temperature CO2 conversion for Mg-containing hydrotalcite-derived catalysts, due to the increased surface alkalinity[90]
0.3–12% Li, Na, K, Cs, Mg, Ca and Ba15% Ni/USY zeoliteIon-exchange for cations and Ni incipient wetness impregnationGHSV = 43,000 h−1
H2/CO2 = 4/1
XCO2 = 62%,
SCH4 = 92%
at 375 °C
(12.1% Cs in 15% Ni/Cs-exchanged USY)
Greater size for monovalent cations improved CO2 adsorption
(Cs+ > Na+ > Li+ > H+), while Mg2+ benefited Ni dispersion
[98]
0.9–13% Mg4.8–15% Ni/USY zeoliteIon-exchange or impregnation for Mg and Ni incipient wetness impregnationGHSV = 43,000 h−1
H2/CO2 = 4/1
XCO2 = 65%,
SCH4 = 93%
at 400 °C
(0.7% Mg in 15% Ni/Mg-exchanged USY)
Finer and better-dispersed Ni nanoparticles after Mg impregnation or Mg ion-exchange[68]
Table 2. Summary of dual-function materials (DFMs) employed in the combined CO2 capture and methanation process.
Table 2. Summary of dual-function materials (DFMs) employed in the combined CO2 capture and methanation process.
CO2 SorbentCatalystCO2 Capture ConditionsCO2 Methanation ConditionsCH4 Production [mmol CH4/g DFM]CommentsReference
1–10% CaO1–10% Ru/Al2O310% CO2/N2, 320 °C, 30 min4% H2/N2, 320 °C, 2 h0.50
(5% Ru, 10% CaO/Al2O3)
High sorbent loading increased CO2 spillover to Ru active sites. 5% Ru, 10%CaO/Al2O3 as best composition. Cyclic operation at 320 °C[19]
10% CaO, MgO, Na2CO3 and K2CO30.1–10% Rh and 5% Ru/Al2O310% CO2/N2, 320 °C, 30 min4% H2/N2, 320 °C, 2 h0.4
(0.1% Rh, 10% CaO/Al2O3),
1.05
(5% Ru, 10% Na2CO3/Al2O3)
0.1% Rh was quite active but Rh is pricy.
Na2CO3 constituted a better adsorbent than CaO for Ru-based DFMs
[113]
10% CaO, MgO, Na2CO3 (6.1% “Na2O”) and K2CO3 (7.01% “Κ2O”)0.5% Rh and 5% Ru/Al2O310% CO2/N2, 320 °C, 30 min10% H2/N2, 320 °C, 1 h0.614
(5% Ru, 6.1% “Na2O”/Al2O3)
Ru-based DFMs with Na2CO3 achieved the fastest hydrogenation rates[21]
5–15% CaO and Na2CO34% Ru/Al2O311% CO2/Ar, 280–400 °C,
1 min
10% H2/Ar, 280–400 °C,
2 min
0.414 at 400 °C
(4% Ru, 15% CaO/Al2O3),
0.383 at 310 °C
(4% Ru, 10% Na2CO3/Al2O3)
Higher sorbent loading promoted CH4 yield. More stable carbonates formed on CaO required higher temperature to decompose[24]
10% LiNO3, NaNO3, Na2CO3 and K2CO31% Ru/Al2O315% CO2/Ar, 230 °C, 30 min12% H2/Ar, 230 °C, 30 minNot reportedLi-promoted DFM presented faster methanation kinetics. Operation at just 230 °C, a lower temperature compared with other DFMs[35]
200% High-capacity MgO sorbent2.5–10% Ru/CeO265% CO2/N2, 300 °C, 1 h5% H2/N2, 300 °C,6.60, 1. cycle
3.36, 10. cycle
(5% Ru/CeO2 + 200% MgO)
High CO2 capture capacity due to a large MgO content. Mechanism: oxygen vacancy catalyzed CO2 hydrogenation through a formate intermediate[25]
10% Na2CO3
(6.1% “Na2O”)
10% Ni/Al2O37.5% CO2/N2 and 7.5% CO2, 4.5% O2, 15% H2O/N2, 320 °C, 20 min15% H2/N2, 320 °C, 1 h0.276, Without O2 during CO2 capture
0, With O2 during CO2 capture
(10% Ni, 6.1% “Na2O”/Al2O3)
Ni-based DFMs required a prereduction/activation step at 650 °C and an O2-free gas during CO2 capture in order to produce CH4[21]
10% Na2CO3
(6.1% “Na2O”)
0.1–1% Ru, Pt, Pd with 10% Ni/Al2O37.5% CO2, 4.5% O2, 15% H2O/N2, 320 °C, 20 min15% H2/N2, 320 °C0.38
(1% Ru, 10% Ni, 6.1% “Na2O”/Al2O3)
Pt and Ru addition as second metals promoted NiO reducibility.
Ni and Ru synergy led to the highest CH4 yield
[23]
5–15% CaO and Na2CO35–15% Ni/Al2O310% CO2/Ar, 280–520 °C,
1 min
10% H2/Ar, 280–520 °C,
2 min
0.142 at 520 °C
(15% Ni, 15% CaO/Al2O3),
0.185 at 400 °C
(10% Ni, 10% Na2CO3/Al2O3)
The adsorbent phase enhanced NiO reducibility. Methanation was favoured at higher temperatures when CaO was used as sorbent[112]
10% CaO, MgO, Na2CO3 and K2CO310% Ni/Al2O39.5% CO2/N2, 320 °C, 25 min10% H2/N2, 320 °C, 20 min0.140
(10% Ni, 10% CaO/Al2O3)
0.178
(10% Ni, 10% Na2CO3/Al2O3)
Easier CO2 release and higher methanation capacity when Na2CO3 was used as sorbent[119]
5% K and La15% Ni/ZrO24.7% CO2/He, 250–450 °C,
5 min
Pure H2, 250–450 °C,
5 min
Not reportedK-promotion greatly enhanced CO2 adsorption and promoted methanation at higher temperatures, while La-promotion favoured faster methanation kinetics[120]
15–30% Mg17–50% Ni/MgO-
Al2O3
15% CO2/N2, 200 and 250 °C,
various times
Pure H2, 200 and 250 °C, various timesNot reportedHydrotalcite-derived layered oxides with high Ni and Mg contents provided a high CH4 yield and favourable methanation kinetics[121]

Share and Cite

MDPI and ACS Style

Tsiotsias, A.I.; Charisiou, N.D.; Yentekakis, I.V.; Goula, M.A. The Role of Alkali and Alkaline Earth Metals in the CO2 Methanation Reaction and the Combined Capture and Methanation of CO2. Catalysts 2020, 10, 812. https://doi.org/10.3390/catal10070812

AMA Style

Tsiotsias AI, Charisiou ND, Yentekakis IV, Goula MA. The Role of Alkali and Alkaline Earth Metals in the CO2 Methanation Reaction and the Combined Capture and Methanation of CO2. Catalysts. 2020; 10(7):812. https://doi.org/10.3390/catal10070812

Chicago/Turabian Style

Tsiotsias, Anastasios I., Nikolaos D. Charisiou, Ioannis V. Yentekakis, and Maria A. Goula. 2020. "The Role of Alkali and Alkaline Earth Metals in the CO2 Methanation Reaction and the Combined Capture and Methanation of CO2" Catalysts 10, no. 7: 812. https://doi.org/10.3390/catal10070812

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