Next Article in Journal
Controlled Preparation of Different Proportions of Metal Fe-Mn from Waste Mn Ferrite by Molten Salt Electrolysis
Previous Article in Journal
Carbon Source Competition in Biological Selenate Reduction under Other Oxyanions Contamination
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Promising Catalytic Systems for CO2 Hydrogenation into CH4: A Review of Recent Studies

1
c5Lab—Sustainable Construction Materials Association, Edifício Central Park, Rua Central Park 6, 2795-242 Linda-a-Velha, Portugal
2
Centro de Química Estrutural, Instituto Superior Técnico, Universidade de Lisboa, Av Rovisco Pais, 1049-001 Lisboa, Portugal
3
Sorbonne Université, Campus Pierre et Marie Curie, Laboratoire de Réactivité de Surface, UMR CNRS 7197, 4 Place Jussieu, F-75005 Paris, France
4
Department of Chemical Engineering, Faculty of Engineering, University of Balamand, P.O. Box 33 Amioun El Koura, Lebanon
5
Centro de Química Estrutural and Departamento de Engenharia Química, Instituto Superior Técnico, Universidade de Lisboa, Av Rovisco Pais, 1049-001 Lisboa, Portugal
*
Author to whom correspondence should be addressed.
Processes 2020, 8(12), 1646; https://doi.org/10.3390/pr8121646
Submission received: 11 November 2020 / Revised: 28 November 2020 / Accepted: 9 December 2020 / Published: 13 December 2020
(This article belongs to the Section Chemical Processes and Systems)

Abstract

:
The increasing utilization of renewable sources for electricity production turns CO2 methanation into a key process in the future energy context, as this reaction allows storing the temporary renewable electricity surplus in the natural gas network (Power-to-Gas). This kind of chemical reaction requires the use of a catalyst and thus it has gained the attention of many researchers thriving to achieve active, selective and stable materials in a remarkable number of studies. The existing papers published in literature in the past few years about CO2 methanation tackled the catalysts composition and their related performances and mechanisms, which served as a basis for researchers to further extend their in-depth investigations in the reported systems. In summary, the focus was mainly in the enhancement of the synthesized materials that involved the active metal phase (i.e., boosting its dispersion), the different types of solid supports, and the frequent addition of a second metal oxide (usually behaving as a promoter). The current manuscript aims in recapping a huge number of trials and is divided based on the support nature: SiO2, Al2O3, CeO2, ZrO2, MgO, hydrotalcites, carbons and zeolites, and proposes the main properties to be kept for obtaining highly efficient carbon dioxide methanation catalysts.

1. Introduction

The promotion of electricity production using renewable sources is in the origin of a general interest for electric energy storage systems able to deal with their well-known intermittency [1,2,3]. Among all, the use of the temporary overproduction of renewable electricity for the synthesis of fuels (energy vectors) has attracted attention [1,3,4,5,6]. CO2 methanation is foreseen to play a key role in the future energy context [4,5], especially because CH4 is the main component of natural gas, which allows its injection in the existing and well-established network (Power-to-Gas) [7].
In addition, important efforts must be carried out in order to constrain CO2 emissions in the next years [8,9,10], being the development of suitable technologies for its capture and conversion to value-added products mandatory for reducing the contribution of sectors such as cement industries, responsible for 7–8% of global CO2 emissions [11,12]. Even if CO2 could be transformed into chemicals and fuels, its usage as feedstock is limited to a few industrial processes (synthesis of urea and derivatives, salicylic acid and carbonates) [13,14,15] due to its thermodynamic stability and, as a result, the high energy substances or electro-reductive processes typically needed for its transformation into other chemicals [16]. CO2 conversion into CH4 (Figure 1) represents a suitable alternative if implemented in cement industries, as they could use natural gas for combustion processes, reducing energy demands and CO2 emissions simultaneously.
Besides green hydrogen production and CO2 capture, CO2 methanation efficiency depends on the used catalysts. Therefore, finding active, selective and stable materials is considered as the core of the process [17]. Taking into account the significant number of publications dealing with this topic, the present work elaborates a review and compares recent studies regarding heterogeneous supported catalysts for conventional thermal CO2 hydrogenation to CH4. The results will be presented and organized based on the support’s nature. This work aims to complement excellent reviews published in the area of carbon dioxide methanation in the last few years [17,18,19,20,21,22,23,24,25,26,27,28,29]. The main findings reported will be summarized and the best performances (T: reaction temperature, XCO2: CO2 conversion, SCH4: selectivity to CH4) will be presented along with the main preparation (active metal(s) loading—wt%, given in italic in all the tables—and incorporation method) and operation (Tred: reduction temperature, H2/CO2 ratio, QT/W: flowrate per mass of catalyst) conditions applied. Additionally, and taking into account that CO2 methanation mechanism is strongly influenced by the catalyst’s composition, the main findings regarding mechanistic approaches will be also included.
Despite the fact that only a few studies have dealt with the use of unsupported catalysts for CO2 methanation, their results must be taken into account. Indeed, Choe et al. [30] and Ren et al. [31] carried out theoretical mechanistic studies using Ni(111). Both proposed that CO2 methanation proceeds with CO as intermediate, suggesting a mechanism based on several elementary steps consisting in two different processes: carbon formation (from CO2 dissociation into CO and O, and the subsequent CO dissociation into C and O) and C hydrogenation to CH4, being concluded that CO dissociation was the rate-determining step of the process. Zhao et al. [32] also developed self-supported cobalt nanoparticles composed of Cox(CoO)1–x with different ratios of x = Co/(Co + CoO) via reduction of precursor Co3O4. The Co0.2(CoO)0.8 sample exhibited the highest catalytic performances and was active even at low temperatures (160–180 °C), which was attributed to the increased amount and moderate binding of adsorbed CO2 on CoO sites. However, these results were obtained with a significantly high contact time.
Going back to the main purpose of this review, the following paragraphs will be divided based on the support nature applied in the CO2 methanation reaction, as shown in Figure 2. Inside each section, a brief elaboration of the type of metal used, the method of the support preparation, the way of the metal active phase deposition and the type of promoter used will be presented equally. Needless to say, a deep analysis of the efficiency of the prepared materials will be explained in order to help researchers in finding the best route for achieving a robust catalyst capable of reaching high catalytic conversions and high rates at even low temperatures.

2. SiO2-Based Catalysts

Regarding the use of supported catalysts, the firstly analyzed material is SiO2. The main catalytic systems found in the literature and the type of metal dispersed over the silica surface can be viewed in Table 1. with the most remarkable findings summarized below.

2.1. Monometallic SiO2-Supported Catalysts

Wu et al. [33] studied SiO2-supported catalysts with 0.5 and 10 wt% Ni incorporated by impregnation. They reported that the verified differences in the Ni0 particles size between the samples with different Ni loadings led to a strong effect on the kinetic parameters of CO2 hydrogenation, the formation pathways of CO and CH4 and the reaction selectivity. Higher Ni loading enhanced the CH4 selectivity while, with 0.5 wt% of Ni, the formation of CO was more relevant. In addition, Ye et al. [34] synthesized two Ni/SiO2 catalysts using different preparation methods: ammonia evaporation and impregnation. While ammonia evaporation led to the formation of nickel phyllosilicate, impregnation promoted the incorporation of Ni as NiO. The higher performances revealed by the sample prepared by ammonia evaporation were attributed to the NiPS phase, playing a relevant role in the reaction due to its lamellar nature, responsible for an increased specific surface area, a higher number of surface acids groups and the anchoring of well-dispersed Ni nanoparticles in the channels, establishing stronger metal-support interactions.

2.2. Bi- and Trimetallic SiO2-Supported Catalysts

Guo et al. [35] prepared Ni–Mg/SiO2 catalysts (10 wt% Ni and 1, 2 or 4 wt% Mg) in order to evaluate the effects of Mg loading and the impregnation strategy (co-impregnation or sequential) in the performances. Authors reported that the incorporation of 1 wt% of Mg as MgO was responsible for an enhancement of Ni species dispersion by suppressing sintering phenomenon on Ni particles. Additionally, MgO favored CO2 adsorption and activation. Complementary, co-impregnation led to the highest activity and stability, which was attributed to the synergistic effects established between Ni and Mg species. Park et al. [37] studied Pd/SiO2, Mg/SiO2, Pd–Mg/SiO2, Pd–Fe/SiO2, Pd–Ni/SiO2, Ni/SiO2 and Pd–Li/SiO2 catalysts prepared by reverse microemulsion. Authors prepared other two samples (Pd/SiO2 and Mg/Pd/SiO2) using the impregnation method for comparison purposes. They obtained the highest performances with 6.2%Pd–3.6%Mg/SiO2 sample (synthesized by reverse microemulsion), consisting of Pd aggregates within an amorphous Mg–Si oxide matrix. They also verified that replacing Mg by Fe or Ni led to the same activity while the selectivity to CH4 was considerably lower, especially in the Pd–Fe/SiO2 sample. Furthermore, Ali et al. [38] studied Ni–Co/SiO2 catalysts prepared by co-precipitation of the Ni and Co precursor salts and the Si source (TEOS) with 80 wt% Ni and 20 wt% Co. Authors evaluated the effect of the calcination temperature (300, 350, 400, 450 and 500 °C) in the metallic species and textural properties, with the best results being obtained with 400 °C and attributed to the lower size of the Ni and Co particles in this sample.
Branco et al. [41] studied the effects of adding lanthanides (La, Ce, Pr, Sm, Dy and Yb) to the formulation of Ni-supported SiO2 catalysts. Authors prepared silica by electrospinning and added the metals by impregnation method. Lanthanides were responsible for an important improvement of the catalytic performances, which was attributed to the enhancement of nickel reducibility and dispersion as well as to the increase of the basicity. Among all, Pr was identified as the most outstanding promoter. Trovarelli et al. [39] prepared SiO2-supported catalysts containing 1 wt% Rh and 1.6 to 28.4 wt% CeO2 (Rh/CeO2, Rh/SiO2, Ce/SiO2, Rh–Ce/SiO2). They concluded that smaller CeO2 crystallites are formed when supporting CeO2 on SiO2. In addition, Rh was found as responsible for a further re-dispersion of CeO2 crystallites. The high activity and selectivity revealed by Rh/CeO2/SiO2 samples was attributed to the presence of surface vacancies at the interface between Rh and CeO2, enhancing CO2 activation. Li et al. [40] studied the performances of LaNi1-xMoxO3 oxides with perovskite-type structure (6.0 wt% Ni and 1.0, 1.5 and 2.0 wt% of Mo) supported over SiO2. For comparison purposes, they also prepared three additional catalysts: LaNiO3/SiO2, LaMoO3/SiO2 and unsupported LaNiO3. As a result, the presence of Mo in the sample contributed to higher catalytic performances and stability. Indeed, the positive effect of this metal was further enhanced when intensifying its loading since it revealed much higher resistance to Ni nanoparticles sintering. Finally, Vogt et al. [42] evaluated the effect of KOH (derived from H2 production process) in the activity of Ni-based SiO2 catalysts. Authors applied in situ (setup combining H2O electrolysis and CO2 methanation steps) and ex situ (KOH impregnation over Ni/SiO2 catalysts) strategies and verified that, while the first increases the obtained catalytic performances, the second leads to catalysts deactivation. They attributed the negative effect of impregnating KOH over the catalysts to the establishment of stronger interactions with CO, thus limiting CO2 methanation. On the contrary, KOH aerosols (in situ) could, according to the authors, increase the hydrogenation rate of CHx species and/or even facilitate water formation/desorption.

2.3. Mesoporous Silica-Based Catalysts

In the last years, mesoporous-based samples and especially silicas such as KIT–6, SBA–15, MCM–41 or KCC–1 have gained a great attention due to their positive feature of managing in confining nickel particles preventing sintering. The main reported findings on such type of supports are discussed below and summarized in Table 2.
  • Mesoporous silica nanoparticle (MSN)-based catalysts
Mesoporous silica nanoparticles (MSNs) present unique characteristics such as an ordered and nanosize porous structure with well-defined and tunable pore sizes (1.5–10 nm) and high textural properties (e.g., pore volume, surface area) [63]. Aziz et al. [43] studied metal-promoted mesostructured silica nanoparticles (MSN, prepared by sol-gel method) using 5 wt% of Rh, Ru, Ni, Fe, Ir, Cu, Zn, V, Cr, Mn, Al or Zr incorporated by impregnation. Authors found that Rh/MSN catalyst was the most active while Ir/MSN presented the poorest activity. Additionally, no catalytic activity was revealed by the Zn/MSN, V/MSN, Cr/MSN, Mn/MSN, Al/MSN and Zr/MSN in the studied temperature range. The same authors [63] prepared 5 wt% Ni-based catalysts supported on MSN, MCM–41, HY zeolite, SiO2 and Al2O3. The activity was found to follow the order: Ni/MSN > Ni/MCM–41 > Ni/HY > Ni/SiO2 > Ni/Al2O3. The better results of the MSN supported catalyst were assigned to the presence of both intra- and inter-particle porosity, which led to a higher concentration of basic sites in this sample. Researchers who worked on these materials also verified the relevance of the basicity in the catalytic performances and highlighted the fact that CO2 methanation activity increased with the concentration of basic sites. Again, same authors [44] studied the influence of Ni loading and the presence of steam in the reactor feed flow. Regarding the metal content, the catalytic results of the prepared materials followed the order: 10%Ni/MSN ≈ 5%Ni/MSN > 3%Ni/MSN > 1%Ni/MSN. However, higher Ni loadings decreased catalysts crystallinity, surface area and basic sites. Consequently, the order of the obtained performances was due to a balance between basic sites and available metallic Ni sites. Additionally, authors emphasized that the presence of steam in the feed stream induced a negative effect on the activity (30% of activity loss after 5 h) and attributed this behavior to the formation of CO2 through the water gas shift (WGS) reaction between intermediate CO molecules and the excess of water, to the acceleration of Ni sintering and to the collapse of the support. Recently, Nguyen et al. [45] reported a highly active 50 wt% Ni-MSN catalyst. MSN support was prepared by sol-gel method while Ni was incorporated by impregnation using urea together with nickel nitrate, which was responsible for an enhancement in the reducibility and dispersion of Ni species and favored CO2 adsorption capacity. They also optimized the calcination time (2, 3 or 4 h) and verified the best results when using the intermediate value (3 h). They finally included 1% of CO in the reaction flow, increasing the overall yield by its effective conversion to CH4.
  • KIT–6-based catalysts
KIT–6 mesoporous silica presents a 3–D channel network with thick pore walls, hydrothermal stability, high surface area and large pore volume [64,65]. Its interconnected 3–D structure promotes the location of active species inside the porous system, favoring metallic dispersion and diffusional processes and hindering pores blockage [64,65]. Zhou et al. [46,47] prepared mesoporous Co/KIT–6 and Co/meso–SiO2 catalysts with the same Co content (wt% not specified) via hydrogen reduction. Co was better dispersed on KIT–6 than on SiO2, explaining the higher activity revealed by Co/KIT–6 sample. In addition, the highly ordered and mesoporous structure of the Co/KIT–6 catalyst improved the selectivity to CH4. Liu et al. [48] studied the effect of Co content on the performances of Co/KIT–6 catalysts, concluding that the best Co loading was 25 wt%. However, for the 15 to 20 wt% Co loading based samples there was a limitation in the number of active sites. In addition, 30 wt% Co catalyst presented larger and less dispersed Co particles with lower specific surface area and a disordered pore structure that hindered the diffusion and migration of CO2 and H2 molecules towards Co active sites. The same authors [49] investigated the influence of the reduction temperature in Co/KIT–6 catalysts activity towards CO2 methanation. They obtained the best results by reducing the sample at 400 °C, due to the larger specific surface area formed, the higher number of Co species reduced and the increased CO2 adsorption and activation capacity. Additionally, Merkache et al. [50] prepared Fe–KIT–6 materials with Si/Fe molar ratios of 10, 30 and 60. The best performances were reported for the sample with higher Fe content (nSi/nFe = 10) and were attributed to the higher concentration of metallic active sites on the KIT–6 structure. Cao et al. [51] reported recently Ni–V/KIT–6 catalysts, with metal loadings varying from 5 to 40 wt% Ni and 0.1 to 2 wt% V. They obtained the best results when using 20 wt% Ni and 0.5 wt% V, achieving a high metal dispersion. The surface basicity of the materials was further boosted by a synergistic effect between Ni and V (present as V2O5 in the catalysts).
  • SBA–15-based catalysts
SBA–15 presents dual porosity composed by mesoporous and intrawall micropores or secondary mesoporous channels. It presents large surface areas, thick pore walls (3–6 nm), uniformly distributed cylindrical channels (5–10 nm) and high hydrothermal stability [66]. Liu et al. [52] prepared Ni-based SBA–15 catalysts by one-pot hydrothermal method with a 3–D network structure and 15 wt% of Ni. For comparison purposes, they also prepared an impregnated sample containing 15 wt% Ni over SBA–15 support. Ni/SBA–15One-pot catalyst presented higher surface area, larger pore volume, better dispersion, higher catalytic activity and anti-sintering properties for Ni0 particles. In addition, Lu et al. [53] prepared Ni/SBA–15 catalysts by grafting method. A chemical bond was formed between silicon and Ni atoms via an oxygen atom (–O–Ni–O–Si–O–), by using Ni ammonia complex ions with NH3/Ni molar ratios of 2 to 4 (equivalent to 4 to 10 wt% Ni). No bulk nickel oxides were depicted in the Ni-grafted SBA–15 sample and the CO2 conversion and methane selectivity were higher than those obtained for equivalent Ni/SBA–15 catalysts prepared by impregnation. Authors also indicated that the increase of the Ni content enhanced the activity and selectivity. Bacariza et al. [54] prepared Ni and Ce/Ni impregnated SBA–15 catalysts with 15 wt% Ni and 15 wt% Ce and with SBA–15 prepared by a conventional and a microwaves-assisted hydrothermal treatment. Despite the synthesis method of the support, relevant performances were detected, with the microwave-method sample presenting slightly higher conversions, especially between 250 and 350 °C. Additionally, an enhancement of the performances was observed by adding Ce, being this attributed to the role of CeO2 species as active sites for CO2 activation. Finally, Li et al. [55] investigated bimetallic Ni–Pd/SBA–15 alloy catalyst for selective hydrogenation of CO2 to methane by changing the atomic ratio of Ni/Pd (Ni0.25Pd0.75, Ni0.50Pd0.50 and Ni0.75Pd0.25). Regarding the metal content, the order of catalytic activity was the following: Ni0.75Pd0.25/SBA–15 > Ni0.50Pd0.5/SBA–15 > Ni0.25Pd0.75/SBA–15 > Ni/SBA–15 > Pd/SBA–15. The excellent catalytic activity of Ni0.75Pd0.25/SBA–15 was related to the synergistic effect between Ni and Pd species. In this way, the charge transfer from Pd to Ni atoms provided a negative charged surface, which promoted the activation of CO2 and the dissociation of intermediates into CH4, crucial for improving the catalytic performance.
  • MCM–41-based catalysts
MCM–41 presents a purely mesoporous 2–D hexagonal structure formed by unidirectional and non-interconnecting cylindrical pores (2–10 nm of diameter) [67,68]. Du et al. [56] reported 1 to 3 wt% Ni containing MCM–41 samples with Ni ions incorporated into the MCM–41 structure. Well-dispersed Ni0 species were formed in the samples, especially after reduction at high temperature (700 °C), leading to superior performances. Additionally, Bacariza et al. [54] prepared Ni and Ce/Ni-based MCM–41 catalysts with 15 wt% Ni and 15 wt% Ce. These samples presented greater activity than equivalent SBA–15 catalytic systems. MCM–41 was found as capable of promoting the metal-support interactions, favoring the CO2 and H2 activation. Additionally, the lower affinity of MCM-41 to water served as another advantage in encountering the water produced as a product in the reaction (according to the literature [44,69,70]). Finally, and as mentioned previously for other type of mesoporous silica supports, the incorporation of Ce to the Ni/MCM–41 sample triggered a boost in its performance during the methanation reaction. To better emphasize on this point, Wang et al. [57] also prepared Ni/MCM–41 promoted with CeO2, containing different CeO2 loadings (0, 10, 20, 30 wt%). The addition of CeO2 improved the dispersion of Ni on the support, providing a larger specific metallic surface area and a higher reducibility. Besides, the interaction between Ni active sites, CeO2 and the support facilitated the activation of adsorbed CO2 species at low temperatures. Thus, within a certain composition range, the incorporation of CeO2 was able to improve the catalytic performances, with 20 wt% CeO2 leading to the best catalytic activity and stability. Recently, Taherian et al. [58] reported the beneficial effects of incorporating Y and Mg to Ni-based MCM–41 catalysts synthesized by direct synthesis method. The incorporation of these two metals to the MCM–41 mesoporous system decreased the average Ni0 size by improving Ni species reducibility and creating new active sites for CO2 activation.
  • KCC–1-based catalysts
KCC–1 is formed by a fibrous surface morphology arranged in a 3–D structure forming nanospheres. Its fibrous structure facilitates the accessibility to the available surface area and improves the hydrothermal and mechanical stability [71]. Hamid et al. [59] prepared a KCC–1 material by microemulsion and used it as catalyst for CO2 methanation in absence of active metals sites. They compared the synthesized sample to MCM–41 and SiO2 materials. As a result, KCC–1, which presented higher basicity and oxygen vacancies than MCM–41 and SiO2, led to the best performances. It was proposed that both CO2 and H2 were activated and dissociated in the KCC–1 support oxygen vacancies, where the CO2 interaction with the material is related to the basicity. The same authors [60] reported promising results when loading KCC–1 mesoporous support with 5 wt% of Ni, Co or Zn, with superior behavior found for Ni/KCC–1. Recently, Lv et al. [61] evaluated the effects of impregnating increasing Ni loadings (5 to 25 wt%) over a fibrous KCC–1 nanospheres synthesized by microemulsion hydrothermal method. 20 wt% Ni sample exhibited better results than those found for equivalent Ni/SiO2 and Ni/MCM–41 catalysts. The formation of dendrimer mesoporous channels not only improved the metallic dispersion due to the confinement effect but also hindered the occurrence of sintering processes. Furthermore, the appearance of intermediate carbonate species was faster when using Ni/KCC–1 instead of Ni/SiO2 or Ni/MCM–41.
  • FDU–12-based catalysts
FDU–12 is a 3–D ordered mesoporous silica with a face-centered cubic structure. It presents a window of ~9 nm and pore cages with sizes ranging from 12 to 42 nm. Its structure presents advantages when comparing to SBA-15 or MCM-41, such as higher resistance towards pores blockage due to its cage-like network [62]. Recently, Liu and Dong [62] reported, for the first time, Ni-based FDU–12 catalysts for CO2 methanation. Authors prepared a series of 10 wt% Ni-containing FDU–12 and evaluated the influence of the impregnation solvent (water or ethylene glycol) and the effects of doping the catalysts with trehalose and Ce (3 wt%). A remarkable reduction of the average Ni0 particle size was observed when using ethylene glycol (5.9 nm) instead of water (10 nm). These values were further improved (from 5.9 to 2.6–3.3 nm) by incorporating increasing amounts of trehalose in the impregnation solution, while Ce did not influence this parameter. The enhancement of the metallic dispersion was responsible for an improvement of the catalytic performances, being a CH4 yield >75% obtained at 425 °C for the 10Ni/FDU–12 catalyst prepared using ethylene glycol and a fraction of trehalose. Complementarily, the already reported positive effect of Ce in the reaction was proved, being the activity boosted at lower reaction temperatures (<350 °C). Finally, authors also verified a reduction of the activation energy with the optimization of the impregnation solvent and Ce incorporation (from 116.23 to 77.49 kJ mol−1).
  • Comparative analysis
Taking into account the various experimental conditions used in the works reported in the literature, a proper comparison among the different types of mesoporous silica supports is difficult to establish. In this way, in an attempt to compare monometallic catalysts prepared by impregnation method, CH4 formation rates as a function of the mass of catalyst (mol CH4 s−1 gcat−1) and metal (mol CH4 s−1 gmetal−1) were determined, far from thermodynamic equilibrium, for a series of Ni and Co-supported SiO2 [34] (reference), MSN [45], KIT–6 [46], SBA–15 [54], MCM–41 [54], KCC–1 [61] and FDU–12 [62] (Figure 3). As seen, the use of Ni instead of Co leads generally to better performances, which insists on the suitability of Ni as CO2 methanation active metal. Additionally, SBA–15, MCM–41 and FDU–12 are the materials leading to the most outstanding results, both per gcatalyst and gmetal. The catalytic performances obtained for the different mesoporous silicas will be likely influenced by several factors such as the metallic dispersion, the activation induced by the support (resulting from the metallic-support interactions), the location of the metal, the ability to activate carbon dioxide (adsorption sites, pores acting as CO2 reservoir) or even the hydrophobicity. Even with well-designed, systematic and specific studies reported in the literature, no definitive conclusions regarding the superior performances revealed by SBA-15, MCM-41 or FDU-12 catalysts can be drawn (e.g., it is not possible to quantify certainly the contribution of each individual factor to the activity).

2.4. Mechanistic Aspects

Wu et al. [33] reported that Ni loading affects the selectivity to methane and the mechanism pathways over Ni/SiO2 catalysts (Figure 4–Left). Indeed, they prepared two 0.5 and 10 wt% Ni/SiO2 samples and proposed that formate species in a monodentate configuration are involved in CO2 hydrogenation on both cases. On one hand, the consecutive pathway, which was favored on small Ni particles (0.5%Ni/SiO2 catalyst, Ni0 average size < 1 nm), was attributed to low H2 coverage on the Ni surface, leading to dissociation of formate intermediates resulting in CO formation and high CO selectivity. On the other hand, the reaction on large Ni particles (10%Ni/SiO2 catalyst, Ni0 average size of 9 nm, stronger H2 adsorption and enhanced H2 coverage), was proposed to be controlled by mixed consecutive and parallel routes, promoting formate species hydrogenation to CO or CH4 as part of a parallel reaction pathway. The sites corresponding to kink, corner or step positions on the Ni/SiO2 surface were proposed as the primary active sites for CO2 hydrogenation. In addition, Park et al. [36,37] proposed a bifunctional mechanism for Pd–MgO/SiO2 catalysts (Figure 4–Right). They considered CO2 adsorption over MgO as a surface carbonate that is sequentially hydrogenated to form methane. Consequently, in this mechanism an active metal (Pd in this case) is required for the dissociation of H2 and the supply of H atoms to the CO2 derived species activated on Mg sites.
Regarding mesoporous silica-based catalysts, Aziz et al. [43] considered CO as a reaction intermediate in CO2 methanation over Ni/MSN catalysts. Indeed, they proposed the dissociation of CO2 into CO and O on the Ni surface and the progressive hydrogenation of CO into methane. Due et al. [56] also suggested that CO2 methanation passes through CO as intermediate over Ni/MCM-41 catalysts. Finally, Hamid et al. [59] suggested a mechanism for CO2 methanation over KCC-1 (Figure 5), where CO2 was activated on oxygen vacancies forming adsorbed carbonyl species. They demonstrated that linear carbonyl species were the precursors for CH4 formation. Additionally, they implied that hydrogen was activated and dissociated in E’ centres (defined as dangling bonds, ≡Si*, formed on the SiOx material for x < 2) acting as active sites and giving rise to atomic H species.

3. Al2O3-Based Catalysts

Alumina-based catalysts are, without doubt, the most commonly used in carbon dioxide methanation reaction. In summary, the main reported catalytic systems are disclosed below in Table 3, while the corresponding conclusions of the published works will be deeply discussed in the upcoming paragraphs.

3.1. Monometallic Al2O3-Supported Catalysts

Garbarino et al. [73] studied Ni/Al2O3 catalysts prepared by impregnation and with 16, 39 and 125 wt% of Ni by wt% of support. Authors performed catalytic tests without a pre-reduction treatment and obtained very low activities, confirming that Ni0 species are indeed required for obtaining significant CO2 conversion degrees. They demonstrated also that the higher the Ni content, the larger the particles but also the greater the number of active sites. Consequently, these two parameters affecting in opposite ways the performances contributed to the following activity results order: 125%Ni/Al2O3 > 16%Ni/Al2O3 > 39%Ni/Al2O3. The same authors [72] synthesized Ni nanoparticles (NPs) with < 8 nm of size and compared their performances with the ones obtained for 125%Ni/Al2O3 catalyst from their previous work. Better outcome was obtained for 125%Ni/Al2O3 (with Ni particles of ~36 nm), while Ni NPs resulted in poorly active metallic sites. They came up with a hypothesis that the support could play a role in the activation of CO2 while Ni nanoparticles could be mainly responsible for the H2 dissociation, which explains the better results obtained for the supported Ni catalyst. Lately, they also reported a study comparing two 3 wt% Ru and 20 wt% Ni based Al2O3 catalysts [75] revealing that, despite the higher metal content of the Ni sample, Ru/Al2O3 led to higher performances due to the favored dispersion of Ru species. Quindimil et al. [74] investigated the influence of the metal nature (Ni, Ru) and loading (4 to 20 wt% for Ni and 1 to 5 wt% for Ru) on CO2 methanation catalysts supported on Al2O3. Authors clearly observed that, in both cases, higher metal loading enhanced the catalytic performances, which could be due to the increase of active metal sites available for H2 dissociation. However, for Ni/Al2O3 series, metallic dispersion decreased with metal loading due to the formation of larger particles, while the dispersion of Ru/Al2O3 was not significantly influenced. Moreover, it was suggested that the reducibility of the catalysts plays a key factor in the catalytic performances. Indeed, the complete reduction of ruthenium species was achieved at 250 °C, whereas for nickel-based catalysts a temperature of 900 °C was required, reducing the available Ni0 sites for H2 dissociation. Besides, at low temperature (T < 300 °C), turnover frequency (TOF) values for Ru/Al2O3 were considerably higher than those found for Ni/Al2O3, which indicates that Ru is more effective than Ni in H2 dissociation. Taking this into account and considering the saturation effect on CO2 conversion with metal loading, the best performances were obtained by 12%Ni/Al2O3 and 4%Ru/Al2O3 formulations, with Ru catalyst leading to higher conversion and CH4 selectivity. As a final remark, no works studying the influence of chloride and/or chlorine doping over Ru/Al2O3 catalysts were found in the literature so far. This strategy could be interesting for Sabatier reaction, as promising results were reported for CO methanation [99,100,101].

3.2. Bimetallic Al2O3-Supported Catalysts

Hwang et al. [76] studied 20%Ni–5%Fe–Al2O3 catalysts prepared by co-precipitation using different precipitation agents ((NH4)2CO3, Na2CO3, NH4OH or NaOH). They verified that Ni0 dispersion was enhanced by using ammonium carbonate, favoring both the CO2 conversion and the CH4 selectivity. Serrer et al. [77] also studied the effect of Fe in Ni–Fe–γ–Al2O3 catalysts synthesized via homogeneous precipitation with urea, a method optimized in previous publications from the same authors [102,103]. During catalyst activation, authors observed a synergistic effect between nickel (13 wt%) and iron (4 wt%) that led to higher fractions of reduced nickel compared to the monometallic 17 wt% Ni/γ–Al2O3 reference. Ni remained in its reduced metal state under CO2 methanation conditions. On the contrary, the oxidation of Fe0 was immediately observed. This indicated that Fe exhibits a highly dynamic behavior through a Fe0 ⇌ Fe2+ ⇌ Fe3+ redox cycle, likely located at the interface between the FeOx clusters and the surface of the metal particles, which promoted CO2 dissociation during the reaction. Burger et al. [79] reported also highly active and thermally stable Mn- and Fe-promoted NiAlOx catalysts prepared by co-precipitation and tested under CO2 methanation conditions using a pressure of 8 bar. They concluded that while Mn improves the CO2 adsorption capacity (especially in terms of medium basic sites) without significantly modifying the available nickel surface area, Fe promotes thermal stability through the formation of a Ni–Fe alloy and slightly increases the CO2 uptake. These authors [104] also reported an improvement of NiFeAlOx catalysts deactivation resistance at elevated temperature and pressure in thermodynamic equilibrium. This phenomenon was attributed to the partial segregation of (γFe, Ni) nanoparticles formed during the activation step giving rise to Fe2+, offering an alternative reaction pathway through CO2 activation on disordered FeyO. They also studied the effects of incorporating Fe by surface redox reaction (SRR) method on Ni-based alumina catalysts [78]. They prepared a 11 wt% Ni/Al2O3 catalyst by deposition-precipitation and a 48 wt% NiAlOx material by co-precipitation method. In the case of Ni/Al2O3-derived catalysts, higher Fe loadings (0.5 to 1.8 wt%) induced better performances (tests done at 8 bar), which was attributed to the enhancement of the electronic properties due to the formation of Ni–Fe alloys. In addition, increasing Fe loadings (3.6 to 8.6 wt%) increased the apparent thermal stability of NiAlOx-derived catalysts, which was ascribed to the partial segregation of the alloyed Ni–Fe particles, in accordance with their previous studies.
Xu et al. [80] investigated a series of Co–Ni doped ordered mesoporous Al2O3 (OMA) oxides with different Co/(Co+Ni) molar ratios fabricated by the one-pot evaporation-induced self-assembly (EISA) method. This synthesis method produced homogeneous incorporation of metals among the Al2O3 framework achieving as well a high Co and Ni dispersion. Bimetallic samples exhibited higher catalytic performances than those obtained for the monometallic Ni and Co OMA catalysts, especially at lower temperatures. This trend was related to the synergetic effect between Co and Ni (Figure 6), since Co presents high activity towards CH4 formation at low temperature, owing to remarkable CO2 activation capacity and allowing the decrease of the apparent activation energy. In addition, they observed that only an appropriate Co/(Co + Ni) molar percentage (20%) could maximally enhance the catalytic performances. Besides, catalysts showed an outstanding thermal stability and anti-sintering properties. Alrafei et al. [81] also optimized Ni (5 to 25 wt%) and Ni–Co catalysts supported on γ–Al2O3. Authors prepared the alumina support in the form of extrudates and incorporated Ni and Co by impregnation (co-impregnation for bimetallic samples). They observed, in accordance with other studies already discussed, that Ni loading strongly influenced the catalytic activity, with values above 15 wt% leading to no remarkable differences. In terms of Co effect, they observed that this metal presented only a positive impact in the performances when Ni loading was 10 wt%. They attributed this to an enhancement of nickel species reducibility and dispersion.
Several studies have dealt with the beneficial incorporation of lanthanides and metal oxides to CO2 methanation catalysts due to their appropriate electronic or acidic/basic properties, among all [105,106]. Indeed, Rahmani et al. [82] reported Ni-based catalysts supported on mesoporous γ–Al2O3 promoted with Ce, Mn, Zr and La oxides and prepared by impregnation, all samples containing 20 wt% Ni and 2 wt% of promoter. As authors identified Ce as the promoter with the most effective results, they studied the effect of Ce loading (2, 4 and 6 wt%) being observed that, despite the small differences between samples, 2 wt% Ce seems to be the most suitable content. Cerium positive effect was explained, as seen in SiO2 based materials section, by the presence of oxygen vacancies in the cerium containing samples, enhancing CO2 activation. In the Mn, Zr and La samples, the presence of the promoters had only an effect in the coverage of Ni sites, reducing the accessibility and causing blockage, so that no enhancement of the performances was found. Liu et al. [83] also reported the effect of adding CeO2 to 15%Ni/Al2O3 catalysts. The catalytic performances were strongly dependent on the CeO2 content, being the sample with 2 wt% Ce the one leading to the best results, as also found by Rahmani et al. [82]. The positive effect of Ce was attributed to the improved reducibility of Ni species (e.g., weakening of Ni–Al2O3 interactions) and enhanced activation of CO2 on Ce sites. Guo et al. [84] developed an innovative catalyst by synthesizing Ni–Al2O3 LDHs (layered double support) and using CeO2 as promoter. Authors obtained the best performances when using ~50 wt% CeO2 due to the small and highly dispersed Ni0 particles formed and due to the existence of oxygen vacancies that directly increased the active sites and promoted CO2 adsorption. Tada et al. [85] verified also the effect of CeO2 loading on the performances of Ru/CeO2/Al2O3 catalysts. They observed that the performances followed the trend: Ru/30%CeO2/Al2O3 > Ru/60%CeO2/Al2O3 > Ru/CeO2 > Ru/Al2O3. With a lower CeO2 content (30 wt%), a larger surface area was achieved due to the smallest average size of CeO2 crystallites in this sample when comparing to the rest of the tested catalysts. Additionally, authors suggested, according to the results of performed mechanistic studies, that samples without CeO2 presented further difficulties in the dissociation and decomposition of intermediate formate species. Along with the previously discussed studies, Ahmad et al. [86] tested the effect of La, Ce, Pr, Eu and Gd as promoters of Ni/γ–Al2O3 catalysts with 5 wt% of the promoter and 12 wt% of Ni added by impregnation. In this case, Pr was pointed out as the best promoter, even if all lanthanides led to an increase of Ni dispersion without remarkable effects on the textural properties when comparing to the 12%Ni/γ–Al2O3 catalyst. Finally, Karam et al. [87] reported Ni–Mg–Al2O3 catalysts prepared by EISA one-pot and optimized the materials in terms of Mg and Ni loadings. The best performances were achieved with 7 wt% Mg and 15 wt% Ni, being higher than those revealed for other Al2O3 samples in the literature. This outcome was due to the important improvement of the metallic dispersion and surface basicity.

3.3. Trimetallic Al2O3-Supported Catalysts and Composites

Wang-Xin et al. [88] prepared highly dispersed Ni–Ce–Zr/γ–Al2O3 catalysts by citric acid assisted impregnation. The addition of citric acid in the preparation procedure promoted the dispersion of the Ni–Ce–Zr oxide species over the γ-Al2O3 surface and improved metal-support interactions. Consequently, Ni–Ce–Zr/γ–Al2O3 catalysts were highly active and the higher the citric acid content used in the synthesis method the better the results. Toemen et al. [89] studied Ru/Mn/Ce/Al2O3 catalysts prepared by sequential impregnation with the aim of optimizing the Ce loading as well as the calcination temperature. Firstly, calcination at 1000 °C was found as the optimum temperature when comparing with the results obtained for the samples calcined at 400, 600 and 800 °C due to the modification of the phases formed during this thermal treatment. Additionally, Ce loadings of 55 to 85 wt% were tested, being the best results obtained for the 65 wt% Ce containing sample, which was attributed to higher dispersion and the favorable morphology of this catalyst.
Zamani et al. [107] carried out a deep study regarding the use of M*/Mn/Cu–Al2O3 (M* = Pd, Rh and Ru) catalysts prepared by impregnation. After comparing three catalysts containing the same Pd, Rh and Ru content over Mn/Cu–Al2O3, the best material was the one containing Ru. Consequently, as a second step, the Ru loading (1.6 to 18.4 wt%) and the calcination temperature (831 to 1168 °C) were optimized by using Design-Expert® software, where 10.9 wt% Ru and a calcination temperature of 1035 °C were described as the best conditions. Bakar et al. [90] studied Ru and Pd supported Mn/NiO/Al2O3 catalysts and reported Ru as a more favorable noble metal for this reaction. They also optimized the Ru/Mn/Ni ratio and the calcination temperature with better performances obtained when the parameters were 5:35:60 and 1000 °C, respectively. Authors defined this optimum catalyst as spherical nanoparticles with aggregated and agglomerated mixtures of metal species on the surface. It must be highlighted that the authors did not report any kind of pre-reduction treatment. However, they considered that unreduced NiO species could be active sites for CO2 methanation. To further confirm their hypothesis, they synthesized a NiO based Al2O3 catalyst and reported ~15 % of CO2 conversion without carrying out a pre-reduction treatment. They asserted the potential role of NiO as an active site for the reaction since, even after the tests, no reduced nickel species were observed in the samples. Franken et al. [108] prepared CoAl2-xMnxO4 (x = 0, 0.1, 0.5, 1) solid solutions by co-precipitation with the aim of understanding the effects of Mn on catalysts properties and performances. CoMn0.5Al1.5O4 led to the best performances while the introduction of Mn into the spinel support was responsible for a reduction in the basicity and an improvement of the hydrogen spillover from Co active sites to the support surface. These modifications led to a reaction mechanism with formate species as intermediates (no CO) accompanied by a remarkable decrease of the activation energy.
Do et al. [91] studied the effect of Ca incorporation (0, 1.0, 3.0, 5.0, 7.0 and 10.0 %mol) to perovskite NiTiO3 (containing ~11 wt% Ni) loaded by impregnation on γ–Al2O3. Ca incorporation favored the nickel dispersion, boosted the CO2 adsorption capacity and improved the textural properties, which resulted not only in a remarkable increase of the performances but also in a significant stability under reaction conditions for 10 days.
Several authors also reported the suitability of Al2O3-based composites and mixed oxides as supports for CO2 methanation catalysts. Indeed, Cai et al. [92] studied 12%Ni/(ZrO2–Al2O3) catalysts checking not only the promoter effect of ZrO2 but also the effects of the ZrO2–Al2O3 composite preparation method (co-precipitation, impregnation of Zr over commercial γ–Al2O3 or drying and impregnation followed by precipitation). In all cases, Ni was incorporated by impregnation. Regarding the preparation method of ZrO2–Al2O3, impregnation followed by precipitation led to the best catalytic performances with high stability due to the enhanced Ni species reducibility and dispersion over this support. Additionally, this sample also presented the highest basicity, which was considered by the authors as a key parameter, also favoring the methanation performances. Regarding the effect of the ZrO2 loading (0, 3, 9 and 15 wt%), 3 wt% ZrO2 enhanced both the activity and the stability. Lin et al. [93] also studied mesoporous Al2O3–ZrO2 modified Ni catalysts prepared via a single-step epoxide-driven sol-gel method and varying the Al/Zr ratio and Ni loading. They observed that the incorporation of ZrO2 into the Ni/Al2O3 structure decreased the interaction between Ni species and Al2O3, promoting nickel reducibility and dispersion. Thus, increasing ZrO2 loading favored the formation of active metallic Ni sites and surface oxygen vacancies, which represent two key factors for improving the catalytic activity at low temperatures. However, increasing both Zr and Ni loading was found as beneficial only for a certain range, with the catalyst containing 20% wt% Ni and an Al/Zr ratio of 1.0 presenting the best results.
Furthermore, Yang et al. [95] prepared CaO–Al2O3 composites by co-precipitation and incorporating Ni by impregnation. The best reported sample contained 15 wt% of Ni and a CaO:Al2O3 ratio of 20:80. Authors concluded that CaO was responsible for restraining the growth of NiO nanoparticles, improving its dispersion and creating a moderate interaction between NiO and Al2O3. Liu et al. [96] also reported ordered mesoporous Ni-Ru-doped CaO–Al2O3 composites synthesized by one-pot evaporation-induced self-assembly method with 10, 1 and 2 wt% of NiO, RuO2 and CaO, respectively. This material was compared to reference samples such as Ni–Al2O3, Ni-Ru–Al2O3 and Ru–Al2O3. Monometallic Ru or Ni catalysts presented considerably similar CO2 conversions due to the close average size values of the metal species in both catalysts (~10 nm). However, the selectivity to methane was quite poor in the Ni sample. The better results obtained for the Ni-Ru bimetallic sample were attributed by the authors to the synergistic effect between Ni and Ru and to the favored H2 activation developed by Ru species. Finally, the addition of CaO enhanced the CO2 adsorption capacity providing Ni–Ru–Ca sample a finer performance.
Xu et al. [94] studied Ru/TiO2–Al2O3 catalysts prepared by impregnation. When comparing the performances of two 5 wt% Ru catalysts supported on Al2O3 and TiO2–Al2O3, the second catalytic system exhibited a much higher activity. This was due to the more dispersed particles found in the Ti-containing sample as a result of establishing a strong interaction between Ru and TiO2. The effect of TiO2 content (5, 10 and 15 wt%) and the calcination temperature after TiO2–Al2O3 synthesis were also evaluated. While TiO2 loading did not influence the results, CO2 conversions increased with higher calcination temperatures (from 600 to 950 °C) and decreased for 1100 °C, being 950 °C the optimum.
Abate et al. [97] prepared γ–Al2O3–ZrO2–TiO2–CeO2 composites as supports for Ni catalysts. The best results were obtained for the sample presenting the composite composition of γ-Al2O3:ZrO2:TiO2:CeO2 = 55:15:15:15, prepared by impregnation-precipitation method using commercial γ-Al2O3 powder as a host, and 20 wt% Ni. The incorporation of Ni to the composites instead of γ–Al2O3 provoked enhanced Ni0 dispersion and reducibility, due to the modification of the metal-support interactions. Mebrahtu et al. [98] also studied Ni catalysts supported on ternary and quaternary alumina–zirconia–titania–ceria mixed oxides. Results showed that, for ternary oxide supported Ni, the catalytic activity depends on textural properties improvement while, for quaternary oxide supported Ni, it depends mainly on reducibility and metallic dispersion. Moreover, TiO2 and ZrO2 incorporation was found to modify mainly textural properties while CeO2 addition promoted Ni dispersion and reducibility. Besides, 15 wt% of CeO2 strengthened the catalyst stability, decreasing the deactivation rate of about one order of magnitude for the best ternary system (20%Ni/90%γ–Al2O3–5%ZrO2–5%TiO2).
Finally, some authors focused on finding new methods to prepare Ni/Al2O3 catalysts. This is the case of Daroughegi et al. [109], who used ultrasound-assisted co-precipitation; Xu et al. [110], who used cold plasma; Song et al. [111], who used microwaves; Schubert et al. [112], who used double flame spray pyrolysis; Aljishi et al. [113] and Xu et al. [114], who used evaporation induced self-assembly technique (EISA) or Shang et al. [115], who used partial hydrolysis of aqueous solutions containing Al(NO3)3 and Ni(NO3)2 with (NH4)2CO3, without templates or organic surfactants. Additionally, Le et al. [116] studied the effect of the type of Al2O3 crystalline phase in the performances of Ni/Al2O3 catalysts.

3.4. Mechanistic Aspects

Beuls et al. [117] proposed some insights for Rh/γ–Al2O3 catalysts mechanism at 50–150 °C and 2 atm. They suggested that CO2 is adsorbed over the material by dissociation, forming CO(ads) and O(ads) and inducing Rh species oxidation. Indeed, Rh species oxidation state played a crucial role in the distribution of the adsorbed species on the catalysts, with Rh–dicarbonyls being identified as those leading to the formation of CH4. In short, authors consider CO as the true reaction intermediate in their catalytic system. Garbarino et al. [73] suggested that CO2 methanation over Ni/Al2O3 catalysts occurs at the expense of CO intermediate on the corners of Ni nanoparticles interacting with alumina, likely due to the formation of formate or other oxygenate species and their further hydrogenation into CH4. Zhang et al. [118] studied the mechanistic implications of different Ni loadings on Ni/ɣ–Al2O3 catalysts by diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) studies. They observed that, while low Ni contents favor the occurrence of reverse water gas shift (RWGS) reaction promoting CO formation, increasing loadings hinder the formation of stable and non-reactive carbonate species and promote formates conversion to CH4, boosting the activity towards CO2 methanation rather than RWGS. Finally, Cárdenas-Arenas et al. [119] proposed that CO2 methanation proceeds over Ni/Al2O3 through the adsorption and consecutive chemisorption of CO2 over hydroxyl groups created due to the H2 reduction of the NiO–Al2O3 interface (Figure 7). They reported that chemisorbed CO2 hampered carbon dioxide molecules chemisorption/dissociation and led, alternatively, to the formation of water and formates species. Even if formate species could be successively hydrogenated to methane, part of them gave rise to CO, decreasing CH4 selectivity.

4. CeO2-Based Catalysts

Even if cerium oxide has been typically used as a promoter for CO2 methanation catalysts, several authors also reported interesting results when using it as support. Indeed, the most notable results can be seen in Table 4. and their main findings are discussed in the following sections.

4.1. Monometallic CeO2-Supported Catalysts

Tada et al. [85,120] studied 10 wt% Ni-based catalysts supported on CeO2, γ–Al2O3, TiO2 and MgO and identified CeO2 as the best choice. The surface coverage by CO2-derived species on CeO2 surface as well as the partial reduction of CeO2 surface was reported to be responsible for the great performances of the 10%Ni/CeO2 catalyst. Atzori et al. [121] also studied NiO–CeO2 mixed oxides highly active for CO2 methanation and prepared by soft-template and impregnation methods. Soft-template produced Ni0 particles of ~7 nm while impregnation contributed to the formation of larger particles (>30 nm) after the pre-reduction treatment. Despite the synthesis method and the differences in the Ni0 size, both samples presented high catalytic performances. Consequently, authors suggested that H2 activation on Ni0 is not the critical step for a high catalytic activity at such conditions, pointing out on the important role of ceria support in the reaction pathway. The proposed explanation was that only the highly uncoordinated Ni atoms at the metal-support interface, whose number strongly depends on the average particle size, remain available to hydrogenate CO2 activated on the nearby ceria sites. Ratchahat et al. [122] also prepared structured Ni/CeO2 catalysts with different configurations: plain, segmented, stacked and multi-stacked. When compared to the plain configuration, the stacked and segmented ones provided an enhancement of convective heat and mass transfer due to the random flow-channels and the addition of a gap distance in the case of the stacked, which contributed to the improvement of the methanation performances at low temperature and increased the contact time. Moreover, the multi-stacked catalyst was tested under industrial-like conditions (high feed rate and pure feed gas component), being the moderate hot spot observed responsible for a re-boosting of the conversion. Furthermore, this catalyst presented high stability (<0.6% CO2 conversion loss after 76 h under reaction conditions), being its shape and surface morphology maintained. Cárdenas-Arenas et al. [123] also reported Ni/CeO2 catalysts prepared by different Ni-incorporation method in order to modify the Ni-Ce interaction. Authors synthesized three dimensionally ordered macroporous (3DOM) structures by co-impregnation, successive impregnation of Ni and Ce precursors or Ni impregnation after CeO2 3DOM synthesis. Besides, materials with uncontrolled structures were also prepared by co-precipitation and Ni impregnation. Authors identified two types of active sites: NiO–CeO2 interface and reduced Ni0 particles efficient for CO2 and H2 dissociation, respectively. The optimal sites proportion required to obtain the maximum conversion was 25% Ni0 for H2 dissociation and 75% NiO–CeO2 for CO2 dissociation, achieved by the Ni-uncontrolled structured catalyst prepared by impregnation. Moreover, pulse experiments with isotopic CO2 confirmed that this catalyst was more effective in hydrogenating the surface carbon species.
Sharma et al. [124] studied Ni, Co, Pd and Ru doped CeO2 catalysts prepared by combustion method. CexRu1-xO2 presented the best catalytic performances while Ni, Co and Pd induced low CH4 selectivity, with Ni and Co catalyzing both methanation and RWGS reaction and Pd producing only CO. Furthermore, Vita et al. [125] studied Ni/GDC (gadolinium-doped ceria) catalysts both in powder and monolith forms. For the powder systems, the activity increased with Ni loading, which was explained by the more favourable metal-support interaction, the increased number of moderate-basic sites and the presence of surface oxygen vacancies. Powdered catalysts presented also lower CH4 productivity than a high loaded monolith (0.5 g cm−3), due to the high surface-to-volume ratio, good interphase mass transfer and low pressure drops of the monolithic system.
Ocampo et al. [126,127] prepared CexZr1-xO2 (CZ) mixed oxides as supports for 5–15 wt% Ni-based catalysts by sol gel method. CZ fluorite structure was maintained after Ni incorporation but the replacement of some Zr4+ by Ni2+ cations led to a reduction in the lattice parameter of the structure, especially in the 10 and 15 wt% Ni samples. Even if all samples presented great performances and stability, 10% Ni–CZ catalyst reported the best results due to the high oxygen storage capacity of CZ and its ability to effectively disperse Ni particles. Additionally, the incorporation of nickel cations into the CZ structure and the higher dispersion of NiO at its surface were found responsible for improving the redox properties of the materials, able to hinder sintering processes. A similar strategy was followed by Ashok et al. [128], who prepared a series of nickel catalysts supported on CexZr1-xO2 (CZ) by ammonia evaporation, impregnation and deposition-precipitation methods. Among all, the Ni/CZ catalyst prepared via ammonium evaporation method led to the best performances and stability. These positive results were explained by the fact that this catalyst was activated at low temperatures, and the incorporation of some Ni cations in the ceria structure led to the creation of oxygen vacancies and to better reducibility of Ni species in this sample.

4.2. Bi- and Trimetallic CeO2-Supported Catalysts

Le et al. [129] studied Ni–Na/SiO2 and Ni–Na/CeO2 catalysts prepared by impregnation in order to compare the effect of the support nature and evaluate Na loading effects. In both cases, increasing Na contents led to lower performances which is ascribed to a reduction in the textural properties and in the CO2 adsorption capacity while higher performances were obtained by CeO2-supported materials. In addition, Sun et al. [131] developed a novel ICCU (integrated carbon capture and utilization) process for carbon capture and methane production, using dual function materials composed by a mechanical mixture of MgO and Ru/CeO2 catalysts with increasing Ru loadings. This ICCU process allows the simultaneous regeneration of the sorbent and the conversion of the captured CO2 to CH4 in a single reactor at intermediate temperature (300 °C). Authors found that MgO+10Ru/CeO2 showed the best performance in the 1st cycle of ICCU due to a higher number of oxygen vacancies. However, after 10 cycles, MgO+5Ru/CeO2 exhibited the highest catalytic activity because of the more favorable metal-support interactions established in this sample. Finally, Pastor-Pérez et al. [130] prepared Ni–Co/CeO2–ZrO2 and Ni–Mn/CeO2–ZrO2 catalysts (~15 wt% Ni and 3.5 wt% Co/Mn) and evaluated the impact of methane incorporation in the feed. Authors verified a negative effect of Mn on the activity, which attributed to a stronger than desired CO2 adsorption over these sites, limiting the subsequent hydrogenation of the reactant towards methane. On the contrary, Co favored the performances through the improvement of the catalyst’s electronic properties and the inhibition of Ni sintering. They also verified that incorporating methane in the feed led to a slight enhancement of the catalytic performances, which they ascribed to low temperature reforming reactions taking place simultaneously with methanation.

4.3. Mechanistic Aspects

In terms of mechanisms, Konishcheva et al. [132] proposed that CO2 methanation proceeds over Ni/CeO2 catalysts through the activation of CO2 on CeO2 surface and the stepwise hydrogenation to methane through hydrocarbonate and formate intermediates by the hydrogen spilled over from Ni. Authors considered the route where CO acts as a reaction intermediate negligible. Cárdenas-Arenas et al. [119] also reported the presence of two independent active sites for CO2 methanation over Ni/CeO2 (Figure 8): NiO–CeO2 interface for CO2 activation and Ni0 particles for H2 dissociation. Even if they identified water desorption as the slowest step of the mechanism, the co-existence of two types of actives sites and the high oxygen mobility of Ni/CeO2 catalysts hindered water inhibitory effect.
Aldana et al. [126] studied Ni–CexZr1-xO2 catalysts and proposed a bifunctional mechanism without CO as intermediate species (Figure 9a). Actually, they observed that H2 molecules were activated and dissociated into H on Ni0 sites while CO2 was adsorbed on the Ce–Zr support oxygen vacancies forming monodentate carbonates. These monodentate carbonates were then sequentially hydrogenated producing formates, formaldehydes, methoxy species and, finally, methane. Additionally, authors proposed a pathway for CO formation and verified its inhibitory role in the reaction (Figure 9b).

5. ZrO2-Based Catalysts

In terms of ZrO2 supported catalysts, the main results found in the literature can be seen in Table 5 while the main conclusions of the published works will be discussed below.

5.1. Monometallic ZrO2-Supported Catalysts

Da Silva et al. [133] studied Ni/ZrO2 catalysts and compared them to Ni/SiO2 samples, both prepared by impregnation and containing 10 wt% Ni. The sample supported on zirconia was more active, which authors explained by the considerably weaker interaction of CO2 with the SiO2 sample and the higher Ni dispersion found in the ZrO2 catalyst. In addition, Lu et al. [134] prepared mesoporous zirconia-modified clays as supports for Ni catalysts by one-pot method (hydrothermal treatment of the mixture of the clay suspension and the ZrO(NO3)2 solution). The bimodal pore structure of the zirconia-modified clays was reported as beneficial for the dispersion of Ni. As concluded by previous authors, ZrO2 favored the dispersion of the Ni species, preventing sintering during the reaction and reducing carbon deposition. The catalyst with 20 wt% of zirconia led to the best performances while the most stable sample was the 15%Ni/ZrO2. Jia et al. [135] studied also 9 wt% Ni-supported ZrO2 catalysts and reported the beneficial effect of carrying out the decomposition of the precursor salts after impregnation (calcination) using dielectric barrier discharge (DBD) plasma instead of a conventional thermal treatment at 500 °C (from 30 to 70% of CO2 conversion at 300 °C when performing thermal and DBD plasma calcination, respectively). The positive effect of plasma was related to the remarkable enhancement of the metallic dispersion when using this non-conventional treatment. Additionally, Zhao et al. [136] prepared 10, 15 and 25 wt% Ni/ZrO2 catalysts by combustion method using several combustion mediums such as urea, glycerol, glycol, ethanol and n-propanol. The highest performances were reported for the sample synthesized using urea, being the results attributed to the favored reducibility of the Ni species, the better dispersion and the greater CO2 adsorption capacity.
Furthermore, Li et al. [137] reported Co-based catalysts supported on ZrO2, Al2O3, SiO2, SiC, TiO2 and activated carbon (AC), all containing 10 wt% of Co. Their catalytic tests were performed at 30 atm of pressure, contrary to what was done in the majority of works reported for CO2 methanation. ZrO2 was the best performing Co-based support, with the order as follows: 10%Ni/ZrO2 > 10%Ni/SiO2 > 10%Ni/Al2O3 ≈ 10%Ni/SiC > 10%Ni/AC > 10%Ni/TiO2. The zirconia support contributed to the better reducibility and dispersion of Ni species. Besides, the presence of oxygen vacancies on the ZrO2 structure prevented the formation of carbon deposits during the tests. The same authors [138] studied later the effect of the impregnation solvent for preparing 2 wt% Co catalysts supported over ZrO2. Among all solvents, citric acid-assisted impregnation led to the best catalytic results. The beneficial effect of citric acid arises from the enhancement of the Co particles dispersion and its interaction with ZrO2 due to the formation of cobalt citrate complex as precursor. This intensified interaction also provided oxygen vacancies and enhanced nickel reducibility, improving CO2 adsorption capacity and, consequently, the obtained performances.
Recently, Nagase et al. [139] reported ~3 wt% Ru/ZrO2 catalysts (using amorphous and crystalline ZrO2) prepared by selective deposition method and compared them with an equivalent Ru/SiO2 sample. Ruthenium particles with average sizes below 10 nm were obtained in all cases with amorphous ZrO2 promoting the activity and selectivity. The effects of the preparation method for Ru/amorphous ZrO2 was studied, by comparing the use of a NaOH or NH3 in the selective deposition method with a conventionally impregnated sample. Impregnated and NH3 samples led to the similar results, but lower than those exhibited by the catalyst prepared using NaOH. The higher activity of the later was related not only to the promoted dispersion, also found when using NH3, but especially to the presence of residual Na species in the catalyst able to activate the intermediate formate species during the reaction.

5.2. Bimetallic ZrO2-Supported Catalysts

Takano et al. [140] prepared 50 wt% Ni/Sm–ZrO2 by mechanical mixture followed by calcination. The best results were obtained for the Zr/Sm ratio of 5 and after calcination at 650 or 800 °C. Sm3+ ions were substituted in the ZrO2 lattice, what promoted the presence of oxygen vacancies, favored the interaction of CO2 weakening the C–O bonds and, consequently, enhanced the CO2 activation.
The same authors [141] prepared Ni supported Yttrium–Zirconia catalysts by co-impregnation and with several Y3+ concentrations. Ni/Y–ZrO2 catalysts showed higher catalytic activity than the Ni/ZrO2 sample, being the performances dependent upon the Y3+ concentration, with the best results obtained for the catalyst with a Y/(Zr+Y) molar ratio of 0.333. Kesavan et al. [142] also studied Ni-based Y-stabilized ZrO2 (YSZ) catalysts with 10 wt% Ni prepared by different methods: wetness impregnation, wetness impregnation with EDTA, electroless plating, mechanical mixing with commercial NiO nano-powder and mechanical mixing with nano-powder transformed in micro-powder NiO. The most efficient and stable catalyst was the one prepared by wet impregnation with Ni(EDTA)2− complex due to the smaller and highly dispersed Ni0 particles formed. Additionally, Kosaka et al. [143] studied Ni-based tubular catalysts again supported over Y–stabilized ZrO2 with different Ni loadings. Even under high gas hourly space velocity (GHSV) conditions, catalysts with higher Ni content showed better catalytic activity, which was also favored by their tubular structure that provided a higher CH4 yield when compared to the same powder catalysts. Authors studied also the effect of Ni loading on the temperature profiles and realized that the high CO2 methanation performances arise from the interaction between heat generation and acceleration of the reaction rate in the catalysts. Through a numerical analysis based on reaction kinetics, heat transfer and fluid dynamics, authors deduced that both high CH4 yield and prevention of hot spots formation can be achieved by properly arranging catalysts with different activities in the reactor.
Ren et al. [144] studied Ni/ZrO2 catalysts impregnated with a second metal (Fe, Co and Cu). Fe was the metal leading to the best performances, which resulted from the improved dispersion and reducibility of Ni species as well as to the partial reduction of ZrO2 in the Ni–Fe/ZrO2 sample. Lu et al. [145] also studied the effects of adding Fe, Co, Ce and La to Ni based ZrO2 modified clays by impregnation. They observed that by adding 1 wt% of promoter to the 15%Ni/ZrO2 catalysts, Ni species dispersion and reducibility and catalysts’ thermal stability were enhanced. Even if the performances were quite similar for all the promoted samples, Fe led to the best results attributed to synergistic effects between Fe, ZrO2 and Ni species. Dumrongbunditkul et al. [146] prepared Co–Cu–ZrO2 catalysts by co-precipitation method and tested them at 30 atm. They also prepared monometallic samples with better results obtained for the Co–ZrO2 than for the Cu–ZrO2 samples. Additionally, the catalyst with Co:Cu:Zr = 20:40:40 molar ratio led to the highest performances due to its high surface area and the more homogeneous dispersion of the mixed metal oxides.
Tan et al. [147] prepared a series of MgO-modified Ni/ZrO2 catalysts with 6 wt% Ni and Mg/Ni molar ratios ranging from 1/6 to 1/1 using citrate-complexing impregnation method. Among all the synthesized materials, the catalyst with Mg/Ni molar ratio of 1/4 showed the best catalytic performances. In accordance with Li et al. [138], the use of citric acid was considered as responsible for a reduction in the Ni particles size. Furthermore, the addition of MgO stabilizes the highly dispersed Ni nanoparticles and has a confinement effect, limiting the growth of Ni and improving the resistance towards sintering and carbon deposition. Takano et al. [148] also studied tetragonal ZrO2 doped with Ca2+ and Ni2+ ions. The incorporation of Ca2+ and Ni2+ ions to the zirconia led to the formation of oxygen vacancies in the ZrO2 lattice and favored the performances.
In order to compare the different promoters reported for Ni-based ZrO2 catalysts prepared by impregnation and calcined under thermal conditions, CH4 production rates were determined for a Ni/ZrO2 catalyst [135] and for bimetallic materials containing Y [141], Fe [144], Co [144], Cu [144], Mg [147] and Ca [148] (Figure 10). As seen, the main effect of the promoters is verified at lower reaction temperatures while Mg, Co or Fe are suitable candidates for developing further optimization studies.

5.3. Mechanistic Aspects

Takano et al. [140,141,148] proposed a mechanism for Ni/ZrO2 catalysts where CO2 is adsorbed as carbonate species on zirconia oxygen vacancy sites rather than CO while hydrogen is activated on Ni particles. Then, H atoms must be supplied from Ni to ZrO2 sites to reduce the adsorbed carbonate species progressively into formates, formaldehydes, methoxy species and, finally, methane.

6. MgO and Hydrotalcite-Based Catalysts

Besides all the catalysts already presented in this work and containing Mg as a promoter, in this chapter catalysts supported on MgO and hydrotalcite(HT)-derived materials will be summarized, with their main results shown in Table 6

6.1. Monometallic MgO and Hydrotalcite-Derived Catalysts

Loder et al. [149] prepared a bifunctional Ni/MgO catalyst for CO2 methanation. They tested the influence of Ni loading (11, 17, 20 and 27 wt% of Ni) and the matrix composition (MgO pure and MgO/CaO). Results confirmed that MgO is an active compound and the CaO did not participate actively in the reaction, since the thermal stability of CaCO3 is much higher than MgCO3. In this way, Ni/MgO catalyst showed better results than those obtained for Ni/MgO–CaO. The catalytic activity increased linearly with the Ni loading since a remarkable number of active sites were found even at high Ni loadings (27 wt%). Based on the reaction kinetics, authors also modelled a Langmuir–Hinshelwood based rate law taking the bifunctional catalytic action of the catalyst into account.
Wierzbicki et al. [150] studied the effect of the Ni loading (10 to 43 wt%) on Ni based hydrotalcite-derived materials with constant M(II)/M(III) (valence II to valence III metals) molar ratio, prepared by pH controlled co-precipitation and the decomposition of the precursor hydrotalcites carried out by thermal treatment. After this treatment, periclase-like structured (homogeneous mixed oxide) materials were formed being Ni present as NiO. Increasing Ni contents weakened the interaction between Ni and the hydrotalcite matrix, improving its reducibility and increasing the number of active sites. This resulted in an enhancement of the performances. Additionally, Bette et al. [151] prepared a ~59 wt% Ni catalyst supported on a (Mg,Al)Ox mixed oxide derived from a (Ni,Mg,Al)–hydrotalcite-like (HT) precursor. The precursor HT structure was synthesized by pH-controlled co-precipitation and the reduction at 900 °C led to metallic Ni particles supported on a (Mg,Al)Ox matrix. Authors reported that the hydrotalcite-like derived material provided anti-sintering and stability properties to the catalysts, what justified the good performances revealed by the supported materials. Finally, Abate et al. [152] prepared Ni–Al–hydrotalcite materials and compared them to a Ni/Al2O3 commercial sample with 80 wt% Ni. Ni content was fixed with an Al/(Al + Ni) = 0.25–0.27 corresponding to a theoretical loading of 75–80 wt% NiO. Better results were obtained for the HT derived sample due to the favored catalysts reducibility and the higher metal surface area and metal dispersion.

6.2. Bi- and Trimetallic MgO and Hydrotalcite-Derived Catalysts

Yan et al. [153] prepared Ni–Mg mixed oxides promoted with Co, Mo, Mn, Fe, Cu and W. Authors also prepared MgO, NiO, NiaMgOx (a = 0.5, 0.8, 1.0, 1.2) and Ni/MgOImpregnated samples. NiO and MgO catalysts revealed quite poor activity, Ni–Mg mixed oxides led to considerably higher levels of CO2 conversion and CH4 selectivity and the Ni/MgOImpregnated sample presented intermediate results between the NiO and the mixed oxides. Being Ni1.0MgOx the best mixed oxide, samples containing transition metals were synthesized using this optimized composition. The performances varied as follows: W > Co > No promoter > Mo > Mn > Fe > Cu. Then, as the W-promoted sample presented the best results, authors optimized the W/Ni molar ratio with the best performances obtained for the 1/1 value. They considered that the remarkable results reported by the NiWMgOx catalyst were related to the great anti-CO-poison ability, strong resistance against coke formation and sintering and the higher amount of CO2 adsorption sites due to the contribution of W species. Varun et al. [154] studied NiO–MgO and M/NiO–MgO (M = 2 wt% Co, Cu or Fe) catalysts and evaluated the composites preparation method and the promoter nature. They reported three NiO–MgO materials synthesized by solution combustion, sonochemical and co-precipitation methods and obtained the best results when using the second due to the highly dispersed particles formed. Regarding the promoter nature (Co, Cu or Fe; incorporated by impregnation), Co led to the best performances, explained by the lowest activation energy determined for this catalyst due to Co reducing nature.
Ho et al. [155] reported hydrotalcite-derived ~60 wt% Ni catalysts promoted with 9.5 wt% La, 6.3 wt% Y or 9.5 wt% Ce. The best catalytic performances were exhibited by the Ni-La sample (Ni–La > Ni–Ce ≈ Ni–Y > Ni) and were attributed to La2O3 species ability to simultaneously promote basicity and Ni0 dispersion. Authors concluded that an equilibrium between Ni0 particle size and basicity is fundamental for obtaining the best CO2 methanation performances. Furthermore, Wierzbicki et al. [156] studied the promoter effect of La on the performances of Ni–HT derived materials, keeping the M(III)/(M(III)+M(II)) molar ratio equal to 0.25. Authors verified that the addition of lanthanum resulted in the formation of a separate phase, not integrated in the hydrotalcite-derived structure. Indeed, the incorporation of 2 wt% of lanthanum favored the catalytic performances due to the enhanced basicity, reducibility and dispersion of Ni species. Lately, the same authors [157] studied different preparation methods for La introduction (e.g., co-precipitation, impregnation and ion exchange using La–EDTA) in order to investigate its influence on the catalytic performances. La affected catalysts CO2 adsorption capacity in terms of medium strength basic sites, which were strongly influenced by the used preparation method. Among all, ion exchange was the most efficient method to increase the amount of medium strength basic sites, leading to the best catalytic performances. Zhang et al. [158] also studied La-promoted Ni/Mg–Al catalysts with different La contents (0 to 8 wt%) and prepared by two different methods: urea hydrolysis and co-precipitation. La-promoted catalysts showed an enhanced catalytic activity, with the one containing 5 wt% of La presenting the best Ni dispersion. Besides, La increased significantly the amount of moderate basic sites that enhanced CO2 adsorption capacity. Furthermore, urea hydrolysis was a more efficient preparation method than co-precipitation, as it promoted Ni dispersion and CO2 adsorption capacity, leading to better catalytic performances.
Wang et al. [159] studied a Ni–Al–hydrotalcite-derived catalyst modified with Fe or Mg and prepared by co-precipitation. Authors prepared samples with different (Fe or Mg)/Al molar ratios (Fe/Al = 0.05 and 0.25, Mg/Al = 0.1 and 1). The performances of Fe catalysts were the same for both ratios while, in Mg samples, the lower the Mg content, the better the results. They suggested that both Fe and Mg favored Ni dispersion while Mg enhanced also the basicity, helping in CO2 adsorption and activation, and Fe improved the reducibility and changed the pore distribution. Mebrathu et al. [162] also studied Ni–Fe bimetallic catalysts derived from hydrotalcites and prepared by co-precipitation, using variable Fe/Ni ratios (0.1 to 1.5, corresponding to 12 wt% Ni and 1.2 to 18 wt% Fe). Bimetallic catalysts showed higher activities and stabilities than monometallic ones, with the sample presenting the lowest Fe content (1.2 wt% Fe, Fe/Ni = 0.1) being the most outstanding. The incorporation of Fe led to weaker Ni-support interactions, promoting nickel reducibility. Additionally, an improvement of the metallic dispersion (with lower Fe loadings) and basicity was associated with Fe incorporation. In accordance with other studies already discussed in this work, the basicity of the support and the metal particle size were found as key catalysts properties. Later, the same authors [160] studied the deactivation mechanism for the best catalyst reported in the previous work (Fe/Ni = 0.1) and for the monometallic Ni sample. They observed the formation of nickel hydroxide by the reaction of Ni species with the water from methanation during the long-term tests (30–40 h). The formation of these species decreased the number of available Ni0 sizes for the reaction, favored sintering processes and facilitated the formation of hardly reducible nickel aluminates. Authors concluded that water partial pressure was the main factor influencing the deactivation rate of the Ni catalyst, while the higher resistance of Ni–Fe material was due to Fe ability to limit the formation of nickel hydroxide.
Finally, He et al. [161] prepared Ni–Al2O3–HT catalysts as well as reference samples with Ni incorporated by impregnation. Ni–Al2O3–HT catalyst presented highly dispersed Ni particles along with strong basic sites, what led to considerably better results than the impregnated samples. Additionally, one sample containing K (K–Ni–Al2O3–HT) was synthesized contributing to results slightly higher than those of the un-promoted catalyst, due to the extra strong basic sites provided by K. Therefore, authors proposed that the combination of small metallic Ni particles with strong basic sites over Al2O3 could effectively promote the conversion of CO2 into CH4. This study contrasts with other works where medium strength basic sites are considered as the most favorable for CO2 methanation reaction, as the strongest ones lead to inactive carbonate phases.

7. Carbon-Based Catalysts

Few studies have focused on the use of carbon-based catalysts for this reaction. In summary, their main results (Table 7) and findings can be found in the following sections.

7.1. Monometallic Carbon-Supported Catalysts

Jiménez et al. [163] prepared Ru impregnated carbon nanofibers (CNFs) with different natures (orientation of graphite planes): platelet, fishbone and ribbon. The catalytic performances were not affected by the nature of the carbon nanofibers used as support. In addition, Li et al. [164] studied Co-based porous carbon (PC) catalysts with controlled crystal morphology and size by using ZIF–67 metal organic framework (MOF) as template and with the aid of surfactants. After carbonization, the obtained samples inherited the original morphology and size of ZIF–67 crystals with a distorted surface. Catalytic tests were carried out at 30 atm and Co nanoparticles inside the carbon matrix ranged between 7 and 20 nm were separated by the graphite-like carbon effectively avoiding metal sintering. Consequently, interesting performances were reported for these samples, especially when compared with similar catalysts supported on commercial activated carbon. Finally, Gödde et al. [165] prepared Ni nanoparticles supported on nitrogen functionalized carbon nanotubes (CNTs) with different Ni loadings (10 to 50 wt%). The optimum loading of 30–40 wt% led to high surface area and small Ni particle sizes, ascribed to the efficient anchoring on the N-doped CNTs. Besides, high stability of the catalysts was demonstrated for 100 h time-on-stream.

7.2. Bi- and Trimetallic Carbon-Supported Catalysts

Romero-Sáez et al. [166] reported Ni–Zr-supported CNTs catalysts by both sequential and co-impregnation methods (metal loadings: 5 wt% Ni and 20 wt% Zr). Authors obtained better results after sequential impregnations, since co-impregnation method resulted in the formation of a core-shell structure where NiO particles were surrounded by ZrO2, limiting the access of the reactants to Ni active sites during the catalytic tests. On the contrary, sequential impregnation favored Ni reducibility and CO2 adsorption capacity. Wang et al. [167] prepared Ni-based catalysts supported on multi-walled CNTs and promoted with Ce by ultrasonic-assisted co-impregnation (12 wt% of Ni and Ce contents of 0, 1.5, 3, 4.5 and 6 wt%). Additionally, authors prepared samples supported on γ–Al2O3 for comparison purposes. They verified, in accordance with other studies previously discussed in this work, that Ce enhanced the Ni0 dispersion, promoted the reduction of metal oxides and favored the CO2 activation. Meanwhile, the confinement effect of CNTs and the promotion effect of cerium could efficiently prevent active species migration and sintering, as well as restrict carbon deposition. The best results were reported for the 12%Ni–4.5%Ce/CNTs. Le et al. [168] prepared Ni based activated carbons (AC) and verified the effect of adding Ce–Zr oxides as promoters. 7%Ni/AC reference sample was prepared by impregnation while the promoted 7%Ni/Ce0.2Zr0.8O2/AC sample was prepared by depositing the Ce0.2Zr0.8O2 mixed oxide prepared by hydrothermal method over the activated carbon and later impregnating 7 wt% Ni over the Ce0.2Zr0.8O2/AC support. Authors verified that the ceria-zirconia solid solution phase could effectively disperse and stabilize nickel species as well as favor activated carbon CO2 adsorption capacity. Consequently, 7%Ni/Ce0.2Zr0.8O2/AC catalyst exhibited higher activity than 7%Ni/AC sample. Finally, Gaidai et al. [169] synthesized carbon fiber (CF) supported Fe–Co catalysts by uniform infiltration of a metal nitrate solution through the carrier (CF) surface, followed by drying. Authors, who performed several treatments in the carbon fibers, found out that oxygenation increased the amount of oxygen in the surface layer, decreasing active metals reducibility. Carbon fibers pre-reduction increased the amount of oxygen containing groups on the surface through pulling them from the bulk, which was detrimental for the catalytic performances.

7.3. Mechanistic Aspects

Lapidus et al. [170] studied CO2 methanation mechanism over copper and nickel-based carbon catalysts (Ni/C and Cu/C). They only found CO and water as products on Cu/C, suggesting that CO2 was so strongly adsorbed on the catalyst’s surface that there was not enough adsorbed hydrogen to convert the formed CO into CH4. In the case of the Ni/C sample, authors observed that CH4, CO and water were produced, indicating the presence of a sufficient amount of adsorbed hydrogen on the Ni catalyst surface. However, they observed that a longer period of time was needed for detecting CH4 in the reactor effluent, whereas CO appeared from the beginning of the reaction. Thus, they proposed that CO was an indispensable reaction intermediate. Wang et al. [167] studied Ni–Ce/CNTs and proposed a mechanism where Ce, in conjunction with the CNTs support, donated electrons and facilitated their transfer from the support to the metallic nickel and the absorbed CO2, accelerating the activation of CO2 molecules and the formation of atomic hydrogen. Then, the activated CO2 dissociated on the surface of Ni formed carbonate species further hydrogenated to methane. They also pointed out that CO could be formed as by-product during this process, without playing a role in CH4 formation.

8. Zeolite-Based Catalysts

Several works have been reported in the last few years dealing with the utilization of zeolite-based materials as catalysts for CO2 methanation, with a recent review recently published on this topic [24]. Consequently, only the main results found in the literature recently are shown in Table 8 in order to allow a proper comparison with the rest of the catalysts presented in the current work.
Starting by an interesting study recently reported by Hussain et al. [171], fibrous silica-mordenite (without any additional active metal) was synthesized by microemulsion and tested under CO2 methanation conditions, comparing its properties and performances with those obtained by a commercial MOR zeolite. Both samples presented CH4 yields below 50% in all the temperature range studied (250–500 °C), but fibrous SiO2–MOR exhibited better performances (conversion, selectivity and stability for 50 h under reaction conditions), especially at higher temperatures. This behavior was explained by the higher basicity, textural properties and oxygen vacancies found in the sample, able to promote the activity towards CO2 methanation and suppress coke deposition, boosting stability with time-on-stream.

8.1. Monometallic Zeolite-Supported Catalysts

Franken et al. [172] recently reported Fe/13X (intact and collapsed zeolite structure) catalysts with increasing metal loadings (1, 5 and 10 wt%) and tested under atmospheric and elevated pressures (5–15 bar). Authors verified that the preservation of the zeolite structure allows obtaining higher metallic dispersion and, hence, more favorable performances. In addition, better performances were obtained for the sample presenting the lowest Fe loading, where highly dispersed metallic particles were formed. Generally, authors verified that higher pressures favored the catalytic performances as expected taking into account the characteristics of the Sabatier reaction. However, the 1 wt% Fe/13X catalyst exhibited lower reaction rates at 15 than at 10 bar, which was ascribed to limitations in terms of intermediate species desorption or modifications on Fe species nature at elevated pressure.
Kitamura Bando et al. [173] studied Rh–Y zeolite and Rh/SiO2 catalysts and performed tests at 30 atm. Authors reported better performances for the Rh–Y but it possessed an intense deactivation after 100 min due to the accumulation of H2O (arising from the reaction) inside the cavities of the zeolite. Later, Graça et al. [184] reported Ni/USY catalysts and evaluated the effects of the Ni incorporation method (ion exchange or impregnation) and Ni loading, with 15 wt% found to be the best compromise. The same authors [188,189] reported that the chosen calcination and reduction temperatures as well as the impregnation solvent can be crucial in these materials. Bacariza et al. evaluated the effects of the compensating cation nature [190], Si/Al ratio [174] and zeolite framework type [175] in the performances of Ni-based zeolites. Among all, larger cations (e.g., Cs+), higher Si/Al ratios and USY zeolite led to the best results. In these studies, an additional relevant property of the support was proposed: the hydrophobicity. In fact, several zeolite samples series presenting a systematic and gradual enhancement of hydrophobicity exhibited improved methane yields, which was attributed to a reduced negative effect of the competitive adsorption of produced water over the same sites adsorbing CO2 [118].
Goodarzi et al. [176] prepared 5 wt% Ni catalysts supported over silicalite-1 (S1) and desilicated silicalite–1 (d–S1) zeolites. Authors observed an improvement of the catalytic performances when using as support d–S1 support, mainly due to the confinement effect induced in Ni0 particles. Chen et al. [178] reported recently a novel Ni–ZSM–5 catalyst synthesized by hydrothermal method and using as Si source a Ni/SiO2 catalyst prepared by impregnation. They compared the properties and performances of this material with conventional Ni/SiO2 and Ni/ZSM–5 catalysts (all with ~10 wt%). Ni–ZSM–5 was strongly stable exhibiting an excellent stability over 40 h time on stream, attributed to an embedment structure able to hinder sintering, while the others reflected severe deactivation after 5 h. Guo et al. [177] compared a Ni/ZSM–5 catalyst with equivalent Ni samples supported over SBA–15, MCM–41, Al2O3 and SiO2. Among them, the Ni/ZSM–5 sample presented the best performances. As already found by Goodarzi et al. [176], desilication provided an encapsulation effect for Ni particles, being CH4 yields significantly improved.
Czuma et al. [180] synthetized an X zeolite from waste fly ashes to be used as support in the preparation of Ni catalysts for CO2 methanation. Despite the partial micropores blockage due to Ni incorporation, the catalyst still presented an adequate surface area, showing a predominant presence of strong basic sites promoting the formation of very stable CO2 species and, consequently, low methane yields. In spite of the CO2 conversions obtained (~50%) being slightly lower than those obtained when using commercial zeolites, the use of fly ash zeolites offers the opportunity to use waste from energy sector, which is a positive economic and ecological aspect. Wei et al. [181] prepared 5 wt% Ni supported 5A and 13X zeolites and studied the effects of the Ni precursor salt (nitrate, citrate or acetate) and calcination temperature (300, 350, 400 and 450 °C) on the properties and performances. While no remarkable enhancements were found in the 5A-based samples, the activity, selectivity and stability were highly improved when using nickel citrate as salt and 350–400 °C as calcination temperature for 5Ni/13X (320 °C: 79% of CO2 conversion, 100% of CH4 selectivity and stability after 200 h of time-on-stream). They attributed the higher performances of 13X samples to the location of Ni species in the zeolite channels, allowing a confinement effect, while the use of citrate and a mild calcination temperature improved Ni species reducibility and metallic dispersion. Recently, Da Costa-Serra et al. [182] reported a series of ITQ–2 (prepared by delamination using ITQ–1 and MCM–22 as starting materials) and ZSM–5 zeolites impregnated with 5 wt% Ni. They evaluated the effect of the zeolite framework type and the Si/Al ratio and elaborated an enhancement of the performances when decreasing the Al content for both structures due to the improvement of the materials’ hydrophobic properties. The best results were revealed by the sample supported over the ITQ–2 with Si/Al = ∞, a purely silica zeolite that led the formation of highly dispersed Ni0 nanoparticles (<3 nm).
As shown, many works have dealt with the preparation of Ni-supported zeolites. Based on a comparison of the zeolite type (USY [174], MOR [175], ZSM–5 [175], BEA [179], 13X [181], 5A [181] or ITQ–2 [182]—Figure 11), one can observe that USY, MOR, ZSM–5 and ITQ–2 present generally better performances, while 5A is the zeolite leading to the lowest performances.
Monometallic zeolites have been also applied in sorption-enhanced methanation. Indeed, Delmelle et al. [191] prepared 5Ni/13X and 5Ni/5A catalysts by impregnation obtaining similar CH4 yields. Borgschulte et al. [192] studied a Ni–5A catalyst prepared by ion exchange with <6 wt% Ni, obtaining a CH4 yield of ~100% when using a H2:CO2 ratio of 8:1. Walspurger et al. [193] mixed a Ni-based commercial catalyst with a hydrophilic 4A zeolite and, as suggested by the previous studies from Delmelle et al. [191] and Borgschulte et al. [192], attributed the beneficial effects of using an hydrophilic zeolite to the displacement of the reaction equilibrium to the formation of CH4. In this context, Isah et al. [178] reported recently a beneficial effect when adding a small fraction (2 wt%) of zeolite over a 10 wt% Ni/Al2O3 catalyst, that improved the textural properties without affecting the metallic surface area available.

8.2. Bimetallic Zeolite-Supported Catalysts

Regarding bimetallic catalysts, several studies reported the beneficial effects of adding promoters such as Mg (Bacariza et al. [183]), Ce (Graça et al. [184], Bacariza et al. [185]) or La (Quindimil et al. [179]) over Ni/Zeolite catalysts due to the general enhancement of the CO2 adsorption capacity and Ni0 dispersion. In addition, Boix et al. [186] reported interesting results when using Pt-Co/MOR catalysts due to the formation of PtCoxOy active species. Recently, Wei et al. [187] reported bimetallic Ni–Ru catalysts supported over 13X and 5A zeolites. Authors kept constant the total metal content (5 wt%) and varied Ru and Ni loadings. They also prepared two monometallic reference samples containing 5 and 2.5 wt% Ni and Ru, respectively. Despite the effects of Ru on the selectivity to CH4, no remarkable enhancement of the catalytic performances was generally achieved when preparing bimetallic Ni–Ru/Zeolite catalysts, being 5Ni/13X and 5Ni/5A the catalysts exhibiting the highest CO2 conversions. In terms of mechanism, Westermann et al. proposed some mechanism insights for Ni/USY [194] and Ce/Ni/USY [195] zeolites prepared by impregnation. In both cases formate species were reported as key intermediates.

9. Other Types of Supported Catalysts

Finally, some catalysts not included in the previous chapters, found in the literature, are presented in Table 9 and discussed below.
In this way, Lu et al. [196] studied the promotion effect of VOx (3, 5 and 8 wt% of V2O5) on 20 wt% Ni catalysts supported on raw (RB) and modified bentonite (B) and prepared by impregnation method. The presence of more dispersed Ni0 particles in the modified bentonite explained the better results obtained for this sample. After adding VOx, the catalytic activity was further improved due to the higher H2 uptakes, the increased Ni0 dispersion and the superior anti-coking and anti-sintering properties of the promoted samples.
Additionally, Le et al. [197] reported Ni-based catalysts using SiC and SiO2 as supports prepared by wet impregnation and deposition-precipitation (DP) methods. Authors observed an enhancement of the metal-support interactions and metallic dispersion when using DP method, which induced a significant effect in the improvement of the CO/CO2 methanation performances over the SiC supported catalyst. SiC turned out to be the most behaving support due to its superior thermal conductivity. Additionally, authors reported the beneficial effect of including Mn in the formulation, due to the favored metallic dispersion and CO2 activation capacity. In addition, Zhi et al. [198] studied Ni and 15%Ni–5%La catalysts supported on SiC and prepared by impregnation. La2O3 promoted the dispersion of Ni species on the SiC support simultaneously increasing the number of active sites on the catalyst surface and leading to better performances than the one reported for the Ni/SiC catalyst.
Kirchner et al. [199] studied different Fe2O3 catalysts and reported the best activity for nanosized γ–Fe2O3. Authors associated the efficiency of this catalyst with reactive surface carbon species. Baysal and Kureti [200] also reported the use of commercial α–Fe2O3 promoted with alkali (Li, Na, K, Rb and Cs), alkali earth (Mg, Ca and Ba), transition (Al, Mn, Cu and Mo) or rare earth (La, Ce and Sm) metals for CO2 methanation at variable pressures. Among all, Mg (2 wt%) was identified as the most promising promoter. Indeed, MgO species, highly dispersed over iron oxide, were responsible for favorable changes in the iron phase composition and in the intermediate carbon species (e.g., Fe carbides).
Wang et al. [201] prepared hierarchically porous network-like Ni/3D draped Co3O4 catalysts, where Ni clusters and nanodeposits were uniformly dispersed across the high surface area of the Co3O4 support. The pre-reduction treatment was found to induce oxygen vacancies formation, improving Ni-Co interaction, enhancing the catalytic performances and lowering the activation energy. Wang et al. [202] studied two perovskite LaCoO3 based catalysts: PdO–LaCoO3 with an encapsulated structure and PdO/LaCoO3 with a surface dispersing structure. The best catalytic performances were achieved by the encapsulated structure, which was attributed to the strong interaction of PdO NPs and LaCoO3 and also to the H2 spillover effect of Pd. Thus, the decomposition of perovskite structure was significantly promoted, benefiting the extraction of metallic Co NPs from LaCoO3 to prepare a well-dispersed catalyst. Besides, the spent catalysts revealed that well-dispersed catalytically active Co2C was in situ formed accompanied with perovskite structure decomposition during the catalytic reaction. LaCoO3 and PdO/LaCoO3 catalysts showed a weak interaction and the perovskite structure remained with few Co2C, explaining the significant difference of catalytic performance.
Branco and Ferreira [210,211,212,213] reported a series of Co and Ni-based bimetallic materials containing f block elements such as Th, U, La, Ce, Sm, Dy or Yb. Starting by Ni–Th and Ni–U catalysts [210], authors studied two preparation methods: controlled oxidation under dry air, using intermetallic binary compounds as precursors, and a modified sol-gel. The intermetallic route induced more favorable interactions between Ni and f block elements, resulting in better performances. Additionally, Th led to better results than U. Furthermore, authors used electrospinning followed by a controlled oxidation treatment for synthesizing Ni–La, Ni–Ce, Ni–Sm, Ni–Dy and Ni–Yb catalysts in the form of nanofibers and nanoparticles [211]. While the morphology of the catalysts did not affect the performances, results were found to increase along the lanthanide series, with Dy and Ce leading to the best performances. Authors also compared Ni–La, Fe–Dy and Co–Sm catalytic systems prepared by electrospinning followed by oxidation [212], being the first catalyst the most significantly active towards CO2 methanation (Fe–Dy and Co–Sm presented >95% selectivity towards CO). Finally, Branco et al. [213] evaluated the effect of the calcination temperature on Co–Lanthanide (La, Ce, Sm, Gd, Dy and Yb) aerogels prepared by epoxide addition method. They observed that higher temperatures promote the formation of perovskite structures (except for Ce and Yb), enhancing catalysts basicity and activity, but presenting lower selectivity towards methane. Among all, the use of Ce as promoter and a calcination temperature of 900 °C led to the best results.
Zhen et al. [203] prepared a series of Ni/MOF–5 with increasing Ni loadings and obtained the highest catalytic activity with 10Ni/MOF–5 catalyst. This material presented high specific surface area and large pore volume that provided highly uniform dispersed Ni particles in the framework. Besides, this catalyst showed high stability and almost no deactivation during a long-term stability test (100 h).
Zhou et al. [204] studied Ni/TiO2 catalysts prepared by impregnation followed by DBD plasma decomposition. Authors found out the structure-sensitivity of the reaction over Ni/TiO2 catalysts and reported the beneficial effect of Ni(111) in the activity achieved. They suggested that, while the existence of multiple Ni facets induces a reaction mechanism passing through formate species as intermediates and with Ni only participating in H2 dissociation, Ni(111) promotes also CO2 dissociation into CO, leading to a mechanism where carbon monoxide acts as an intermediate for the final production of CH4. Kim et al. [214] studied Ru/TiO2 catalysts and the effect of mixing anatase and rutile TiO2 phases. Mixtures were done by using several ratios and at different stages: before RuO2 deposition and before or after annealing. A synergetic effect was verified when the mixing was done before the annealing, leading to a higher catalytic activity. In this case, RuO2 nanoparticles migrated towards the TiO2 phase to be stabilized and lately reduced. Petala et al. [206] studied the effects of alkali (Li, Na, K, Cs) incorporation on Ru/TiO2 based catalysts (0.5 and 5 wt% of Ru). The effect of the alkali promoters in the observed performances was stronger over the low Ru content catalyst (0.5%Ru/TiO2) while, in the 5%Ru/TiO2 sample no significant effects were verified. Catalytic activity was strongly improved with the addition of small alkali contents (0.2 wt%), following the order: TiO2 (unpromoted) < Li ~ K < Cs < Na, with Na-promoted sample being 3 times more active that the unpromoted catalyst. Zuzeng et al. [207] evaluated the effects of doping TiO2 with Y (0, 1.0, 2.0 and 3.0 wt%) to obtain Co/Y–TiO2 catalysts (20 wt% Co). The best catalytic activity was exhibited by the 20%Co/2%Y–TiO2 and the incorporation of Y increased the specific surface area, enhanced the reducibility, generated oxygen vacancies and improved CO2 adsorption capacity through the creation of medium basic sites. Finally, Marwood et al. [215] used 2%Ru/TiO2 to perform CO2 methanation mechanism studies and observed the formation of CO as a reaction intermediate. However, they suggested that CO was not formed as a result of the dissociative adsorption of CO2 but as a product of formate species.
Cerdá-Moreno et al. [208] reported Ni–Sepiolite catalysts, prepared by precipitation and impregnation, and a Ni–Todorokite synthesized by sequential method and with 5 wt% Ni. Authors studied not only the properties and performances but also proposed the mechanism over these two materials based on Operando DRIFTS studies. Overall, todorokite led to higher performances, even at lower reaction temperatures, mainly due to its ability to activate CO2 forming intermediate species (carbonates) further hydrogenated to formates and, later, CH4 (associative mechanism). Oppositely, sepiolite did not adsorb CO2, being CO the main reaction intermediate (dissociative mechanism). Regarding the preparation method followed for Ni/Sepiolite catalysts, precipitation improved the reducibility of the Ni species formed and the metallic dispersion.
Liang et al. [209] reported Ni/Attapulgite catalysts prepared by impregnation with varied Ni loading (5, 15 and 25 wt%). Higher Ni content reduced the textural properties of the material, the metallic-support interactions and the number of basic sites, while the average Ni0 size increased. However, and in line with several works already discussed, increasing the metal loading improved the activity and, mainly, the selectivity to CH4 as a result of the CO formation suppression.
In order to compare among the different types of new or not widely explored supports and as previously done, CH4 production rates were determined far from the equilibrium for Ni-based bentonite [196], SiC [197], Co3O4 [201], MOF–5 [203], sepiolite [208], todorokite [208] and attapulgite [209] (Figure 12). Based on the plotted data, the use and optimization of Co2O3, todorokite and sepiolite-based catalysts in CO2 methanation reaction constitute an interesting topic for further research in this area.

10. Final Remarks and Conclusions

As concluded from the results presented in previous chapters, supported nickel catalysts are the most studied for CO2 methanation reaction. However, their main drawback arises from their sintering processes and the deactivation at low temperatures, which is due to the interaction of nickel metal particles with CO and the formation of mobile nickel sub-carbonyls. Consequently, nickel catalysts are typically modified by adding promoters and stabilizers (e.g., Ce, Zr, La, Ti, Mg) in order to avoid deactivation processes [216]. Additionally, ruthenium could be considered as the alternative metal used on CO2 methanation catalysts since, although more expensive, is highly active and stable for this reaction. Even if several studies have also reported interesting results when using alternative metals (e.g., iron, cobalt), they typically required pressures above the atmospheric, as their activity is generally lower than that exhibited by Ni catalysts.
Regarding the supports, it can be concluded that their nature is crucial for the achievement of significant catalytic performances. Indeed, the dispersion and reducibility of the active metals, the activation of CO2 and H2, and the reaction mechanisms are deeply influenced by the choice of the support. As seen, SiO2 and Al2O3 have been widely and extensively studied due to their good initial activities and relatively cheap prices. However, SiO2 and Al2O3 based catalysts often suffer from sintering and serious carbon deposition at high temperatures. Additionally, CeO2 and ZrO2 have been revealed responsible for CO2 activation due to the presence of oxygen vacancies in their structure as well as promoting the reducibility and dispersion of other metals. Mesoporous SiO2 materials allow the favorable dispersion of Ni particles due to their porous structure, as occurs also in some carbon and zeolite-based supports, with the last being easily tunable in terms of hydrophobicity and basicity. MgO and hydrotalcite-derived materials present beneficial effects in terms of interaction with CO2 due to the enhanced basicity of these materials even if the reducibility of the metal species is sometimes compromised due to the strong metal-support interactions. Finally, some other materials were found promising for CO2 methanation, such as Fe and Co oxides, SiC, MOFs, sepiolite, todorokite or TiO2.
To conclude, this work aims to become a starting point or a roadmap for researchers intending to enhance heterogeneous catalysts properties and performances towards CO2 methanation. Indeed, a remarkable number of publications regarding supported catalysts for thermal Sabatier reaction were summarized not only in terms of main conclusions but also in terms of achieved catalytic performances and preparation conditions used so that researchers could have an idea about the main properties to guarantee and the main drawbacks of the chosen supports.
Based on the previous findings, we claim that the main properties to be considered for CO2 methanation catalysts must be the ones summarized in Figure 13 and described below.
  • Metallic dispersion. Based on the analyzed studies, catalysts presenting higher metallic dispersion and smaller particles present typically better performances due to the favored H2 dissociation capacity. These properties can be enhanced by tuning the preparation conditions (e.g., method, solvent, calcination and reduction temperatures), adding promoters (e.g., CeO2, MgO, ZrO2, La2O3) or, among all, through encapsulation strategies;
  • Basicity. CO2 can be, based on the previously summarized works, adsorbed and activated on basic sites. Generally, authors have identified medium strength basic sites as the most promising for CO2 methanation, since stronger sites lead to the formation of inactive carbonate species. These types of sites can be obtained by adding a basic promoter, such as MgO, or using a support able to interact with CO2 (e.g., CeO2, zeolites, Al2O3);
  • Oxygen vacancies. Several authors referred to the presence of oxygen vacancies and its responsibility in the enhancement of CO2 activation. This can be achieved by using pure or modified CeO2, ZrO2 or TiO2 oxides or even KCC–1 mesoporous material;
  • Metal-support interactions. Typically, this property is related to the metal dispersion and average particle sizes. Indeed, strong metal-support interactions (SMSI) can promote the formation of smaller and well-dispersed metallic particles, typically resisting sintering and carbon formation. However, the use of higher reduction temperatures could be required. SMSI can be obtained by tuning the preparation conditions. Indeed, while impregnation typically leads to weaker interactions, the insertion of the active metals into the support framework (e.g., ZrO2, CeO2, Al2O3, TiO2) represents a promising alternative;
  • Reducibility. Even if few works suggested a possible role of metal oxides in the reaction (e.g., NiO), metallic states of transition or noble metals are typically considered as the active phases for the CO2 methanation reaction. This property, hardly dependent on the metal-support interactions, could be responsible for a reduction in the number of available active sites. It is important to deem in mind that increasing the reduction temperature to maximize the amount of reduced species can lead to severe sintering processes, which can negatively influence the observed performances. Authors found improvements in reducibility by adding promoters (e.g., CeO2, MgO, ZrO2, La2O3);
  • Hydrophobicity. Water inhibitory effect in the CO2 methanation reaction was proved by several research studies. From the literature about zeolite supports, it was pointed out that the lower the affinity of the support for water adsorption (higher hydrophobicity), the higher the methane yields produced. The adsorption of water on the same sites for CO2 adsorption may comprehensively induce a negative kinetic effect on the methanation process. Consequently, it is important to use supports and/or promoters presenting low and weak affinity with water;
  • Textural properties. Mesoporous materials have gained attention in the last years for this reaction as they can be responsible for encapsulation effects able to reduce metallic particle sizes and strengthen metal-support interactions, resulting in higher catalytic performances. Apart from the use of conventional mesoporous materials, carbons or zeolites as supports, strategies for obtaining ordered mesoporous structures such as Al2O3 represent promising strategies.
Finally, further research on the topics presented below could be useful towards a better understanding of CO2 methanation reaction and its application at industrial scale:
  • Supports. The optimization and further development of new catalysts based on promising but not deeply explored materials such as MOFs or mesoporous silicas would be a promising pathway [217,218,219]. In addition, the utilization of waste materials (e.g., fly ash, rice husk) as support precursors constitutes an interesting route towards higher cost-efficient catalysts [220];
  • Active metals. The systematic analysis of alternative transition (e.g., Co, Fe) or noble (e.g., Ru, Rh) active metals for CO2 methanation is a topic of high interest. Complementary, further efforts towards the identification and optimization of synergistic effects in bimetallic systems (e.g., Ni-Co, Ni–Fe) through advanced characterization techniques would be highly valuable;
  • Mechanistic approaches. Modelling and in situ/Operando spectroscopy studies carried out using density-functional theory (DFT), DRIFTS or Operando FTIR under conventional and more realistic methanation conditions (e.g., incorporation of pollutants/minor compounds present in flue gases, biogas or hydrogen streams in the feed) will be advantageous. In addition, further studies dealing with the elucidation of the deactivation mechanisms over different types of catalysts, using as basis long-term and aging experiments, will be helpful;
  • Catalytic testing. Although few works dealt with the effects of incorporating CO, CH4 or even steam in the reactor feed, evaluating CO2 methanation performances under realistic conditions will be key for identifying strategies to obtain catalysts with high resistance to, among all, oxygen, steam or H2S. In addition, further research on the preparation of scale-up catalysts (e.g., monoliths) will be important.

Author Contributions

Conceptualization M.C.B.; writing—original draft preparation M.C.B. and D.S.; writing—review and editing M.C.B., D.S., L.K., J.M.L. and C.H.; supervision: J.M.L. and C.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Fundação para a Ciência e Tecnologia (FCT), grants number SFRK/BD/52369/2013 and UID/QUI/00100/2020. M.C.B. and D.S. thank also Sustainable Construction Materials Association (c5Lab) for their contracts.

Conflicts of Interest

Authors declare no conflict of interest.

References

  1. Denholm, P.; Ela, E.; Kirby, B.; Milligan, M. The Role of Energy Storage with Renewable Electricity Generation; Technical Report; National Renewable Energy Laboratory: Golden, CO, USA, 2010; pp. 1–61. [Google Scholar]
  2. Weitzel, T.; Glock, C.H. Energy management for stationary electric energy storage systems: A systematic literature review. Eur. J. Oper. Res. 2018, 264, 582–606. [Google Scholar] [CrossRef]
  3. Chen, H.; Cong, T.N.; Yang, W.; Tan, C.; Li, Y.; Ding, Y. Progress in electrical energy storage system: A critical review. Prog. Nat. Sci. 2009, 19, 291–312. [Google Scholar] [CrossRef]
  4. Blanco, H.; Faaij, A. A review at the role of storage in energy systems with a focus on Power to Gas and long-term storage. Renew. Sustain. Energy Rev. 2018, 81, 1049–1086. [Google Scholar] [CrossRef]
  5. Schaaf, T.; Grünig, J.; Schuster, M.R.; Rothenfluh, T.; Orth, A. Methanation of CO2 - storage of renewable energy in a gas distribution system. Energy Sustain. Soc. 2014, 4, 2. [Google Scholar] [CrossRef]
  6. Ghaib, K.; Ben-Fares, F.-Z. Power-to-Methane: A state-of-the-art review. Renew. Sustain. Energy Rev. 2018, 81, 433–446. [Google Scholar] [CrossRef]
  7. Sterner, M. Bioenergy and Renewable Power Methane in Integrated 100% Renewable Energy Systems: Limiting Global Warming By Transforming Energy Systems; Kassel University Press GmbH: Kassel, Germany, 2009; ISBN 978-3-89958-799-9. [Google Scholar]
  8. International Renewable Energy Agency (IRENA). Global Energy Transformation: A Roadmap to 2050 (2019 edition); International Renewable Energy Agency: Abu Dhabi, UAE, 2019. [Google Scholar]
  9. U.S. Energy Information Administration (EIA). International Energy Outlook 2019; U.S. Energy Information Administration (EIA): Washington, DC, USA, 2019.
  10. Pachauri, R.K.; Meyer, L.A. Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Intergovernmental Panel on Climate Change (IPCC): Geneva, Switzerland, 2014. [Google Scholar]
  11. Andrew, R.M. Global CO2 emissions from cement production, 1928–2018. Earth Syst. Sci. Data 2019, 11, 1675–1710. [Google Scholar] [CrossRef] [Green Version]
  12. IEA. Technology Roadmap-Low-Carbon Transition in the Cement Industry–Analysis, IEA, Paris. 2018. Available online: https://www.iea.org/reports/technology-roadmap-low-carbon-transition-in-the-cement-industry (accessed on 30 June 2020).
  13. Aresta, M.; Dibenedetto, A.; Quaranta, E. State of the art and perspectives in catalytic processes for CO2 conversion into chemicals and fuels: The distinctive contribution of chemical catalysis and biotechnology. J. Catal. 2016, 343, 2–45. [Google Scholar] [CrossRef]
  14. Aresta, M.; Dibenedetto, A. Utilisation of CO2 as a chemical feedstock: opportunities and challenges. Dalton Trans. 2007, 2975. [Google Scholar] [CrossRef]
  15. Aresta, M.; Dibenedetto, A.; Angelini, A. The changing paradigm in CO2 utilization. J. CO2 Util. 2013, 3–4, 65–73. [Google Scholar] [CrossRef]
  16. Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703–3727. [Google Scholar] [CrossRef] [Green Version]
  17. Frontera, P.; Macario, A.; Ferraro, M.; Antonucci, P. Supported Catalysts for CO2 Methanation: A Review. Catalysts 2017, 7, 59. [Google Scholar] [CrossRef]
  18. Ghaib, K.; Nitz, K.; Ben-Fares, F.-Z. Chemical Methanation of CO2: A Review. ChemBioEng Rev. 2016, 3, 266–275. [Google Scholar] [CrossRef]
  19. Jin Lee, W.; 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. [Google Scholar] [CrossRef]
  20. Rönsch, S.; Schneider, J.; Matthischke, S.; Schlüter, M.; Götz, M.; Lefebvre, J.; Prabhakaran, P.; Bajohr, S. Review on methanation – From fundamentals to current projects. Fuel 2016, 166, 276–296. [Google Scholar] [CrossRef]
  21. Malara, A.; Frontera, P.; Antonucci, P.; Macario, A. Smart recycling of carbon oxides: Current status of methanation reaction. Curr. Opin. Green Sustain. Chem. 2020, 26, 100376. [Google Scholar] [CrossRef]
  22. Ashok, J.; Pati, S.; Hongmanorom, P.; Tianxi, Z.; Junmei, C.; Kawi, S. A review of recent catalyst advances in CO2 methanation processes. Catal. Today 2020. [Google Scholar] [CrossRef]
  23. 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]
  24. Bacariza, M.C.; Graça, I.; Lopes, J.M.; Henriques, C. Tuning Zeolite Properties towards CO2 Methanation: An Overview. Chem. Cat. Chem. 2019, 11, 2388–2400. [Google Scholar] [CrossRef]
  25. 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. [Google Scholar] [CrossRef]
  26. Lv, C.; Xu, L.; Chen, M.; Cui, Y.; Wen, X.; Li, Y.; Wu, C.; 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. [Google Scholar] [CrossRef]
  27. Solis-Garcia, A.; Fierro-Gonzalez, J.C. Mechanistic Insights into the CO2 Methanation Catalyzed by Supported Metals: A Review. Available online: https://www.ingentaconnect.com/content/asp/jnn/2019/00000019/00000006/art00005 (accessed on 28 April 2020).
  28. Kuznecova, I.; Gusca, J. Property based ranking of CO and CO2 methanation catalysts. Energy Procedia 2017, 128, 255–260. [Google Scholar] [CrossRef]
  29. Erdőhelyi, A. Hydrogenation of Carbon Dioxide on Supported Rh Catalysts. Catalysts 2020, 10, 155. [Google Scholar] [CrossRef] [Green Version]
  30. Choe, S.-J.; Kang, H.-J.; Kim, S.-J.; Park, S.-B.; Park, D.-H.; Huh, D.-S. Adsorbed Carbon Formation and Carbon Hydrogenation for CO2 Methanation on the Ni(111) Surface: ASED-MO Study. Bull. Korean Chem. Soc. 2005, 26, 1682–1688. [Google Scholar] [CrossRef] [Green Version]
  31. Ren, J.; Guo, H.; Yang, J.; Qin, Z.; Lin, J.; Li, Z. Insights into the mechanisms of CO2 methanation on Ni(111) surfaces by density functional theory. Appl. Surf. Sci. 2015, 351, 504–516. [Google Scholar] [CrossRef] [Green Version]
  32. Zhao, K.; Calizzi, M.; Moioli, E.; Li, M.; Borsay, A.; Lombardo, L.; Mutschler, R.; Luo, W.; Züttel, A. Unraveling and optimizing the metal-metal oxide synergistic effect in a highly active Cox(CoO)1–x catalyst for CO2 hydrogenation. J. Energy Chem. 2021, 53, 241–250. [Google Scholar] [CrossRef]
  33. Wu, H.C.; Chang, Y.C.; Wu, J.H.; Lin, J.H.; Lin, I.K.; Chen, C.S. Methanation of CO2 and reverse water gas shift reactions on Ni/SiO2 catalysts: The influence of particle size on selectivity and reaction pathway. Catal. Sci. Technol. 2015, 5, 4154–4163. [Google Scholar] [CrossRef]
  34. Ye, R.-P.; Gong, W.; Sun, Z.; Sheng, Q.; Shi, X.; Wang, T.; Yao, Y.; Razink, J.J.; Lin, L.; Zhou, Z.; et al. Enhanced stability of Ni/SiO2 catalyst for CO2 methanation: Derived from nickel phyllosilicate with strong metal-support interactions. Energy 2019, 188, 116059. [Google Scholar] [CrossRef]
  35. 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]
  36. 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]
  37. 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]
  38. Ali, S.A.M.; Hamid, K.H.K.; Ismail, K.N. Effect of Calcination Temperature on The Structure and Catalytic Performance Of 80Ni20CO/SiO2 Catalyst for CO2 Methanation. In 3rd Electronic and Green Materials International Conference 2017 (egm 2017); Abdullah, M.M.A., Ramli, M.M., AbdRahim, S.Z., Isa, S.S.M., Saad, M.N.M., Ismail, R.C., Ghazli, M.F., Eds.; Amer Inst Physics: Melville, NY, USA, 2017; Volume 1885, p. UNSP 020272-1. ISBN 978-0-7354-1565-2. [Google Scholar]
  39. Trovarelli, A.; Deleitenburg, C.; Dolcetti, G.; Lorca, J.L. CO2 Methanation Under Transient and Steady-State Conditions over Rh/CeO2 and CeO2-Promoted Rh/SiO2: The Role of Surface and Bulk Ceria. J. Catal. 1995, 151, 111–124. [Google Scholar] [CrossRef]
  40. Li, S.; Guo, S.; Gong, D.; Kang, N.; Fang, K.-G.; Liu, Y. Nano composite composed of MoOx-La2O3Ni on SiO2 for storing hydrogen into CH4 via CO2 methanation. Int. J. Hydrogen Energy 2019, 44, 1597–1609. [Google Scholar] [CrossRef]
  41. Branco, J.B.; Brito, P.E.; Ferreira, A.C. Methanation of CO2 over nickel-lanthanide bimetallic oxides supported on silica. Chem. Eng. J. 2020, 380, 122465. [Google Scholar] [CrossRef]
  42. 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. Chem. Cat. Chem. 2020, 12, 2792–2800. [Google Scholar] [CrossRef]
  43. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Sidik, S.M. Methanation of carbon dioxide on metal-promoted mesostructured silica nanoparticles. Appl. Catal. A 2014, 486, 115–122. [Google Scholar] [CrossRef]
  44. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Saad, M.W.A. CO2 methanation over Ni-promoted mesostructured silica nanoparticles: Influence of Ni loading and water vapor on activity and response surface methodology studies. Chem. Eng. J. 2015, 260, 757–764. [Google Scholar] [CrossRef]
  45. Nguyen, P.-A.; Luu, C.-L.; Nguyen, T.-T.-V.; Nguyen, T.; Hoang, T.-C. Improving the performance of nickel catalyst supported on mesostructured silica nanoparticles in methanation of CO2-rich gas by urea–nitrate combustion. Chem. Pap. 2020. [Google Scholar] [CrossRef]
  46. Zhou, G.; Wu, T.; Xie, H.; Zheng, X. Effects of structure on the carbon dioxide methanation performance of Co-based catalysts. Int. J. Hydrogen Energy 2013, 38, 10012–10018. [Google Scholar] [CrossRef]
  47. Zhou, G.; Wu, T.; Zhang, H.; Xie, H.; Feng, Y. Carbon Dioxide Methanation on Ordered Mesoporous Co/KIT-6 Catalyst. Chem. Eng. Commun. 2014, 201, 233–240. [Google Scholar] [CrossRef]
  48. Liu, H.; Xu, S.; Zhou, G.; Xiong, K.; Jiao, Z.; Wang, S. CO2 hydrogenation to methane over Co/KIT-6 catalysts: Effect of Co content. Fuel 2018, 217, 570–576. [Google Scholar] [CrossRef]
  49. Liu, H.; Xu, S.; Zhou, G.; Huang, G.; Huang, S.; Xiong, K. CO2 hydrogenation to methane over Co/KIT-6 catalyst: Effect of reduction temperature. Chem. Eng. J. 2018, 351, 65–73. [Google Scholar] [CrossRef]
  50. Merkache, R.; Fechete, I.; Maamache, M.; Bernard, M.; Turek, P.; Al-Dalama, K.; Garin, F. 3D ordered mesoporous Fe-KIT-6 catalysts for methylcyclopentane (MCP) conversion and carbon dioxide (CO2) hydrogenation for energy and environmental applications. Appl. Catal. A-Gen. 2015, 504, 672–681. [Google Scholar] [CrossRef]
  51. Cao, H.; Wang, W.; Cui, T.; Wang, H.; Zhu, G.; Ren, X. Enhancing CO2 Hydrogenation to Methane by Ni-Based Catalyst with V Species Using 3D-mesoporous KIT-6 as Support. Energies 2020, 13, 2235. [Google Scholar] [CrossRef]
  52. Liu, Q.; Tian, Y. One-pot synthesis of NiO/SBA-15 monolith catalyst with a three-dimensional framework for CO2 methanation. Int. J. Hydrogen Energy 2017, 42, 12295–12300. [Google Scholar] [CrossRef]
  53. Lu, B.; Ju, Y.; Abe, T.; Kawamoto, K. Grafting Ni particles onto SBA-15, and their enhanced performance for CO2 methanation. RSC Adv. 2015, 5, 56444–56454. [Google Scholar] [CrossRef]
  54. Bacariza, M.C.; Graça, I.; Bebiano, S.S.; Lopes, J.M.; Henriques, C. Micro- and mesoporous supports for CO2 methanation catalysts: A comparison between SBA-15, MCM-41 and USY zeolite. Chem. Eng. Sci. 2018, 175, 72–83. [Google Scholar] [CrossRef] [Green Version]
  55. Li, Y.; Zhang, H.; Zhang, L.; Zhang, H. Bimetallic NiPd/SBA-15 alloy as an effective catalyst for selective hydrogenation of CO2 to methane. Int. J. Hydrogen Energy 2019, 44, 13354–13363. [Google Scholar] [CrossRef]
  56. Du, G.; Lim, S.; Yang, Y.; Wang, C.; Pfefferle, L.; Haller, G.L. Methanation of carbon dioxide on Ni-incorporated MCM-41 catalysts: The influence of catalyst pretreatment and study of steady-state reaction. J. Catal. 2007, 249, 370–379. [Google Scholar] [CrossRef]
  57. Wang, X.; Zhu, L.; Liu, Y.; Wang, S. CO2 methanation on the catalyst of Ni/MCM-41 promoted with CeO2. Sci. Total Environ. 2018, 625, 686–695. [Google Scholar] [CrossRef]
  58. Taherian, Z.; Khataee, A.; Orooji, Y. Promoted nickel-based catalysts on modified mesoporous silica support: The role of yttria and magnesia on CO2 methanation. Microporous Mesoporous Mater. 2020, 306, 110455. [Google Scholar] [CrossRef]
  59. Hamid, M.Y.S.; Firmansyah, M.L.; Triwahyono, S.; Jalil, A.A.; Mukti, R.R.; Febriyanti, E.; Suendo, V.; Setiabudi, H.D.; Mohamed, M.; Nabgan, W. Oxygen vacancy-rich mesoporous silica KCC-1 for CO2 methanation. Appl. Catal. A 2017, 532, 86–94. [Google Scholar] [CrossRef] [Green Version]
  60. Shahul Hamid, M.Y.; Triwahyono, S.; Jalil, A.A.; Che Jusoh, N.W.; Izan, S.M.; Tuan Abdullah, T.A. Tailoring the Properties of Metal Oxide Loaded/KCC-1 toward a Different Mechanism of CO2 Methanation by in Situ IR and ESR. Inorg. Chem. 2018, 57, 5859–5869. [Google Scholar] [CrossRef] [PubMed]
  61. Lv, C.; Xu, L.; Chen, M.; Cui, Y.; Wen, X.; Wu, C.; Yang, B.; Wang, F.; Miao, Z.; Hu, X.; et al. Constructing highly dispersed Ni based catalysts supported on fibrous silica nanosphere for low-temperature CO2 methanation. Fuel 2020, 278, 118333. [Google Scholar] [CrossRef]
  62. Liu, Q.; Dong, H. In Situ Immobilizing Ni Nanoparticles to FDU-12 via Trehalose with Fine Size and Location Control for CO2 Methanation. ACS Sustainable Chem. Eng. 2020, 8, 2093–2105. [Google Scholar] [CrossRef]
  63. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Mukti, R.R.; Taufiq-Yap, Y.H.; Sazegar, M.R. Highly active Ni-promoted mesostructured silica nanoparticles for CO2 methanation. Appl. Catal. B 2014, 147, 359–368. [Google Scholar] [CrossRef]
  64. Wang, W.; Qi, R.; Shan, W.; Wang, X.; Jia, Q.; Zhao, J.; Zhang, C.; Ru, H. Synthesis of KIT-6 type mesoporous silicas with tunable pore sizes, wall thickness and particle sizes via the partitioned cooperative self-assembly process. Microporous Mesoporous Mater. 2014, 194, 167–173. [Google Scholar] [CrossRef]
  65. Soni, K.; Rana, B.S.; Sinha, A.K.; Bhaumik, A.; Nandi, M.; Kumar, M.; Dhar, G.M. 3-D ordered mesoporous KIT-6 support for effective hydrodesulfurization catalysts. Appl. Catal. B 2009, 90, 55–63. [Google Scholar] [CrossRef]
  66. Singh, S.; Kumar, R.; Setiabudi, H.D.; Nanda, S.; Vo, D.-V.N. Advanced synthesis strategies of mesoporous SBA-15 supported catalysts for catalytic reforming applications: A state-of-the-art review. Appl. Catal. A 2018, 559, 57–74. [Google Scholar] [CrossRef]
  67. Vallet-Regi, M.; Rámila, A.; Del Real, R.P.; Pérez-Pariente, J. A New Property of MCM-41:  Drug Delivery System. Chem. Mater. 2001, 13, 308–311. [Google Scholar] [CrossRef]
  68. Corma, A.; Fornes, V.; Navarro, M.T.; Perezpariente, J. Acidity and Stability of MCM-41 Crystalline Aluminosilicates. J. Catal. 1994, 148, 569–574. [Google Scholar] [CrossRef]
  69. Gao, J.; Wang, Y.; Ping, Y.; Hu, D.; Xu, G.; Gu, F.; Su, F. A thermodynamic analysis of methanation reactions of carbon oxides for the production of synthetic natural gas. RSC Adv. 2012, 2, 2358. [Google Scholar] [CrossRef]
  70. Ocampo, F. Développement de catalyseurs pour la réaction de méthanation du dioxyde de carbone; University of Strasbourg: Strasbourg, France, 2011. [Google Scholar]
  71. Polshettiwar, V.; Cha, D.; Zhang, X.; Basset, J.M. High-Surface-Area Silica Nanospheres (KCC-1) with a Fibrous Morphology. Angew. Chem. Int. Ed. 2010, 49, 9652–9656. [Google Scholar] [CrossRef] [PubMed]
  72. Riani, P.; Garbarino, G.; Lucchini, M.A.; Canepa, F.; Busca, G. Unsupported versus alumina-supported Ni nanoparticles as catalysts for steam/ethanol conversion and CO2 methanation. J. Mol. Catal. A 2014, 383–384, 10–16. [Google Scholar] [CrossRef]
  73. Garbarino, G.; Riani, P.; Magistri, L.; Busca, G. A study of the methanation of carbon dioxide on Ni/Al2O3 catalysts at atmospheric pressure. Int. J. Hydrogen Energy 2014, 39, 11557–11565. [Google Scholar] [CrossRef]
  74. Quindimil, A.; De-La-Torre, U.; Pereda-Ayo, B.; Davó-Quiñonero, A.; Bailón-García, E.; Lozano-Castelló, D.; González-Marcos, J.A.; Bueno-López, A.; González-Velasco, J.R. Effect of metal loading on the CO2 methanation: A comparison between alumina supported Ni and Ru catalysts. Catal. Today 2019. [Google Scholar] [CrossRef]
  75. Garbarino, G.; Bellotti, D.; Riani, P.; Magistri, L.; Busca, G. Methanation of carbon dioxide on Ru/Al2O3 and Ni/Al2O3 catalysts at atmospheric pressure: Catalysts activation, behaviour and stability. Int. J. Hydrogen Energy 2015, 40, 9171–9182. [Google Scholar] [CrossRef]
  76. Hwang, S.; Hong, U.G.; Lee, J.; Seo, J.G.; Baik, J.H.; Koh, D.J.; Lim, H.; Song, I.K. Methanation of carbon dioxide over mesoporous Ni–Fe–Al2O3 catalysts prepared by a coprecipitation method: Effect of precipitation agent. J. Ind. Eng. Chem. 2013, 19, 2016–2021. [Google Scholar] [CrossRef]
  77. Serrer, M.-A.; Gaur, A.; Jelic, J.; Weber, S.; Fritsch, C.; Clark, A.H.; Saraçi, E.; Studt, F.; Grunwaldt, J.-D. Structural dynamics in Ni–Fe catalysts during CO2 methanation – role of iron oxide clusters. Catal. Sci. Technol. 2020, 10, 7542–7554. [Google Scholar] [CrossRef]
  78. Burger, T.; Augenstein, H.M.S.; Hnyk, F.; Döblinger, M.; Köhler, K.; Hinrichsen, O. Targeted Fe-Doping of Ni−Al Catalysts via the Surface Redox Reaction Technique for Unravelling its Promoter Effect in the CO2 Methanation Reaction. ChemCatChem 2020, 12, 649–662. [Google Scholar] [CrossRef]
  79. Burger, T.; Koschany, F.; Thomys, O.; Köhler, K.; Hinrichsen, O. CO2 methanation over Fe- and Mn-promoted co-precipitated Ni-Al catalysts: Synthesis, characterization and catalysis study. Appl. Catal. A: General 2018, 558, 44–54. [Google Scholar] [CrossRef]
  80. Xu, L.; Lian, X.; Chen, M.; Cui, Y.; Wang, F.; Li, W.; Huang, B. CO2 methanation over CoNi bimetal-doped ordered mesoporous Al2O3 catalysts with enhanced low-temperature activities. Int. J. Hydrogen Energy 2018, 43, 17172–17184. [Google Scholar] [CrossRef]
  81. Alrafei, B.; Polaert, I.; Ledoux, A.; Azzolina-Jury, F. Remarkably stable and efficient Ni and Ni-Co catalysts for CO2 methanation. Catal. Today 2020, 346, 23–33. [Google Scholar] [CrossRef]
  82. Rahmani, S.; Rezaei, M.; Meshkani, F. Preparation of promoted nickel catalysts supported on mesoporous nanocrystalline gamma alumina for carbon dioxide methanation reaction. J. Ind. Eng. Chem. 2014, 20, 4176–4182. [Google Scholar] [CrossRef]
  83. Liu, H.; Zou, X.; Wang, X.; Lu, X.; Ding, W. Effect of CeO2 addition on Ni/Al2O3 catalysts for methanation of carbon dioxide with hydrogen. J. Nat. Gas Chem. 2012, 21, 703–707. [Google Scholar] [CrossRef]
  84. Guo, X.; He, H.; Traitangwong, A.; Gong, M.; Meeyoo, V.; Li, P.; Li, C.; Peng, Z.; Zhang, S. Ceria imparts superior low temperature activity to nickel catalysts for CO2 methanation. Catal. Sci. Technol. 2019, 9, 5636–5650. [Google Scholar] [CrossRef]
  85. Tada, S.; Ochieng, O.J.; Kikuchi, R.; Haneda, T.; Kameyama, H. Promotion of CO2 methanation activity and CH4 selectivity at low temperatures over Ru/CeO2/Al2O3 catalysts. Int. J. Hydrogen Energy 2014, 39, 10090–10100. [Google Scholar] [CrossRef]
  86. Ahmad, W.; Younis, M.N.; Shawabkeh, R.; Ahmed, S. Synthesis of lanthanide series (La, Ce, Pr, Eu & Gd) promoted Ni/gamma-Al2O3 catalysts for methanation of CO2 at low temperature under atmospheric pressure. Catal. Commun. 2017, 100, 121–126. [Google Scholar] [CrossRef]
  87. Karam, L.; Bacariza, M.C.; Lopes, J.M.; Henriques, C.; Massiani, P.; El Hassan, N. Assessing the potential of xNi-yMg-Al2O3 catalysts prepared by EISA-one-pot synthesis towards CO2 methanation: An overall study. Int. J. Hydrogen Energy 2020. [Google Scholar] [CrossRef]
  88. Nie, W.-X.; Zou, X.-J.; Wang, X.-G.; Ding, W.-Z.; Lu, X.-G. Preparation of Highly Dispersed Ni-Ce-Zr Oxides over Mesoporous Y-Alumina and Their Catalytic Properties for CO2 Methanation. Acta Phys.-Chim. Sin. 2016, 32, 2803–2810. [Google Scholar] [CrossRef]
  89. Toemen, S.; Bakar, W.A.W.A.; Ali, R. Investigation of Ru/Mn/Ce/Al2O3 catalyst for carbon dioxide methanation: Catalytic optimization, physicochemical studies and RSM. J. Taiwan Inst. Chem. Eng. 2014, 45, 2370–2378. [Google Scholar] [CrossRef]
  90. Wan Abu Bakar, W.A.; Ali, R.; Mohammad, N.S. The effect of noble metals on catalytic methanation reaction over supported Mn/Ni oxide based catalysts. Arab. J. Chem. 2015, 8, 632–643. [Google Scholar] [CrossRef] [Green Version]
  91. 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]
  92. Cai, M.; Wen, J.; Chu, W.; Cheng, X.; Li, Z. Methanation of carbon dioxide on Ni/ZrO2-Al2O3 catalysts: Effects of ZrO2 promoter and preparation method of novel ZrO2-Al2O3 carrier. J. Nat. Gas Chem. 2011, 20, 318–324. [Google Scholar] [CrossRef]
  93. Lin, J.; Ma, C.; Wang, Q.; Xu, Y.; Ma, G.; Wang, J.; Wang, H.; Dong, C.; Zhang, C.; Ding, M. Enhanced low-temperature performance of CO2 methanation over mesoporous Ni/Al2O3-ZrO2 catalysts. Appl. Catal. B 2019, 243, 262–272. [Google Scholar] [CrossRef]
  94. Xu, J.; Lin, Q.; Su, X.; Duan, H.; Geng, H.; Huang, Y. CO2 methanation over TiO2-Al2O3 binary oxides supported Ru catalysts. Chin. J. Chem. Eng. 2016, 24, 140–145. [Google Scholar] [CrossRef]
  95. 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, 2041821. [Google Scholar] [CrossRef] [Green Version]
  96. 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. Hydrogen Energy 2018, 43, 239–250. [Google Scholar] [CrossRef]
  97. Abate, S.; Mebrahtu, C.; Giglio, E.; Deorsola, F.; Bensaid, S.; Perathoner, S.; Pirone, R.; Centi, G. Catalytic Performance of gamma-Al2O3-ZrO2-TiO2-CeO2 Composite Oxide Supported Ni-Based Catalysts for CO2 Methanation. Ind. Eng. Chem. Res. 2016, 55, 4451–4460. [Google Scholar] [CrossRef]
  98. Mebrahtu, C.; Abate, S.; Perathoner, S.; Chen, S.; Centi, G. CO2 methanation over Ni catalysts based on ternary and quaternary mixed oxide: A comparison and analysis of the structure-activity relationships. Catal. Today 2018, 304, 181–189. [Google Scholar] [CrossRef]
  99. Djinović, P.; Galletti, C.; Specchia, S.; Specchia, V. CO Methanation Over Ru–Al2O3 Catalysts: Effects of Chloride Doping on Reaction Activity and Selectivity. Top Catal. 2011, 54, 1042. [Google Scholar] [CrossRef]
  100. Konishcheva, M.V.; Potemkin, D.I.; Snytnikov, P.V.; Stonkus, O.A.; Belyaev, V.D.; Sobyanin, V.A. The insights into chlorine doping effect on performance of ceria supported nickel catalysts for selective CO methanation. Appl. Catal. B 2018, 221, 413–421. [Google Scholar] [CrossRef]
  101. Gao, Z.; Wang, L.; Ma, H.; Li, Z. Durability of catalytic performance of the chlorine-doped catalyst Ni(Clx)/ZrO2 for selective methanation of CO in H2-rich gas. Appl. Catal. A 2017, 534, 78–84. [Google Scholar] [CrossRef]
  102. Serrer, M.-A.; Kalz, K.F.; Saraҫi, E.; Lichtenberg, H.; Grunwaldt, J.-D. Role of Iron on the Structure and Stability of Ni3.2Fe/Al2O3 during Dynamic CO2 Methanation for P2X Applications. ChemCatChem 2019, 11, 5018–5021. [Google Scholar] [CrossRef] [Green Version]
  103. Mutz, B.; Belimov, M.; Wang, W.; Sprenger, P.; Serrer, M.-A.; Wang, D.; Pfeifer, P.; Kleist, W.; Grunwaldt, J.-D. Potential of an Alumina-Supported Ni3Fe Catalyst in the Methanation of CO2: Impact of Alloy Formation on Activity and Stability. ACS Catal. 2017, 7, 6802–6814. [Google Scholar] [CrossRef]
  104. Burger, T.; Ewald, S.; Niederdränk, A.; Wagner, F.E.; Köhler, K.; Hinrichsen, O. Enhanced activity of co-precipitated NiFeAlOx in CO2 methanation by segregation and oxidation of Fe. Appl. Catal. A 2020, 604, 117778. [Google Scholar] [CrossRef]
  105. Rosid, S.J.M.; Toemen, S.; Iqbal, M.M.A.; Bakar, W.A.W.A.; Mokhtar, W.N.A.W.; Aziz, M.M.A. Overview performance of lanthanide oxide catalysts in methanation reaction for natural gas production. Environ. Sci. Pollut. Res. 2019, 26, 36124–36140. [Google Scholar] [CrossRef]
  106. Bernardon, C.; Ben Osman, M.; Laugel, G.; Louis, B.; Pale, P. Acidity versus metal-induced Lewis acidity in zeolites for Friedel–Crafts acylation. Comptes Rendus Chim. 2017, 20, 20–29. [Google Scholar] [CrossRef]
  107. Zamani, A.H.; Ali, R.; Abu Bakar, W.A.W. Optimization of CO2 methanation reaction over M*/Mn/Cu–Al2O3 (M*: Pd, Rh and Ru) catalysts. J. Ind. Eng. Chem. 2015, 29, 238–248. [Google Scholar] [CrossRef]
  108. Franken, T.; Terreni, J.; Borgschulte, A.; Heel, A. Solid solutions in reductive environment – A case study on improved CO2 hydrogenation to methane on cobalt based catalysts derived from ternary mixed metal oxides by modified reducibility. J. Catal. 2020, 382, 385–394. [Google Scholar] [CrossRef]
  109. Daroughegi, R.; Meshkani, F.; Rezaei, M. Enhanced activity of CO2 methanation over mesoporous nanocrystalline Ni-Al2O3 catalysts prepared by ultrasound-assisted co-precipitation method. Int. J. Hydrogen Energy 2017, 42, 15115–15125. [Google Scholar] [CrossRef]
  110. Xu, Y.; Chen, Y.; Li, J.; Zhou, J.; Song, M.; Zhang, X.; Yin, Y. Improved low-temperature activity of Ni-Ce/gamma-Al2O3 catalyst with layer structural precursor prepared by cold plasma for CO2 methanation. Int. J. Hydrogen Energy 2017, 42, 13085–13091. [Google Scholar] [CrossRef]
  111. Song, F.; Zhong, Q.; Yu, Y.; Shi, M.; Wu, Y.; Hu, J.; Song, Y. Obtaining well-dispersed Ni/Al2O3 catalyst for CO2 methanation with a microwave-assisted method. Int. J. Hydrogen Energy 2017, 42, 4174–4183. [Google Scholar] [CrossRef]
  112. Schubert, M.; Pokhrel, S.; Thome, A.; Zielasek, V.; Gesing, T.M.; Roessner, F.; Maedler, L.; Baeumer, M. Highly active Co-Al2O3-based catalysts for CO2 methanation with very low platinum promotion prepared by double flame spray pyrolysis. Catal. Sci. Technol. 2016, 6, 7449–7460. [Google Scholar] [CrossRef] [Green Version]
  113. Aljishi, A.; Veilleux, G.; Lalinde, J.A.H.; Kopyscinski, J. The effect of synthesis parameters on ordered mesoporous nickel alumina catalyst for CO2 methanation. Appl. Catal. A 2018, 549, 263–272. [Google Scholar] [CrossRef]
  114. Xu, L.; Wang, F.; Chen, M.; Zhang, J.; Yuan, K.; Wang, L.; Wu, K.; Xu, G.; Chen, W. CO2 methanation over a Ni based ordered mesoporous catalyst for the production of synthetic natural gas. RSC Adv. 2016, 6, 28489–28499. [Google Scholar] [CrossRef]
  115. Shang, X.; Deng, D.; Wang, X.; Zou, X.; Ding, W.; Lu, X. Low Temperature Synthesis of Mesoporous gamma-Alumina supported Nickel Oxides and their Catalytic Application for CO2 Methanation. In Proceedings of the 2015 International Symposium on Energy Science and Chemical Engineering (isesce 2015); He, Y., Ed.; Atlantis Press: Paris, France, 2016; Volume 45, pp. 117–123, ISBN 978-94-6252-140-7. [Google Scholar]
  116. Le, T.A.; Kim, T.W.; Lee, S.H.; Park, E.D. CO and CO2 methanation over Ni catalysts supported on alumina with different crystalline phases. Korean J. Chem. Eng. 2017, 34, 3085–3091. [Google Scholar] [CrossRef]
  117. Beuls, A.; Swalus, C.; Jacquemin, M.; Heyen, G.; Karelovic, A.; Ruiz, P. Methanation of CO2: Further insight into the mechanism over Rh/γ-Al2O3 catalyst. Appl. Catal. B 2012, 113–114, 2–10. [Google Scholar] [CrossRef]
  118. Zhang, Z.; Tian, Y.; Zhang, L.; Hu, S.; Xiang, J.; Wang, Y.; Xu, L.; Liu, Q.; Zhang, S.; Hu, X. Impacts of nickel loading on properties, catalytic behaviors of Ni/γ–Al2O3 catalysts and the reaction intermediates formed in methanation of CO2. Int. J. Hydrogen Energy 2019, 44, 9291–9306. [Google Scholar] [CrossRef]
  119. 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 2020, 265, 118538. [Google Scholar] [CrossRef]
  120. 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. Hydrogen Energy 2012, 37, 5527–5531. [Google Scholar] [CrossRef]
  121. Atzori, L.; Cutrufello, M.G.; Meloni, D.; Cannas, C.; Gazzoli, D.; Monaci, R.; Sini, M.F.; Rombi, E. Highly active NiO-CeO2 catalysts for synthetic natural gas production by CO2 methanation. Catal. Today 2018, 299, 183–192. [Google Scholar] [CrossRef]
  122. Ratchahat, S.; Sudoh, M.; Suzuki, Y.; Kawasaki, W.; Watanabe, R.; Fukuhara, C. Development of a powerful CO2 methanation process using a structured Ni/CeO2 catalyst. J. CO2 Util. 2018, 24, 210–219. [Google Scholar] [CrossRef]
  123. 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. 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]
  124. Sharma, S.; Hu, Z.; Zhang, P.; McFarland, E.W.; Metiu, H. CO2 methanation on Ru-doped ceria. J. Catal. 2011, 278, 297–309. [Google Scholar] [CrossRef]
  125. Vita, A.; Italiano, C.; Pino, L.; Frontera, P.; Ferraro, M.; Antonucci, V. Activity and stability of powder and monolith-coated Ni/GDC catalysts for CO2 methanation. Appl. Catal. B 2018, 226, 384–395. [Google Scholar] [CrossRef]
  126. Aldana, P.A.U.; Ocampo, F.; Kobl, K.; Louis, B.; Thibault-Starzyk, F.; Daturi, M.; Bazin, P.; Thomas, S.; Roger, A.C. Catalytic CO2 valorization into CH4 on Ni-based ceria-zirconia. Reaction mechanism by operando IR spectroscopy. Catal. Today 2013, 215, 201–207. [Google Scholar] [CrossRef]
  127. Ocampo, F.; Louis, B.; Roger, A.-C. Methanation of carbon dioxide over nickel-based Ce0.72Zr0.28O2 mixed oxide catalysts prepared by sol–gel method. Appl. Catal. A 2009, 369, 90–96. [Google Scholar] [CrossRef]
  128. Ashok, J.; Ang, M.L.; Kawi, S. Enhanced activity of CO2 methanation over Ni/CeO2-ZrO2 catalysts: Influence of preparation methods. Catal. Today 2017, 281, 304–311. [Google Scholar] [CrossRef]
  129. 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]
  130. Pastor-Pérez, L.; Patel, V.; Le Saché, E.; Reina, T.R. CO2 methanation in the presence of methane: Catalysts design and effect of methane concentration in the reaction mixture. J. Energy Inst. 2020, 93, 415–424. [Google Scholar] [CrossRef]
  131. 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]
  132. Konishcheva, M.V.; Potemkin, D.I.; Badmaev, S.D.; Snytnikov, P.V.; Paukshtis, E.A.; Sobyanin, V.A.; Parmon, V.N. On the Mechanism of CO and CO2 Methanation Over Ni/CeO2 Catalysts. Top. Catal. 2016, 59, 1424–1430. [Google Scholar] [CrossRef]
  133. da Silva, D.C.D.; Letichevsky, S.; Borges, L.E.P.; Appel, L.G. The Ni/ZrO2 catalyst and the methanation of CO and CO2. Int. J. Hydrogen Energy 2012, 37, 8923–8928. [Google Scholar] [CrossRef]
  134. Lu, H.; Yang, X.; Gao, G.; Wang, K.; Shi, Q.; Wang, J.; Han, C.; Liu, J.; Tong, M.; Liang, X.; et al. Mesoporous zirconia-modified clays supported nickel catalysts for CO and CO2 methanation. Int. J. Hydrogen Energy 2014, 39, 18894–18907. [Google Scholar] [CrossRef]
  135. Jia, X.; Zhang, X.; Rui, N.; Hu, X.; Liu, C. Structural effect of Ni/ZrO2 catalyst on CO2 methanation with enhanced activity. Appl. Catal. B 2019, 244, 159–169. [Google Scholar] [CrossRef]
  136. Zhao, K.; Wang, W.; Li, Z. Highly efficient Ni/ZrO2 catalysts prepared via combustion method for CO2 methanation. J. CO2 Util. 2016, 16, 236–244. [Google Scholar] [CrossRef]
  137. Li, W.; Nie, X.; Jiang, X.; Zhang, A.; Ding, F.; Liu, M.; Liu, Z.; Guo, X.; Song, C. ZrO2 support imparts superior activity and stability of Co catalysts for CO2 methanation. Appl. Catal. B 2018, 220, 397–408. [Google Scholar] [CrossRef]
  138. Li, W.; Liu, Y.; Mu, M.; Ding, F.; Liu, Z.; Guo, X.; Song, C. Organic acid-assisted preparation of highly dispersed Co/ZrO2 catalysts with superior activity for CO2 methanation. Appl. Catal. B 2019, 254, 531–540. [Google Scholar] [CrossRef]
  139. Nagase, H.; Naito, R.; Tada, S.; Kikuchi, R.; Fujiwara, K.; Nishijima, M.; Honma, T. Ru nanoparticles supported on amorphous ZrO2 for CO2 methanation. Catal. Sci. Technol. 2020, 10, 4522–4531. [Google Scholar] [CrossRef]
  140. Takano, H.; Izumiya, K.; Kumagai, N.; Hashimoto, K. The effect of heat treatment on the performance of the Ni/(Zr-Sm oxide) catalysts for carbon dioxide methanation. Appl. Surf. Sci. 2011, 257, 8171–8176. [Google Scholar] [CrossRef]
  141. Takano, H.; Kirihata, Y.; Izumiya, K.; Kumagai, N.; Habazaki, H.; Hashimoto, K. Highly active Ni/Y-doped ZrO2 catalysts for CO2 methanation. Appl. Surf. Sci. 2016, 388, 653–663. [Google Scholar] [CrossRef]
  142. Kesavan, J.K.; Luisetto, I.; Tuti, S.; Meneghini, C.; Iucci, G.; Battocchio, C.; Mobilio, S.; Casciardi, S.; Sisto, R. Nickel supported on YSZ: The effect of Ni particle size on the catalytic activity for CO2 methanation. J. CO2 Util. 2018, 23, 200–211. [Google Scholar] [CrossRef]
  143. Kosaka, F.; Yamaguchi, T.; Ando, Y.; Mochizuki, T.; Takagi, H.; Matsuoka, K.; Fujishiro, Y.; Kuramoto, K. Effect of Ni content on CO2 methanation performance with tubular-structured Ni-YSZ catalysts and optimization of catalytic activity for temperature management in the reactor. Int. J. Hydrogen Energy 2020, 45, 12911–12920. [Google Scholar] [CrossRef]
  144. Ren, J.; Qin, X.; Yang, J.-Z.; Qin, Z.-F.; Guo, H.-L.; Lin, J.-Y.; Li, Z. Methanation of carbon dioxide over Ni–M/ZrO2 (M = Fe, Co, Cu) catalysts: Effect of addition of a second metal. Fuel Process. Technol. 2015, 137, 204–211. [Google Scholar] [CrossRef]
  145. Lu, H.; Yang, X.; Gao, G.; Wang, J.; Han, C.; Liang, X.; Li, C.; Li, Y.; Zhang, W.; Chen, X. Metal (Fe, Co, Ce or La) doped nickel catalyst supported on ZrO2 modified mesoporous clays for CO and CO2 methanation. Fuel 2016, 183, 335–344. [Google Scholar] [CrossRef]
  146. Dumrongbunditkul, P.; Witoon, T.; Chareonpanich, M.; Mungcharoen, T. Preparation and characterization of Co-Cu-ZrO2 nanomaterials and their catalytic activity in CO2 methanation. Ceram. Int. 2016, 42, 10444–10451. [Google Scholar] [CrossRef]
  147. 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]
  148. 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. Hydrogen Energy 2015, 40, 8347–8355. [Google Scholar] [CrossRef]
  149. Loder, A.; Siebenhofer, M.; Lux, S. The reaction kinetics of CO2 methanation on a bifunctional Ni/MgO catalyst. J. Ind. Eng. Chem. 2020, 85, 196–207. [Google Scholar] [CrossRef]
  150. Wierzbicki, D.; Baran, R.; Debek, R.; Motak, M.; Grzybek, T.; Galvez, M.E.; Da Costa, P. The influence of nickel content on the performance of hydrotalcite-derived catalysts in CO2 methanation reaction. Int. J. Hydrogen Energy 2017, 42, 23548–23555. [Google Scholar] [CrossRef]
  151. Bette, N.; Thielemann, J.; Schreiner, M.; Mertens, F. Methanation of CO2 over a (Mg,Al)O-x Supported Nickel Catalyst Derived from a (Ni,Mg,Al)-Hydrotalcite-like Precursor. ChemCatChem 2016, 8, 2903–2906. [Google Scholar] [CrossRef]
  152. Abate, S.; Barbera, K.; Giglio, E.; Deorsola, F.; Bensaid, S.; Perathoner, S.; Pirone, R.; Centi, G. Synthesis, Characterization, and Activity Pattern of Ni-Al Hydrotalcite Catalysts in CO2 Methanation. Ind. Eng. Chem. Res. 2016, 55, 8299–8308. [Google Scholar] [CrossRef]
  153. Yan, Y.; Dai, Y.; He, H.; Yu, Y.; Yang, Y. A novel W-doped Ni-Mg mixed oxide catalyst for CO2 methanation. Appl. Catal. B 2016, 196, 108–116. [Google Scholar] [CrossRef]
  154. Varun, Y.; Sreedhar, I.; Singh, S.A. Highly stable M/NiO–MgO (M = Co, Cu and Fe) catalysts towards CO2 methanation. Int. J. Hydrogen Energy 2020. [Google Scholar] [CrossRef]
  155. Ho, P.H.; de Luna, G.S.; Angelucci, S.; Canciani, A.; Jones, W.; Decarolis, D.; Ospitali, F.; Aguado, E.R.; Rodríguez-Castellón, E.; Fornasari, G.; et al. Understanding structure-activity relationships in highly active La promoted Ni catalysts for CO2 methanation. Appl. Catal. B 2020, 278, 119256. [Google Scholar] [CrossRef]
  156. Wierzbicki, D.; Debek, R.; Motak, M.; Grzybek, T.; Galvez, 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]
  157. 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]
  158. Zhang, L.; Bian, L.; Zhu, Z.; Li, Z. La-promoted Ni/Mg-Al catalysts with highly enhanced low-temperature CO2 methanation performance. Int. J. Hydrogen Energy 2018, 43, 2197–2206. [Google Scholar] [CrossRef]
  159. Wang, X.; Zhen, T.; Yu, C. Application of Ni-Al-hydrotalcite-derived catalyst modified with Fe or Mg in CO2 methanation. Appl. Petrochem. Res. 2016, 6, 217–223. [Google Scholar] [CrossRef] [Green Version]
  160. Mebrahtu, C.; Perathoner, S.; Giorgianni, G.; Chen, S.; Centi, G.; Krebs, F.; Palkovits, R.; Abate, S. Deactivation mechanism of hydrotalcite-derived Ni–AlOx catalysts during low-temperature CO2 methanation via Ni-hydroxide formation and the role of Fe in limiting this effect. Catal. Sci. Technol. 2019, 9, 4023–4035. [Google Scholar] [CrossRef]
  161. 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]
  162. 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]
  163. Jiménez, V.; Sánchez, P.; Panagiotopoulou, P.; Valverde, J.L.; Romero, A. Methanation of CO, CO2 and selective methanation of CO, in mixtures of CO and CO2, over ruthenium carbon nanofibers catalysts. Appl. Catal. A 2010, 390, 35–44. [Google Scholar] [CrossRef]
  164. Li, W.; Zhang, A.; Jiang, X.; Chen, C.; Liu, Z.; Song, C.; Guo, X. Low Temperature CO2 Methanation: ZIF-67-Derived Co-Based Porous Carbon Catalysts with Controlled Crystal Morphology and Size. ACS Sustain. Chem. Eng. 2017, 5, 7824–7831. [Google Scholar] [CrossRef]
  165. Gödde, J.; Merko, M.; Xia, W.; Muhler, M. Nickel nanoparticles supported on nitrogen–doped carbon nanotubes are a highly active, selective and stable CO2 methanation catalyst. J. Energy Chem. 2021, 54, 323–331. [Google Scholar] [CrossRef]
  166. Romero-Sáez, M.; Dongil, A.B.; Benito, N.; Espinoza-González, R.; Escalona, N.; Gracia, F. CO2 methanation over nickel-ZrO2 catalyst supported on carbon nanotubes: A comparison between two impregnation strategies. Appl. Catal. B 2018, 237, 817–825. [Google Scholar] [CrossRef]
  167. Wang, W.; Chu, W.; Wang, N.; Yang, W.; Jiang, C. Mesoporous nickel catalyst supported on multi-walled carbon nanotubes for carbon dioxide methanation. Int. J. Hydrogen Energy 2016, 41, 967–975. [Google Scholar] [CrossRef]
  168. Le, M.C.; Le Van, K.; Nguyen, T.H.T.; Nguyen, N.H. The Impact of Ce-Zr Addition on Nickel Dispersion and Catalytic Behavior for CO2 Methanation of Ni/AC Catalyst at Low Temperature. J. Chem. 2017, 4361056. [Google Scholar] [CrossRef] [Green Version]
  169. Gaidai, S.V.; Gryn’ko, V.S.; Zhludenko, M.G.; Dyachenko, A.G.; Tkach, V.M.; Ishchenko, O.V. Activity of Carbon-Fiber-Supported Fe-Co Catalysts in the CO2 Methanation Reaction. J. Superhard Mater. 2017, 39, 122–128. [Google Scholar] [CrossRef]
  170. Lapidus, A.L.; Gaidai, N.A.; Nekrasov, N.V.; Tishkova, L.A.; Agafonov, Y.A.; Myshenkova, T.N. The mechanism of carbon dioxide hydrogenation on copper and nickel catalysts. Pet. Chem. 2007, 47, 75–82. [Google Scholar] [CrossRef]
  171. Hussain, I.; Jalil, A.A.; Hassan, N.S.; Hambali, H.U.; Jusoh, N.W.C. Fabrication and characterization of highly active fibrous silica-mordenite (FS@SiO2-MOR) cockscomb shaped catalyst for enhanced CO2 methanation. Chem. Eng. Sci. 2020, 228, 115978. [Google Scholar] [CrossRef]
  172. Franken, T.; Heel, A. Are Fe based catalysts an upcoming alternative to Ni in CO2 methanation at elevated pressure? J. CO2 Util. 2020, 39, 101175. [Google Scholar] [CrossRef]
  173. Kitamura Bando, K.; Soga, K.; Kunimori, K.; Ichikuni, N.; Okabe, K.; Kusama, H.; Sayama, K.; Arakawa, H. CO2 hydrogenation activity and surface structure of zeolite-supported Rh catalysts. Appl. Catal. A 1998, 173, 47–60. [Google Scholar] [CrossRef]
  174. 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]
  175. Bacariza, M.C.; Maleval, M.; Graça, I.; Lopes, J.M.; Henriques, C. Power-to-methane over Ni/zeolites: Influence of the framework type. Microporous Mesoporous Mater. 2019, 274, 102–112. [Google Scholar] [CrossRef]
  176. Goodarzi, F.; Kang, L.; Wang, F.R.; Joensen, F.; Kegnæs, S.; Mielby, J. Methanation of Carbon Dioxide over Zeolite-Encapsulated Nickel Nanoparticles. ChemCatChem 2018, 10, 1566–1570. [Google Scholar] [CrossRef] [Green Version]
  177. Guo, X.; Traitangwong, A.; Hu, M.; Zuo, C.; Meeyoo, V.; Peng, Z.; Li, C. Carbon Dioxide Methanation over Nickel-Based Catalysts Supported on Various Mesoporous Material. Energy Fuels 2018, 32, 3681–3689. [Google Scholar] [CrossRef]
  178. Isah, A.; Akanyeti, İ.; Oladipo, A.A. Methanation of CO2 over zeolite–promoted Ni/Al2O3 nanocatalyst under atmospheric pressure. Reac. Kinet. Mech. Cat. 2020, 130, 217–228. [Google Scholar] [CrossRef]
  179. Quindimil, A.; De-La-Torre, U.; Pereda, 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 2018, 238, 393–403. [Google Scholar] [CrossRef]
  180. Czuma, N.; Zarębska, K.; Motak, M.; Gálvez, M.E.; Da Costa, P. Ni/zeolite X derived from fly ash as catalysts for CO2 methanation. Fuel 2020, 267, 117139. [Google Scholar] [CrossRef]
  181. Wei, L.; Haije, W.; Kumar, N.; Peltonen, J.; Peurla, M.; Grenman, H.; de Jong, W. Influence of nickel precursors on the properties and performance of Ni impregnated zeolite 5A and 13X catalysts in CO2 methanation. Catal. Today 2020. [Google Scholar] [CrossRef]
  182. da Costa-Serra, J.F.; Cerdá-Moreno, C.; Chica, A. Zeolite-Supported Ni Catalysts for CO2 Methanation: Effect of Zeolite Structure and Si/Al Ratio. Appl. Sci. 2020, 10, 5131. [Google Scholar] [CrossRef]
  183. 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]
  184. 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 2014, 147, 101–110. [Google Scholar] [CrossRef]
  185. 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]
  186. Boix, A.V.; Ulla, M.A.; Petunchi, J.O. Promoting effect of Pt on Co mordenite upon the reducibility and catalytic behavior of CO2 hydrogenation. J. Catal. 1996, 162, 239–249. [Google Scholar] [CrossRef]
  187. Wei, L.; Kumar, N.; Haije, W.; Peltonen, J.; Peurla, M.; Grénman, H.; de Jong, W. Can bi-functional nickel modified 13X and 5A zeolite catalysts for CO2 methanation be improved by introducing ruthenium? Mol. Catal. 2020, 494, 111115. [Google Scholar] [CrossRef]
  188. Bacariza, M.C.; Graça, I.; Westermann, A.; Ribeiro, M.F.; Lopes, J.M.; Henriques, C. CO2 Hydrogenation Over Ni-Based Zeolites: Effect of Catalysts Preparation and Pre-reduction Conditions on Methanation Performance. Top Catal. 2015, 59, 314–325. [Google Scholar] [CrossRef]
  189. Bacariza, M.C.; Amjad, S.; Teixeira, P.; Lopes, J.M.; Henriques, C. Boosting Ni Dispersion on Zeolite-Supported Catalysts for CO2 Methanation: The Influence of the Impregnation Solvent. Energy Fuels 2020, 34, 14656–14666. [Google Scholar] [CrossRef]
  190. 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]
  191. Delmelle, R.; Duarte, R.B.; Franken, T.; Burnat, D.; Holzer, L.; Borgschulte, A.; Heel, A. Development of improved nickel catalysts for sorption enhanced CO2 methanation. Int. J. Hydrogen Energy 2016, 41, 20185–20191. [Google Scholar] [CrossRef] [Green Version]
  192. Borgschulte, A.; Gallandat, N.; Probst, B.; Suter, R.; Callini, E.; Ferri, D.; Arroyo, Y.; Erni, R.; Geerlings, H.; Züttel, A. Sorption enhanced CO2 methanation. Phys. Chem. Chem. Phys. 2013, 15, 9620–9625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Walspurger, S.; Elzinga, G.D.; Dijkstra, J.W.; Sarić, M.; Haije, W.G. Sorption enhanced methanation for substitute natural gas production: Experimental results and thermodynamic considerations. Chem. Eng. J. 2014, 242, 379–386. [Google Scholar] [CrossRef] [Green Version]
  194. 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 2015, 174–175, 120–125. [Google Scholar] [CrossRef]
  195. Westermann, A.; Azambre, B.; Bacariza, M.C.; Graça, I.; Ribeiro, M.F.; Lopes, J.M.; Henriques, C. The promoting effect of Ce in the CO2 methanation performances on NiUSY zeolite: A FTIR In Situ/Operando study. Catal. Today 2017, 283, 74–81. [Google Scholar] [CrossRef]
  196. Lu, X.; Gu, F.; Liu, Q.; Gao, J.; Liu, Y.; Li, H.; Jia, L.; Xu, G.; Zhong, Z.; Su, F. VOx promoted Ni catalysts supported on the modified bentonite for CO and CO2 methanation. Fuel Process. Technol. 2015, 135, 34–46. [Google Scholar] [CrossRef]
  197. Le, T.A.; Kang, J.K.; Park, E.D. CO and CO2 Methanation Over Ni/SiC and Ni/SiO2 Catalysts. Top Catal. 2018, 61, 1537–1544. [Google Scholar] [CrossRef]
  198. Zhi, G.; Guo, X.; Wang, Y.; Jin, G.; Guo, X. Effect of La2O3 modification on the catalytic performance of Ni/SiC for methanation of carbon dioxide. Catal. Commun. 2011, 16, 56–59. [Google Scholar] [CrossRef]
  199. Kirchner, J.; Anolleck, J.K.; Loesch, H.; Kureti, S. Methanation of CO2 on iron based catalysts. Appl. Catal. B-Environ. 2018, 223, 47–59. [Google Scholar] [CrossRef]
  200. Baysal, Z.; Kureti, S. CO2 methanation on Mg-promoted Fe catalysts. Appl. Catal. B 2020, 262, 118300. [Google Scholar] [CrossRef]
  201. Wang, Y.; Arandiyan, H.; Scott, J.; Dai, H.; Amal, R. Hierarchically Porous Network-Like Ni/Co3O4: Noble Metal-Free Catalysts for Carbon Dioxide Methanation. Adv. Sustain. Syst. 2018, 2. [Google Scholar] [CrossRef]
  202. Wang, K.; Li, W.; Huang, J.; Huang, J.; Zhan, G.; Li, Q. Enhanced active site extraction from perovskite LaCoO3 using encapsulated PdO for efficient CO2 methanation. J. Energy Chem. 2021, 53, 9–19. [Google Scholar] [CrossRef]
  203. Zhen, W.; Li, B.; Lu, G.; Ma, J. Enhancing catalytic activity and stability for CO2 methanation on Ni@MOF-5 via control of active species dispersion. Chem. Commun. 2015, 51, 1728–1731. [Google Scholar] [CrossRef] [PubMed]
  204. Zhou, R.; Rui, N.; Fan, Z.; Liu, C. Effect of the structure of Ni/TiO2 catalyst on CO2 methanation. Int. J. Hydrogen Energy 2016, 41, 22017–22025. [Google Scholar] [CrossRef]
  205. Liu, J.; Li, C.; Wang, F.; He, S.; Chen, H.; Zhao, Y.; Wei, M.; Evans, D.G.; Duan, X. Enhanced low-temperature activity of CO2 methanation over highly-dispersed Ni/TiO2 catalyst. Catal. Sci. Technol. 2013, 3, 2627–2633. [Google Scholar] [CrossRef]
  206. 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 2018, 224, 919–927. [Google Scholar] [CrossRef]
  207. Qin, Z.; Wang, X.; Dong, L.; Su, T.; Li, B.; Zhou, Y.; Jiang, Y.; Luo, X.; Ji, H. CO2 methanation on Co/TiO2 catalyst: Effects of Y on the support. Chem. Eng. Sci. 2019, 210, 115245. [Google Scholar] [CrossRef]
  208. Cerdá-Moreno, C.; Chica, A.; Keller, S.; Rautenberg, C.; Bentrup, U. Ni-sepiolite and Ni-todorokite as efficient CO2 methanation catalysts: Mechanistic insight by operando DRIFTS. Appl. Catal. B 2020, 264, 118546. [Google Scholar] [CrossRef]
  209. Liang, C.; Gao, Z.; Lian, H.; Li, X.; Zhang, S.; Liu, Q.; Dong, D.; Hu, X. Impacts of metal loading in Ni/attapulgite on distribution of the alkalinity sites and reaction intermediates in CO2 methanation reaction. Int. J. Hydrogen Energy 2020, 45, 16153–16160. [Google Scholar] [CrossRef]
  210. Branco, J.B.; Ferreira, A.C. Methanation of CO2 Over Bimetallic Ni–5f Block Element (Th, U) Oxides. Eur. J. Inorg. Chem. 2019, 2019, 1039–1045. [Google Scholar] [CrossRef]
  211. Ferreira, A.C.; Branco, J.B. Methanation of CO2 over nanostructured nickel-4f block element bimetallic oxides. Int. J. Hydrogen Energy 2019, 44, 6505–6513. [Google Scholar] [CrossRef]
  212. Ferreira, A.C.; Branco, J.B. New approach for the synthesis of nanostructured binary f-block intermetallic compounds: CO2 methanation studies. Intermetallics 2019, 108, 32–38. [Google Scholar] [CrossRef]
  213. Branco, J.B.; da Silva, R.P.; Ferreira, A.C. Methanation of CO2 over Cobalt-Lanthanide Aerogels: Effect of Calcination Temperature. Catalysts 2020, 10, 704. [Google Scholar] [CrossRef]
  214. Kim, A.; Debecker, D.P.; Devred, F.; Dubois, V.; Sanchez, C.; Sassoye, C. CO2 methanation on Ru/TiO2 catalysts: On the effect of mixing anatase and rutile TiO2 supports. Appl. Catal. B 2018, 220, 615–625. [Google Scholar] [CrossRef]
  215. Marwood, M.; Doepper, R.; Renken, A. In-situ surface and gas phase analysis for kinetic studies under transient conditions The catalytic hydrogenation of CO2. Appl. Catal. A 1997, 151, 223–246. [Google Scholar] [CrossRef] [Green Version]
  216. Wei, W.; Jinlong, G. Methanation of carbon dioxide: An overview. Front. Chem. Sci. Eng. 2010, 5, 2–10. [Google Scholar] [CrossRef]
  217. Remya, V.R.; Kurian, M. Synthesis and catalytic applications of metal–organic frameworks: A review on recent literature. Int. Nano. Lett. 2019, 9, 17–29. [Google Scholar] [CrossRef] [Green Version]
  218. Safaei, M.; Foroughi, M.M.; Ebrahimpoor, N.; Jahani, S.; Omidi, A.; Khatami, M. A review on metal-organic frameworks: Synthesis and applications. TrAC Trends Anal. Chem. 2019, 118, 401–425. [Google Scholar] [CrossRef]
  219. Liang, J.; Liang, Z.; Zou, R.; Zhao, Y. Heterogeneous Catalysis in Zeolites, Mesoporous Silica, and Metal–Organic Frameworks. Adv. Mater. 2017, 29, 1701139. [Google Scholar] [CrossRef]
  220. Balakrishnan, M.S.; Batra, V.J.; Hargreaves, J.S.; Pulford, D.I. Waste materials – catalytic opportunities: An overview of the application of large scale waste materials as resources for catalytic applications. Green Chem. 2011, 13, 16–24. [Google Scholar] [CrossRef]
Figure 1. CO2 methanation in the Power-to-Gas context.
Figure 1. CO2 methanation in the Power-to-Gas context.
Processes 08 01646 g001
Figure 2. Structure of the present review article in terms of chapters and corresponding topics.
Figure 2. Structure of the present review article in terms of chapters and corresponding topics.
Processes 08 01646 g002
Figure 3. CH4 production rates per mass of catalyst (left) or metal (right) determined, far from thermodynamic equilibrium, for a series of Ni and Co-supported silica materials prepared by impregnation method as a function of the reaction temperatures. Italic numbers correspond to the wt% of metal in the catalysts.
Figure 3. CH4 production rates per mass of catalyst (left) or metal (right) determined, far from thermodynamic equilibrium, for a series of Ni and Co-supported silica materials prepared by impregnation method as a function of the reaction temperatures. Italic numbers correspond to the wt% of metal in the catalysts.
Processes 08 01646 g003
Figure 4. Mechanisms proposed for Ni/SiO2 (left—reproduced with permission from Wu et al., Catalysis Science and Technology; published by Royal Society of Chemistry, 2015) [33] and Pd-MgO/SiO2 (right—reproduced with permission from Park and McFarland, Journal of Catalysis; published by Elsevier, 2009) [37] catalysts, respectively.
Figure 4. Mechanisms proposed for Ni/SiO2 (left—reproduced with permission from Wu et al., Catalysis Science and Technology; published by Royal Society of Chemistry, 2015) [33] and Pd-MgO/SiO2 (right—reproduced with permission from Park and McFarland, Journal of Catalysis; published by Elsevier, 2009) [37] catalysts, respectively.
Processes 08 01646 g004
Figure 5. Mechanism proposed for KCC-1 catalysts [59]. Reproduced with permission from Hamid et al., Applied Catalysis A: General; published by Elsevier, 2017.
Figure 5. Mechanism proposed for KCC-1 catalysts [59]. Reproduced with permission from Hamid et al., Applied Catalysis A: General; published by Elsevier, 2017.
Processes 08 01646 g005
Figure 6. Synergistic effect proposed for xCo–yNi–Al2O3 catalysts [80]. Reproduced with permission from Xu et al., International Journal of Hydrogen Energy; published by Elsevier, 2018.
Figure 6. Synergistic effect proposed for xCo–yNi–Al2O3 catalysts [80]. Reproduced with permission from Xu et al., International Journal of Hydrogen Energy; published by Elsevier, 2018.
Processes 08 01646 g006
Figure 7. Mechanism proposed for Ni/Al2O3 catalysts [119]. Reproduced with permission from Cárdenas-Arenas et al., Applied Catalysis B: Environmental; published by Elsevier, 2020.
Figure 7. Mechanism proposed for Ni/Al2O3 catalysts [119]. Reproduced with permission from Cárdenas-Arenas et al., Applied Catalysis B: Environmental; published by Elsevier, 2020.
Processes 08 01646 g007
Figure 8. Mechanism proposed for Ni/CeO2 catalysts [119]. Reproduced with permission from Cárdenas-Arenas et al., Applied Catalysis B: Environmental; published by Elsevier, 2020.
Figure 8. Mechanism proposed for Ni/CeO2 catalysts [119]. Reproduced with permission from Cárdenas-Arenas et al., Applied Catalysis B: Environmental; published by Elsevier, 2020.
Processes 08 01646 g008
Figure 9. Mechanism proposed for (a) CO2 methanation and (b) CO formation over Ni-CexZr1-xO2 catalysts [126]. Reproduced with permission from Aldana et al., Catalysis Today; published by Elsevier, 2013.
Figure 9. Mechanism proposed for (a) CO2 methanation and (b) CO formation over Ni-CexZr1-xO2 catalysts [126]. Reproduced with permission from Aldana et al., Catalysis Today; published by Elsevier, 2013.
Processes 08 01646 g009
Figure 10. CH4 production rates per gram of catalyst determined, far from thermodynamic equilibrium, for a series of Ni-based ZrO2 materials with without promoters all prepared by impregnation method as a function of the reaction temperatures.
Figure 10. CH4 production rates per gram of catalyst determined, far from thermodynamic equilibrium, for a series of Ni-based ZrO2 materials with without promoters all prepared by impregnation method as a function of the reaction temperatures.
Processes 08 01646 g010
Figure 11. CH4 production rates per mass of catalyst (left) or Ni (right) determined, far from thermodynamic equilibrium, for a series of Ni/zeolites prepared by impregnation method as a function of the reaction temperatures. Italic numbers correspond to the wt% of Ni in the catalysts.
Figure 11. CH4 production rates per mass of catalyst (left) or Ni (right) determined, far from thermodynamic equilibrium, for a series of Ni/zeolites prepared by impregnation method as a function of the reaction temperatures. Italic numbers correspond to the wt% of Ni in the catalysts.
Processes 08 01646 g011
Figure 12. CH4 production rates per mass of catalyst (left) or Ni (right) determined, far from thermodynamic equilibrium, for a series of Ni-supported catalysts prepared using non-conventional supports a function of the reaction temperatures. Italic numbers correspond to the wt% of Ni in the catalysts.
Figure 12. CH4 production rates per mass of catalyst (left) or Ni (right) determined, far from thermodynamic equilibrium, for a series of Ni-supported catalysts prepared using non-conventional supports a function of the reaction temperatures. Italic numbers correspond to the wt% of Ni in the catalysts.
Processes 08 01646 g012
Figure 13. Favorable properties for obtaining active, selective and stable CO2 methanation catalysts.
Figure 13. Favorable properties for obtaining active, selective and stable CO2 methanation catalysts.
Processes 08 01646 g013
Table 1. SiO2-based materials reported in literature for CO2 methanation reaction.
Table 1. SiO2-based materials reported in literature for CO2 methanation reaction.
CatalystPreparation MethodTred
(°C)
H2:CO2QT/W
(mL g−1 h−1)
Best Catalytic Performances Ref.
T
(°C)
XCO2
(%)
SCH4
(%)
10Ni/SiO2Impregnation5004:1120,0003501090[33]
40Ni/SiO2Impregnation5004:110,0003706290[34]
40Ni/SiO2Ammonia-evaporation5004:110,0003708095[34]
10Ni-MgO/SiO2Co-impregnation4504:124,0004007298[35]
6.2Pd-MgO/SiO2Reverse microemulsion4504:173204505995[36,37]
80Ni-Co/SiO2Co-precipitation3504:130,0003504998[38]
1Rh/CeO2/SiO2Impregnation5003:1n.a.23010n.a.[39]
6Ni-La-Mo/SiO2Impregnation7004:115,00030075100[40]
Table 2. Mesoporous SiO2-based materials reported in literature for CO2 methanation reaction.
Table 2. Mesoporous SiO2-based materials reported in literature for CO2 methanation reaction.
CatalystPreparation MethodSA a
(m2 g−1)
Tred
(°C)
H2:CO2QT/W
(mL g−1 h−1)
Best Catalytic PerformancesRef.
T
(°C)
XCO2
(%)
SCH4
(%)
5Rh/MSNImpregnation933 (3.6)5004:150,00035099100[43]
10Ni/MSNImpregnation662 (3.5)5004:150,00035085100[44]
50Ni/MSNImpregnation134 (2.3)4504:115,00037596100[45]
20Co/KIT–6Impregnation369 (6.4)4004.6:122,00028049100[46,47]
25Co–KIT–6Impregnation323 (~8)4004:160,0003605395[48,49]
Fe–KIT–6HS b435 (6.5)3504:150,000500163[50]
20Ni–V2O5/KIT–6Impregnationn.a.5504:196,00035087100[51]
15Ni-SBA–15One-pot HS b574 (4.5)5004:110,0004207696[52]
10Ni/SBA–15Grafting551 (6.3)5504:120,0004508092[53]
15Ni/SBA–15Impregnation235 (8.7)4704:186,2004006593[54]
CeO2/15Ni/SBA–15Impregnation320 (8.3)4704:186,2003507097[54]
74Ni–Pd/SBA–15Impregnation535 (6.4)6004:160004309697[55]
3Ni–MCM–41One-pot HS b1480 (2.9)7004:157604001796[56]
15Ni/MCM–41Impregnation847 (1.9)4704:186,2004007093[54]
CeO2/15Ni/MCM–41Impregnation589 (2.0)4704:186,2004007595[54]
20Ni–CeO2/MCM–41Deposition-precipitation302 (4.6)4704:1300038086100[57]
Ni–Y2O3/MgO–MCM–41DS method c445 (3.4)6004:190004006585[58]
KCC–1Microemulsion773 (4.6)5504:150,0004504984[59]
5Co/KCC–1Impregnation318 (4.9)5004:1450040072n.a.[60]
5Ni/KCC–1Impregnation537 (4.7)5004:1450040093>95[60]
20Ni/KCC–1Impregnation216 (3.4)5004:112,0003758298[61]
10Ni/FDU–12Impregnation 506 (8.6)6004:160,0004257997[62]
10Ni–CeO2/FDU–-12Impregnation500 (8.7)6004:160,0004008197[62]
a Surface area determined by Brunauer–Emmett–Teller (BET) method. Values in brackets in this column correspond to pore diameters (nm); b Hydrothermal synthesis; c Direct synthesis method.
Table 3. Al2O3-based materials reported in literature for CO2 methanation reaction.
Table 3. Al2O3-based materials reported in literature for CO2 methanation reaction.
CatalystPreparation MethodTred
(°C)
H2:CO2QT/W
(mL g−1 h−1)
Best Catalytic Performances Ref.
T
(°C)
XCO2
(%)
SCH4
(%)
55Ni/Al2O3Impregnationn.a.5:1n.a.5007186[72,73]
12Ni/Al2O3Impregnation5005:130,00042580100[74]
3Ru/Al2O3Impregnationn.a.5:1n.a.3509596[75]
4Ru/Al2O3Impregnation3005:130,00037585100[74]
30Ni–Fe–Al2O3Co-precipitation7004:196002205899[76]
13Ni–Fe–γ–Al2O3Co-precipitation5004:1353,0004506692[77]
36Ni–Fe–Al2O3Co-precipitation5004:1150,0003259899[78]
36Ni–Mn–Al2O3Co-precipitation5004:1150,0003009299[79]
8Ni–Co–Al2O3EISA one-pot8004:115,0004008098[80]
10Ni–Co/γ–Al2O3Impregnation4004:113032590100[81]
20Ni/CeO2/γ–Al2O3Impregnation4503.5:1900035080100[82]
15Ni–CeO2/Al2O3Co-impregnation5004:115,00035085100[83]
CeO2/42Ni–Al2O3Hydrothermal 5004:1n.a.2509099[84]
2Ru/CeO2/Al2O3Impregnation5004:17235090100[85]
12Ni–Pr/γ–Al2O3Impregnation7504:1600030098100[86]
15Ni–Mg–Al2O3EISA one-pot8004:186,1004007096[87]
15Ni–CeO2–ZrO2/γ–Al2O3Impregnation6004:1300030090100[88]
5Ru/Mn/CeO2/Al2O3Impregnationn.a.4:16362009891[89]
5Ru/Mn/Ni/Al2O3Impregnationn.a.4:150040010072[90]
(CaO/11NiTiO3)/γ–Al2O3Impregnation7004:1n.a.40053n.a.[91]
12Ni/(ZrO2–Al2O3)Impregnation4503.5:181003607070[92]
10Ni/(Al2O3–ZrO2)Epoxide-driven sol-gel5004:1600034077100[93]
5Ru/(TiO2–Al2O3)Impregnation4004:160,00037582100[94]
15Ni/(CaO–Al2O3)Impregnation5004:1n.a.4506692[95]
10Ni–Ru–(CaO–Al2O3)EISA one-pot6004:130,00038084100[96]
20Ni/(γ–Al2O3–ZrO2–TiO2–CeO2)Impregnation5004:120,0003008298[97]
20Ni/(γ–Al2O3–ZrO2–TiO2–CeO2)Impregnation5004:160,00035090n.a.[98]
Table 4. CeO2-based materials reported in literature for CO2 methanation reaction.
Table 4. CeO2-based materials reported in literature for CO2 methanation reaction.
CatalystPreparation MethodTred
(°C)
H2:CO2QT/W
(mL g−1 h−1)
Best Catalytic Performances Ref.
T
(°C)
XCO2
(%)
SCH4
(%)
10Ni/CeO2Impregnation6004:1n.a.30090100[85,120]
32Ni-CeO2Soft-template4004:172,00030087100[121]
10Ni/CeO2Impregnation 5004:1n.a.30092100[122]
8.5Ni/CeO2Impregnation5004:130,00037580100[123]
Ce0.95Ru0.05O2Combustion 5004:1n.a.4505599[124]
50Ni/CexGd1-xO2Solution-combustion8004:1n.a.45072n.a.[125]
5Ni-Ce0.72Zr0.28O2Sol-gel4004:122,0004008599[126,127]
10Ni-CexZr1-xO2Ammonia evaporation4504:122,00027555100[128]
13Ni-NaO/CeO2Impregnation50050:160,0002509796[129]
16Ni-Co/CeO2-ZrO2Impregnation4504:112,0003507098[130]
Table 5. ZrO2-based materials reported in literature for CO2 methanation reaction.
Table 5. ZrO2-based materials reported in literature for CO2 methanation reaction.
CatalystPreparation MethodTred
(°C)
H2:CO2QT/W
(mL g−1 h−1)
Best Catalytic PerformancesRef.
T
(°C)
XCO2
(%)
SCH4
(%)
10Ni/ZrO2Impregnationn.a.125:1140,000450100n.a.[133]
15Ni/ZrO2Impregnation5004:1n.a.3508099[134]
10Ni/ZrO2Impregnation a5004:110,0003508097[135]
15Ni/ZrO2Combustion5004:148,00040085100[136]
10Co/ZrO2Impregnation4004:1360040093100[137]
2Co/ZrO2Impregnation4004:172,0004008599[138]
3Ru/ZrO2Selective deposition method3004:1n.a.30082100[139]
50Ni/Sm-ZrO2Mechanical mixture3004:1300035095100[140]
Ni/Y–ZrO2Co-impregnation4004:16300 b40038 c[141]
10Ni/Y–ZrO2Impregnation with EDTA 5004:160,0003756096[142]
75Ni/Y–ZrO2Extrusion 6004:1n.a.300d9298[143]
30Ni–Fe/ZrO2Co-impregnation4004:1498027010095[144]
30Ni–Co/ZrO2Co-impregnation4004:1498027010095[144]
30Ni–Cu/ZrO2Co-impregnation4004:149803308888[144]
15Ni–Fe/ZrO2Co-impregnation5004:1n.a.4007896[145]
Co–Cu/ZrO2Co-precipitation3003:114,4003006883[146]
6Ni–MgO/ZrO2Impregnation with citric acid4504:115,00030095100[147]
Ni–CaO/ZrO2Co-impregnation4004:16742 b35085100[148]
a Plasma decomposition instead of thermal calcination; b In these works ɣ–Al2O3, typical support for CO2 methanation catalysts with ability to activate CO2, was loaded in the reactor together with the corresponding Ni-based ZrO2 catalyst. Consequently, the mass of ɣ–Al2O3 was also considered in the QT/W calculations; c CH4 yield; d Temperature of the furnace.
Table 6. MgO and HT-derived materials reported in literature for CO2 methanation reaction.
Table 6. MgO and HT-derived materials reported in literature for CO2 methanation reaction.
CatalystPreparation MethodTred
(°C)
H2:CO2QT/W
(mL g−1 h−1)
Best Catalytic Performances Ref.
T
(°C)
XCO2
(%)
SCH4
(%)
27Ni/MgOImpregnation n.a.4:1370032587>99[149]
42.5Ni–HTderivedCo-precipitation9004:1n.a.3008299[150]
59Ni–HTderivedCo-precipitation9004:111003307495[151]
80Ni–HTderivedCo-precipitation5004:120,00030085100[152]
Ni–WOx–MgOCo-precipitation5004:160,00030085100[153]
2Co/NiO–MgOImpregnation1208:160,00032592100[154]
56Ni–La2O3–HTderivedCo-precipitation6004:180,0002757096[155]
15Ni–La2O3–Ni–HTderivedCo-precipitation9004:1n.a.3007598[156]
21Ni–La2O3–HTderivedIon-exchanged9004:1n.a.30082100[157]
15Ni–La2O3–HTderivedUrea hydrolysis 7004:1450003009095[158]
Ni–Fe–HTderivedCo-precipitation7004:11200030095100[159]
12Ni–Fe–HTderivedCo-precipitation5004:1n.a.3008397[160]
78Ni–K2O–Al2O3–HTderivedCo-precipitation5004:175,00035085100[161]
Table 7. Carbon-based materials reported in literature for CO2 methanation reaction.
Table 7. Carbon-based materials reported in literature for CO2 methanation reaction.
CatalystPreparation MethodTred
(°C)
H2:CO2QT/W
(mL g−1 h−1)
Best Catalytic PerformancesRef.
T
(°C)
XCO2
(%)
SCH4
(%)
0.5Ru/CNFsImpregnation3003.3:180,00050052100[163]
32.5Co–PCCarbonization4004:172,0002705399[164]
30Ni/NCNTsImpregnation4204:150,0003405196[165]
5Ni/ZrO2/CNTsImpregnation5005:120,0004005596[166]
12Ni–CeO2/CNTsUltrasonic-assisted co-impregnation3504:130,00035084100[167]
7Ni/Ce0.2Zr0.8O2/ACImpregnation6004:140,00030085100[168]
30Fe–Co/CFUniform infiltration30027:166003906588[169]
Table 8. Zeolite-based materials reported in literature for CO2 methanation reaction.
Table 8. Zeolite-based materials reported in literature for CO2 methanation reaction.
CatalystPreparation MethodTred
(°C)
H2:CO2QT/W
(mL g−1 h−1)
Best Catalytic Performances Ref.
T
(°C)
XCO2
(%)
SCH4
(%)
SiO2–MOR aMicroemulsion5004:145,0005006568[171]
1Fe/13XImpregnation4004:1n.a.3508976[172]
6Rh–YIon exchange4503:160001506100[173]
15Ni/USYImpregnation4704:186,2004007397[174]
15Ni/MORImpregnation4704:186,2004406995[175]
15Ni/ZSM-5Impregnation4704:186,2004506894[175]
5Ni/d-S1Impregnation5004:172,0004505791[176]
10Ni/ZSM-5Impregnation5004:1n.a.4007699[177]
10Ni–ZSM-5Hydrothermal method b4003:1n.a.40066100[178]
10Ni/BEAImpregnation5004:130,0004507397[179]
15Ni/XImpregnation 4704:112,0004505390[180]
5Ni/13XEvaporation impregnation5004:113,33332080100[181]
5Ni/5AEvaporation impregnation5004:113,3334006595[181]
5Ni/ITQ-2Impregnation4504:190004008299[182]
MgO/13Ni/USYImpregnation7004:186,2004006393[183]
CeO2/14Ni/USYImpregnation4704:186,2004006895[184]
20Ni–CeO2/USYCo-impregnation4704:186,2003057899[185]
0.5Pt–Co–MORIon exchange3504:1n.a.3504115[186]
10Ni/La2O3/USYImpregnation5004:130,00040075100[179]
2.5Ni–2.5Ru/13XEvaporation impregnation5004:113,3334006592[187]
4Ni–1Ru/5AEvaporation impregnation5004:113,3334006094[187]
a Fibrous silica–mordenite; b Using as Si source a Ni/SiO2 material with Ni incorporated by impregnation.
Table 9. Different supported materials reported in literature for CO2 methanation.
Table 9. Different supported materials reported in literature for CO2 methanation.
CatalystPreparation MethodTred
(°C)
H2:CO2QT/W
(mL g−1 h−1)
Best Catalytic Performances Ref.
T
(°C)
XCO2
(%)
SCH4
(%)
20Ni/BentoniteImpregnation5504:130,0004007490[196]
20Ni–VOx/BentoniteImpregnation5504:130,0003808987[196]
20Ni/SiCDeposition-precipitation60050:160,000325100100[197]
15Ni/SiCImpregnation4004:1n.a.4008099[198]
15Ni–La2O3/SiCCo-impregnation4004:1n.a.3508599[198]
γ–Fe2O3Commercial450200:1150,0004005070[199]
Mg/α–Fe2O3Impregnation3504:1150,0004004965[200]
12Ni/Co3O4Impregnation 4504:148,000250100100[201]
3Pd–LaCoO3One pot n.a.3:118,00030062>99[202]
10Ni/MOF-5Impregnation n.a.4:1750032075100[203]
6.2Ni/TiO2Impregnation a7004:1n.a.35073100[204]
15Ni/TiO2Deposition-precipitation4504:124002609699[205]
5Ru/TiO2Impregnation3004:190,00030080100[206]
20Co/Y2O3–TiO2Deposition-precipitation4004:1360035086100[207]
5Ni–SepiolitePrecipitation4504:1900040088100[208]
15Ni–TodorokiteSequential method4504:1900030090100[208]
25Ni/AttapulgiteImpregnation6004:111,4004008599[209]
a Calcination under DBD plasma.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bacariza, M.C.; Spataru, D.; Karam, L.; Lopes, J.M.; Henriques, C. Promising Catalytic Systems for CO2 Hydrogenation into CH4: A Review of Recent Studies. Processes 2020, 8, 1646. https://doi.org/10.3390/pr8121646

AMA Style

Bacariza MC, Spataru D, Karam L, Lopes JM, Henriques C. Promising Catalytic Systems for CO2 Hydrogenation into CH4: A Review of Recent Studies. Processes. 2020; 8(12):1646. https://doi.org/10.3390/pr8121646

Chicago/Turabian Style

Bacariza, M. Carmen, Daniela Spataru, Leila Karam, José M. Lopes, and Carlos Henriques. 2020. "Promising Catalytic Systems for CO2 Hydrogenation into CH4: A Review of Recent Studies" Processes 8, no. 12: 1646. https://doi.org/10.3390/pr8121646

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