Next Article in Journal
Power Measurement Using Adaptive Chirp Mode Decomposition for Electrical Vehicle Charging Load
Next Article in Special Issue
A Review of On-Board Carbon Capture and Storage Techniques: Solutions to the 2030 IMO Regulations
Previous Article in Journal
Application of Particle Swarm Optimization (PSO) Algorithm in Determining Thermodynamics of Solid Combustibles
Previous Article in Special Issue
A Three-Phase Phase-Modular Single-Ended Primary-Inductance Converter Rectifier Operating in Discontinuous Conduction Mode for Small-Scale Wind Turbine Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Looking for an Optimal Composition of Nickel-Based Catalysts for CO2 Methanation

1
Dipartimento di Ingegneria Civile, Chimica e Ambientale, Università di Genova, Via Opera Pia 15, 16145 Genova, Italy
2
Consorzio Interuniversitario Nazionale di Scienza e Tecnologia dei Materiali (INSTM), UdR di Genova, Via Dodecaneso 31, 16146 Genova, Italy
3
Dipartimento di Chimica e Chimica Industriale, Università di Genova, Via Dodecaneso 31, 16146 Genova, Italy
4
CNR-SCITEC G. Natta, Via Golgi 19, 20133 Milano, Italy
*
Author to whom correspondence should be addressed.
Energies 2023, 16(14), 5304; https://doi.org/10.3390/en16145304
Submission received: 1 June 2023 / Revised: 3 July 2023 / Accepted: 5 July 2023 / Published: 11 July 2023
(This article belongs to the Special Issue Green Technologies for Energy Transition)

Abstract

:
A detailed critical analysis of the scientific literature data concerning catalysts for CO2 methanation based on nickel supported over oxides was performed. According to the obtained information, it seems that an ionic support is necessary to allow a good nickel dispersion to produce very small nickel metal particles. Such small metal particles result in being very active toward methanation, limiting the production of carbonaceous materials. The use of support and/or surface additives gives rise to medium surface basicity, allowing medium-strong adsorption of CO2, and it is also advisable to increase the reaction rate. A medium nickel loading would allow the free support geometric surface to be covered densely by small nickel metal particles without the production of larger Ni crystals. It is also advisable to work at temperatures where Ni(CO)4 formation is not possible (e.g., >573 K). The promising properties of systems based on doped Ni/Al2O3, doped with basic and re-active oxides such as MnOx or/and CeO2, and those based on Ni/CeO2 were underlined.

1. Introduction

The methanation of captured CO2 using green hydrogen is considered a possible way to valorize and reuse it, also in view of the so-called “power to Gas” technologies [1,2]. The new process to be designed for CO2 methanation [3] is strictly related to ones developed decades ago for Substitute Natural Gas (SNG) production from CO-rich syngases [4]. It possibly implies a CO2 purification step followed by a catalytic reaction step. Depending on the performances and design of the catalytic reactor and on the final use of the produced methane, the following steps may be needed: condensation and removal of coproduced water, product separation from unreacted CO2 and/or hydrogen (that can be recycled to the reactor), and separation from byproducts, such as CO and/or other hydrocarbons or oxygenated compounds. Thus, as always for industrial chemical plants, the key step for an efficient process is to develop high activity and high selectivity catalysts, just to reduce the onerousness of the separation and recycling steps and to maximize productivity. Although artificial intelligence-based methods are under development for discovering new and better catalysts for industrial chemical reactions [5], up to now, empirical methods based on human intelligence still represent the most reliable approach, in our opinion.

2. Thermodynamics of CO2 Hydrogenation

The CO2 methanation reaction,
CO2 + 4 H2 ⇆ CH4 + 2 H2O
also denoted as the “Sabatier reaction”, is an exothermic equilibrium reaction: thus, it is most favored at low temperatures and high pressures. However, according to its large exothermicity (ΔH°298 = −165 kJ/mol), the theoretical thermodynamic methane yield approaches totality at temperatures up to 473 K, also at 1 atm [6,7]. At higher temperatures, CO2 conversions and methane yields allowed by thermodynamics decrease. However, methane synthesis is competitive with the production of CO through the so-called “reverse Water Gas Shift” reaction, rWGS,
CO2 + H2 ⇆ CO + H2O
which is weakly endothermic (ΔH°298 = +41.16 kJ/mol) and almost independent from reaction pressure; thus, its equilibrium is shifted to products the higher the temperature is. Thus, thermodynamically allowed CO2 conversion will increase again, and at higher temperatures, CO will become the predominant product [6]. However, given the strong dependence of methanation on, and the almost independence of rWGS from, total pressure, at higher pressures the temperature range where methanation is predominant will increase to higher temperatures. In Figure 1, the resulting CO2 conversions and CH4 and CO yields at thermodynamic equilibrium are reported, as a function of temperature, at 0.1, 2.03, and 20.3 MPa.
The production of higher hydrocarbons (paraffins and olefins), as it occurs in the so-called CO2-Fischer Tropsch process, through the following reactions,
n CO2 + (3n + 1) H2 = CnH2n + 2 + 2n H2O     n > 1
m CO2 + 3m H2 = CmH2m + 2m H2O     m > 1
also represent competitive reactions, in principle. These reactions, which are also exothermic equilibrium reactions most favored at low temperature and high pressure, are less favored than methanation and may be realized in conditions where rWGS is still not favored. The higher hydrocarbon yield, forecasted at thermodynamic equilibrium at atmospheric pressure, is extremely low, and methane also remains the most favored hydrocarbon product at higher pressures [9].
According to Gao et al. [6] and our own thermodynamic calculations, carbon formation theoretically does not occur when the starting molar ratio is H2/CO2 = 4, while it should occur at T < 873 K when H2/CO2 = 2. In contrast, the calculations of Schmider et al. [10] report the formation of significant amounts of carbon by CO2 hydrogenation with H2/CO2 = 4 at atmospheric pressure, with a maximum amount at around 800 K, which almost disappears if reaction pressure is raised to 2 MPa.
Thus, the overall available thermodynamic data indicate that, in principle, the best conditions for an industrial CO2 methanation process are to realize the reaction in the presence of a selective catalyst for methanation, with low or no activity for reverse water gas shift and higher hydrocarbon synthesis, with a moderate H2 excess (H2/CO2 > 4) working at T < 650 K or less at a moderately high pressure (1–2 MPa) to push conversion and selectivity to methane, limiting coproduction of CO and carbon deposition. In addition, the adoption of high space velocities will limit the significant production of higher hydrocarbons, achieving high CH4 productivity.
To test and develop optimal catalysts, however, conditions in which possible limits become evident may give more information. To evidence and compare intrinsic catalytic activity and selectivity toward methane, as well as stability with respect to carbon deposition phenomena, studies realized at atmospheric pressure are more informative than those realized at high pressures [11], where the production of CO and carbon are essentially thermodynamically not possible.

3. Nickel Catalysts for CO2 Hydrogenation Reactions

As is well known, transition metals are the most catalytically active materials for hydrogenation reactions, at least in sulfur-free environments [12]. As mentioned above, CO2 hydrogenation can produce CO by reverse water gas shift, methane, and any other hydrocarbon and oxygenated organic compound. According to the literature, while Cu- and Pd-based catalysts are very active for methanol and higher alcohol synthesis, and Fe, Co, and Ru catalysts are active for higher hydrocarbon synthesis, Ni and Ru are the most active for COx methanation reactions. In fact, commercial catalysts are available for methanation in CO-rich syngases [12]: they are essentially based on Ru/Al2O3 for low-temperature applications, mainly for the methanation of COx residues in hydrogen and ammonia synthesis gas. Ni/Al2O3 is used for the same application and for high-temperature applications for the production of SNG [13].
An enormous amount of research work was recently carried out concerning catalysts for CO2 methanation. Several reviews collect such studies [14,15,16,17]. Most of these papers, however, when concerning catalyst composition, appear to be mere collections of data with poor criticism. According to the literature, the commercial catalysts for CO-rich syngas methanation are also active for pure CO2 methanation, although the optimal composition for these quite different applications can be remarkably different. In fact, although CO methanation and CO2 methanation are certainly related reactions, the two reactants CO and CO2 have very different chemical and adsorption properties. CO adsorbs strongly, also at very low temperatures, on metal surfaces, including nickel [18], with which it can even react in mild conditions, producing the volatile carbonyl compound Ni(CO)4 [19]. Instead, the interaction of CO with oxides, such as catalyst carriers, is essentially weak [20]. In contrast, CO2 interaction with metal surfaces may be relatively weak [21] and may dissociatively occur at moderate temperatures, such as on Ni [22]. Instead, CO2 adsorption on metal oxides occurs at room and even lower temperatures and may be very strong with basic oxides, producing a number of surface and subsurface species [23,24] and sometimes bulk carbonates [21]. The stronger reactivity of CO2 with oxide carriers suggests that the carrier can have a more relevant role in the kinetics of CO2 hydrogenation than of CO hydrogenation.
Taking into account that nickel is by far cheaper than ruthenium and other noble metals, and that the reaction temperature for CO2 methanation may not necessarily be so low, a large part of the current research is centered on nickel catalysts [25,26]. The main approach is to increase the low-temperature activity of Ni-based catalysts to retain total selectivity to methane, limiting deactivation phenomena.

3.1. Unsupported Nickel Catalysts

Unsupported Ni (nano)particles can be produced in several ways, as summarized by Abboud et al. [27]. The preparation of ferromagnetic Ni particles [28] can help in the recovery of the catalyst by magnetic methods or by using induction heating systems. Thermal stability against sintering is a main point concerning unsupported metal particles. Lucchini et al. [29] reported relevant sintering and consequent deactivation of differently prepared unsupported nickel nanoparticles upon CO catalytic hydrogenation experiments in the 550–800 K range.
CO2 methanation was investigated on unsupported Ni metal particles produced by the reduction of NiO with H2 by Garbarino et al. [30], Pandey et al. [31], Varvoutis et al. [32], and Kristiani and Takeishi [33], while Lee et al. tested Raney nickel [34]. Ni particles sized at 100–200 nm show moderate catalytic activity, definitely lower than for supported nickel [30,32]. Low Turn Over Frequency [33] but also low selectivities to methane with both coproduction of CO and a quite quick deactivation was associated with the production of encapsulating carbon and carbon nanotubes [30]; see Figure 2.
Similarly, Hao et al. [35] reported low activity and low selectivity to methane over “bare nickel”. According to Pandey [31], the moderate activity of pure nickel is significantly increased by the addition of small amounts of iron. On Raney nickel, selectivity to methane is reported to be not far from 100% for the purest Ni sample, which is also the most active [34]. Very small (d < 8 nm) unsupported nickel nanoparticles produced by the reduction of NiCl2*6H2O with NaBH4 were found almost inactive in CO2 methanation by Riani et al. [36]. However, it must be taken into account that such a preparation method may leave B impurities on the metal samples [37], which can affect catalytic behavior. Lan et al. [38] reported that the increase in the (111) to (200) crystal planes, obtained by exposing the catalyst to a magnetic field during preparation, enormously increases the catalytic activity of unsupported nickel at 453 K, with total methane selectivity.

3.2. Nickel Supported on Simple Metal Oxide Carriers

Several papers report a comparison of the catalytic activity of nickel catalysts supported on different simple oxide carriers for CO2 methanation, with partially agreeing data. According to Muroyama et al. [39], the CO2 conversion for most catalysts drastically increases at 498–523 K and reaches a maximal value at 573–623 K, approaching thermodynamically allowed values for the given conditions. The order of CH4 yield at 523 K was as follows: Ni/Y2O3 > Ni/Sm2O3 > Ni/ZrO2 > Ni/CeO2 > Ni/Al2O3 > Ni/La2O3. The catalytic activity can be partly explained by the basic property of the catalysts. However, the same authors reported different scales for selectivity, with Ni/Al2O3 being the most selective and performant in methanation at 673 K. In at least partial agreement, Hatzisymeon et al. [40] reported the following sequence of activity: Ni/CeO2 > Ni/Y2O3− stabilized ZrO2 > Ni/Al2O3 > Ni/TiO2, all with 5 wt% Ni. These authors attribute the trend to the decreasing particle size of Ni metal particles. According to Ma et al. [41], the CO2 methanation activity order for 5 wt% nickel catalysts on relatively low surface area supports is Ni/CeO2 > Ni/Al2O3  >  Ni/TiO2  >  Ni/ZrO2. According to Gac et al. [42], the activity trend for 10–40 wt% Ni-supported catalysts is Ni/CeO2 > Ni/ZrO2 > Ni/Al2O3, all with very high methane selectivity. However, these authors underline that Al2O3-supported catalysts have the highest resistance against sintering. Lee et al. [43] reported the following activity trend: Ni/CeO2 ≈ Ni/Y2O3 > Ni/Al2O3 > Ni/TiO2. These authors suggest that the reducibility of both metal and support at low temperatures, strong metal–support interaction, and small Ni particle size are important factors for low-temperature CO2 methanation. Li et al. [44] reported the activity scale of Ni/CeO2 > Ni/TiO2 > Ni/SiO2, with additional lower selectivity to methane for the Ni/SiO2 catalyst, attributed to the absence of strong metal–support interaction.

3.2.1. Ni/SiO2 Catalysts

As it is well known, amorphous silicas are largely covalent supports [45] which are usually supposed to be essentially non-dispersive with respect to other more ionic materials. Depending on the preparation method and type of materials, amorphous silicas have high morphological stability towards sintering and loss of surface area up to 873 K or more [46,47]. Silica is active in reversible adsorption, including of CO2, with an important role of surface silanol groups as hydrogen bond donors [48], but without any Lewis acidity and basicity [49]. This is the reason for poor dispersing ability towards supported phases, as well as for poor adsorption of CO2 without the formation of carbonate-like species.
Preparations of Ni/SiO2 catalysts have been described by several authors. In general, it has been found that supporting nickel salts on amorphous silica followed by calcination gives rise to segregated NiO particles, although quite a small part of Ni2+ can be highly dispersed on ion exchange sites, i.e., exchanging protons in silanol group nests [50,51,52]. After reduction, such Ni/SiO2 catalysts usually mainly present large Ni metal particles with poor dispersion [53] and poor interaction with the support, although the copresence of dispersed unreduced nickel has also been reported [52]. Stability towards sintering of the Ni/SiO2 catalyst is limited [27], although the use of meso/microporous silica supports and opportune preparation methods can improve it [54].
The CO2 hydrogenation reaction over Ni/SiO2 catalysts has been the object of several studies. According to Gac et al. [55], small particles (d = 15.8 nm) of nickel on silica can show very high catalytic activity and selectivity to methane at low temperatures, while selectivity to CO increases with increased particle size and reaction temperature above 673 K. Wu et al. [56] reported a different result, showing that low-loading (0.5 wt% Ni/SiO2) catalysts essentially work as a rWGS catalyst while high-loading and high particle size samples (10% Ni/SiO2) are effective methanation catalysts. Galhardo et al. [57] reported high selectivity to CO for 15 wt% Ni/SiO2 at both low and high temperatures, while near 673 K selectivity to methane is slightly higher than that to CO. These authors report data suggesting that the formation of carbon deposits on Ni metal particles produces Ni carbide species favoring the rWGS activity and deactivating methanation. Li et al. [44] reported lower methane selectivity for Ni/SiO2 with respect to other supported Ni catalysts. Chen et al. [58] investigated CO2 hydrogenation on Ni nanoparticles supported on surface-modified, cage-type mesoporous silica at 773 K, observing high selectivity to CO and low selectivity to methane. Riani et al. [52] reported that low-loading Ni/SiO2 catalysts with particle sizes < 20 nm showed high selectivity to CO and moderate activity for the rWGS reaction, likely mainly due to nickel species dispersed in silica exchange sites, as evidenced by visible spectroscopy. The same authors reported that high-loading Ni/SiO2 catalysts with particle size ~ 100 nm show both methanation and rWGS activity, but there is evident short-term deactivation for methanation (see the decreased methane yield in the decreasing temperature steps at the right of Figure 3), attributed to large, segregated Ni metal particles, covered by a carbon veil.
Hao et al. [35] reported low activity for methanation and high activation energy (125 kJ/mol) for Ni/SiO2 when compared with Ni/CeO2 (Eact near 100 kJ/mol). Guo et al. [60] compared the methanation activity of different 10 wt% nickel catalysts supported on amorphous and mesoporous silica (SBA-15-type and MCM-41-type), showing that they have selectivity to methane approaching 100%, Ni/MCM-41 being definitely less active. In contrast, Bakariza et al. [61] found 15 wt% Ni/MCM-41 more active than Ni/SBA-15. Aziz et al. [62] reported high methanation activity and selectivity for 5% Ni/SiO2 materials, with better performances when using sol–gel prepared mesostructured silica nanoparticles. Yang et al. [63] reported near-equilibrium conversion with 92% methane selectivity on a Ni/MCM-41 silica catalyst, whose activity is improved by zirconium doping. Vogt et al. [64] reported high selectivity to methane for a Ni/SiO2 catalyst with 6 nm Ni particle size but lower selectivity, decreasing with time on stream, for methane for a Ni/SiO2 catalyst with 1.8 nm Ni particle size. Zhu et al. [65] emphasized the progressive deactivation of a 20 wt% Ni/SiO2 catalyst in a 30 h time on stream experiment. However, they also reported a significant increase in activity, selectivity to methane, and stability by further adding, in a second step, additional silica, alumina, or zirconia over Ni/SiO2. Guo and Lu [66] reported that co-impregnation of 1 wt% magnesium improves the activity and stability of 10% Ni/SiO2, but the addition of more Mg results in lower activity and selectivity to methane.

3.2.2. Ni/Al2O3 Catalysts

In contrast to silica, spinel-type aluminas are considered to be ionic solids with high surface, Lewis acidity and basicity, and a strong ability to disperse supported species [67,68]. Thus, Ni dispersion on alumina is very high. Impregnation of γ-Al2O3 with Ni salts followed by calcination results in the formation of a NiAl2O4-like surface layer which is hardly reducible, forming, upon reduction, very small Ni metal clusters [69] which are very selective for CO2 methanation [52,70], as evident from Figure 3. Instead, even at a high Ni loading, the formation of bulk NiAl2O4 does not occur until at least 873 K, given how the reaction between NiO and γ-Al2O3 is still slow and incomplete even at 1273 K [71]. However, the surface spinel layer stabilizes the Ni/γ-Al2O3 catalysts up to high temperatures [72], preventing bulk transformation and loss of surface area.
Surface characterization studies performed by IR spectroscopy of adsorbed CO allowed for the distinguishing of three types of surface Ni metal sites depending on Ni loading on reduced Ni-γ-Al2O3. For a low Ni loading, nearly isolated metal sites are able to adsorb CO in the form of polycarbonyls; at medium loadings, small nickel metal particles adsorb CO on-top (terminal carbonyls) only, and at higher coverages, larger metal particles adsorb CO also as bridging species [73].
In their screening work, among oxide-supported nickel catalysts, Muroyama et al. [39] found Ni/Al2O3 to be not very active at low temperatures (473–523 K) with respect to other oxide-supported Ni catalysts, but to be the most performant and selective to methane at 673 K. In an early study, Van Herwijnen et al. [74] reported a kinetic study of methanation of diluted CO2 (1.4 vol%) in hydrogen, over 33.6 wt% NiO/γ-Al2O3 at 473–503 K, with full conversion at sufficiently high contact times already at 500 K at atmospheric pressure. For γ-alumina supports with 180–220 m2/g, the best activity and selectivity are reported for catalysts with 15–25 wt% Ni [75,76], which roughly corresponds to the coverage of the support by a “geometric monolayer” of surface Ni aluminate species, while at a lower loading, both activity and selectivity decrease in favor of rWGS activity to CO. Otherwise, at an even higher loading, where larger but still strongly support-bonded Ni metal particles are formed, activity decreases, but deactivation is significant, while high selectivity is usually retained [77].
On the other hand, Ni-Al2O3 catalysts prepared by coprecipitation, such as by coprecipitated Ni-Al hydrotalcite, were reported to be definitely more active and more selective than impregnated Ni/γ-Al2O3 at medium-low loadings [78], but may be less active and selective at a near monolayer loading [75]. On the other hand, coprecipitated catalysts with very high Ni loadings (50–90 wt% Ni) were reported by Beierlein [77] to be definitely more active in methanation and stable than high-loading samples (>30 wt% Ni) produced by impregnation. These authors attribute this activity trend to the effect of higher nickel surface area. Indeed, when working with Ni-Al hydrotalcite, the Ni loading is usually very high. When working with a catalyst with a Ni/Al atomic ratio = 3, high selectivity to CO was found at low contact time, with methane selectivity increasing with increasing contact time [79].

3.2.3. Ni/CeO2 Catalysts

Most authors of screening papers report ceria as the best oxide support for nickel catalysts used in CO2 methanation. Hao et al. [35] reported excellent activity and near 100% methane selectivity over 7.5% Ni/CeO2 with 25 m2/g surface area, although the Turn Over Frequency increases by decreasing particle size as a result of decreasing loading. Methanation activity is already significant at 498 K with near 100% selectivity. Similarly, Bian et al. [80] reported activity already in the 473–523 K range for Ni/CeO2 with a near 100% methane selectivity, which somehow depended on ceria particle morphologies: Ni on ceria nanorods appeared to be more active than Ni on ceria nanocubes. Using ceria nanorods, Varvoutis et al. [32] showed high activity already in the 473–573 K range and equilibrium conversion above 623 K with total methane selectivity for samples with 8 to 33% Ni/CeO2 (SBET 72 to 51 m2/g by increasing loading). On the other hand, low-loading Ni/CeO2, with 1–2% nickel, are reported to work as excellent catalysts for the competitive reverse water gas shift reaction to CO [81]. The role of the preparation method has also been emphasized by several authors: Ye et al. [82] reported much higher activity and selectivity of sol–gel prepared Ni/CeO2 than from an analogous catalyst produced by impregnation. As a key for explaining the best performances of Ni/CeO2, the role of the partial reducibility of CeO2, with the detection of Ce3+ ions, has been emphasized [83]. Indeed, among typical oxide carriers, ceria is the most reducible, and also has the ability to absorb hydrogen-forming bulk and surface hydride species [84]. In fact, it has been reported that the Ni/CeO2 system is quite an unstable one, with the ability to dissolve Ni2+ ions in the bulk at higher temperatures and longer times [85]. It is, in fact, known that NiO can dissolve in CeO2 in significant amounts, forming a solid solution with a fluorite structure [86,87], at least for small particle size nanomaterials. A similar deactivation phenomenon with loss of metal surface area is reported, e.g., for Pd/CeO2 for the Water Gas shift reaction [88].
In their review, Xie et al. [89] summarized recent attempts to further improve Ni/CeO2 catalysts for CO2 methanation by modifying the preparation method (constructing more efficient structures and producing more efficient morphologies) and adding additional components, such as alkali, alkali–metal, and rare earth oxides and by alloying Ni with other transition metals. According to these authors, from a practical industrial point of view, challenges remain in developing a durable and low-cost CeO2-based methanation catalyst with considerable low-temperature activity.

3.2.4. Ni/ZrO2 Catalysts

As shown above, most screening studies report high methanation activity for Ni/ZrO2. Chen et al. [90] investigated the activity of 15 wt% Ni/ZrO2 catalysts with predominant monoclinic or tetragonal zirconia support: the monoclinic sample showed lower initial conversion temperature (down to 358 K) and higher basicity than the tetragonal one, both being very selective to methane formation, but with a slight selectivity decrease above 623 K. A similar result was reported by Jia et al. [91] for a 10 wt% Ni/ZrO2 monoclinic + tetragonal zirconia support. Tan et al. [92] used a monoclinic + tetragonal zirconia support with a surface area of 96.6 m2/g to prepare a series of Ni/ZrO2 catalysts. While all catalysts showed selectivity to methane near 100%, the most active appeared to be that with a 6 wt% Ni loading. According to these authors, the intrinsic activity of the catalyst is related to the Ni dispersion and the size of metallic Ni; thus, excellent activity was obtained with highly dispersed particles. According to Vargas et al. [93], really highly dispersed 1 wt% Ni/m-ZrO2 prepared by ultrasound-assisted synthesis is very active and selective in methanation of CO2. Ma et al. confirmed the high activity of Ni/m-ZrO2, more active than a catalyst based on cubic zirconia [94], attributing high activity to the high local electron density of nickel sites. Zhao et al. [95] compared the activities of different 15 wt% Ni/ZrO2 catalysts produced by combustion with different combustion compounds with a sample produced by impregnation of a mainly monoclinic ZrO2 support. The catalyst produced by the urea combustion method (resulting in being tetragonal) was reported to be the most active and stable, although, in the 573–773 K range, the impregnated sample was reported to be even more selective to methane. Concerns about the use of Ni/ZrO2 catalytic materials is due to the potential of NiO to enter monoclinic, tetragonal, and cubic zirconia phases forming solid solutions [96,97] which influence the solid state chemistry of zirconia, apparently favoring the metastable tetragonal phase.

3.2.5. Nickel Catalysts Supported on Other Rare Earth Oxides

In a recent review, Spennati et al. [98] emphasized the role of rare earth oxides in the composition of Ni-based catalysts for CO2 methanation. According to the screening work of Muroyama et al. [39], where pure support effect has been considered for 10 wt% Ni/oxide catalysts, Ni/Y2O3 and, to a lesser extent, Ni/Sm2O3, show the highest methanation activities at 523 K, although they lose methane selectivity above 573 K in favor of CO. Thus, the methanation activity of Ni/Y2O3 is closely similar to that of Ni/Al2O3 and Ni/ZrO2 at 623 K. Other screening studies confirm the high activity of Ni/Y2O3, but no more than Ni/CeO2 [43]. Li et al. [99] confirmed the very high methanation activity of 10 wt% Ni/Y2O3, if prepared in the presence of large amounts of citric acid in particular. Ni/Sm2O3 catalysts were also reported as interesting, with high low-temperature activity and almost total selectivity to methane, mainly when the sample is highly charged (63 wt% Ni) [100], with a slight deactivation attributed to carbonate formation.
Relatively high-loading Ni/La2O3 catalysts [101] are reported to be selective in the methanation of CO2, while low-loading Ni/La2O3 appear to work as rWGS catalysts with high CO selectivity [102]. However, studies report that La2O3 is unstable in reaction conditions, suffering from carbonation or hydration with the formation of oxycarbonate and hydroxide production [103], respectively. On the other hand, 10 wt% Ni supported on the oxycarbonate La2O2CO3 has been reported to display enhanced activity and stability with respect to Ni/La2O3 [104].

3.2.6. Other Ni/Oxide Catalysts

As commented before, most screening studies find a relatively low activity of Ni/TiO2 with respect to other oxide-supported Ni catalysts. In their study, Unwiset et al. [105] found catalytic activity only above 573 K, increasing with the Ni loading, and high selectivity to methane for Ni/TiO2-anatase samples. The XRD study of the catalysts shows that part of the nickel was incorporated in the anatase framework in the solid solution. When looking at solid state stability, the Ni/TiO2-anatase system is considered to be quite an unstable one, since Ni2+ is a promoter of the anatase to rutile phase transformation [106] that usually also occurs at low temperatures with a significant loss of surface area. For example, the TiO2 phase transformation was reported to affect the performance of the Ni/TiO2 catalyst for biomass fast pyrolysis volatiles [107]. On the other hand, Messou et al. [108] found Ni/rutile more active than Ni/anatase, although Ni over a mixed phase support was even more active, showing some kind of synergism between the two TiO2 phases. A few studies also report on the methanation activity of Ni/MgO catalysts [109,110], where high activity and selectivity to methane can also be found. It must be mentioned, however, that the full reciprocal solubility of NiO and MgO (which are isostructural) may represent an element of instability of the system if prepared by impregnation, or of limited efficiency per nickel atom if prepared by coprecipitation.

3.3. Effect of Addition of Doping with a Second Oxide on Ni/γ-Al2O3 and Ni/CeO2

The data summarized above indicate that Ni/CeO2 and Ni/γ-Al2O3 are likely the most attractive systems for CO2 methanation: alumina is usually the preferred support by catalyst producers because of its moderate costs, its well-known chemistry, and its well-proven stability. On the other hand, Ni-Al2O3 catalysts are also already largely used in the industry for CO hydrogenation both to produce Synthetic Natural Gas and to remove CO and CO2 from hydrogen and ammonia synthesis gas, as well as for several hydrogenations in the treatment of hydrocarbon streams [12]. Otherwise, Ni/CeO2 seems to give more active catalysts, in particular at low temperatures. The use of CeO2 is increasingly of interest in heterogeneous catalysis [111].

3.3.1. Effect of a Second Oxide Component on Ni/γ-Al2O3

Silica is a well-known stabilizer of the γ-Al2O3 against phase transition and surface area loss [49,112]. However, the addition of silica in small amounts to γ-Al2O3 shows a slightly negative effect on CO2 hydrogenation activity, although it may have a positive effect when applied in small amounts on the stability of the methanation activity [52]. Liang et al. reported on the effect of Na, K, Mg, Ca, Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn, Zr, La, and Ce on the activity of Ni/γ-Al2O3 [113], observing that Zn only promotes methanation activity at low temperature, while Cr, Fe, Mn, La, Ce promote stability in prolonged CO2 methanation experiments. In contrast, Cai et al. [114] reported a marked promotion effect of zirconia on 12 wt% Ni/γ-Al2O3, in particular when the 3 wt% ZrO2/γ-Al2O3 carrier was prepared by the impregnation–precipitation method. In their recent review, Spennati et al. [98] emphasized the positive role of doping Ni/Al2O3 with rare earth oxide components. Ahmad et al. [115] reported the activity scale Pr > La > Gd > Ce > Eu for 12 wt% Ni/γ-Al2O3 doped with 5 wt% rare earth metal. Guilera et al. [116] co-impregnated alumina with nickel and Ce, La, Sm, Y, and Zr compounds and found La and Ce gave the best catalytic performance, while Sm and Zr showed a lower positive effect. Looking at the optimal amount of Ceria, Kim et al. [117] report maximum activation for methanation with 15 wt% CeO2 for 15 wt% Ni/Al2O3 with 92 m2/g surface area. As for the Ni/La2O3 –γ-Al2O3 system, Garbarino et al. [59] reported 14 wt% La2O3 as the best loading for 13.6 wt% Ni/γ-Al2O3 (17 wt% as NiO). In Figure 3, the positive effect of La2O3 addition to Ni/γ-Al2O3 is evident. This effect was in part attributed to the increased strength of absorption of CO2 in the form of carbonate species over the La2O3-Al2O3 support, with respect to bicarbonate species on pure Al2O3. Riani et al. [118] reported later that the coaddition of silica and lanthana to Ni/γ-Al2O3 can give rise to catalysts with even higher low-temperature methanation activity. A catalyst obtained by a precursor with a weight composition of 16.7% NiO, 37.0% La2O3, 4.6% SiO2, and 41.7% Al2O3 was found to be very active at 523 K.
Beierlein et al. [77] confirmed the negative effect of sodium impurities in the methanation activity and selectivity of Ni/γ-Al2O3. Potassium addition on Ni/γ-Al2O3 was reported by Gandara Loe et al. [119] to significantly increase CO selectivity, finally producing an excellent catalyst for rWGS at T > 873 K. As for alkali earth elements, Liang et al. [120] reported that Ba (1–5 wt% loading) doping of Ni/γ-Al2O3 results in a remarkable selectivity increase, while Sr is less effective and Mg and Ca enhance rWGS reaction, thus reducing methane selectivity. However, Garbarino et al. [121] reported that Ni/Ca-aluminate industrial methane steam reforming catalyst (18 wt% NiO in the catalyst precursor) has excellent activity and selectivity to methane, which are both decreased by 1.8 wt% K2O doping. Garbarino et al. [122] found excellent low-temperature activity for Ca-doped 19 wt% Ni/γ-Al2O3, further enhanced by the addition of V2O5.
A very interesting point concerns the Ni/MgO-Al2O3 system, which can be prepared by coprecipitation starting from Ni-Mg-Al hydrotalcite-type double hydroxides. Liu et al. reported that Ni/MgAl catalysts (36.9 wt% Ni, (Ni + Mg)/Al = 2) exhibit outstanding performance in CO2 methanation, compared with Ni/γ-Al2O3 and Ni/MgO [123], associated with a very small Ni metal particle size, allowing better low-temperature CO2 methanation activity. Very high CO2 methanation activity and selectivity, higher than that of Ni/γ-Al2O3, was recently reported by Fasolini et al. [124] for ex-hydrotalcite 27 wt% Ni/MgO-Al2O3.
Another element that has been reported as a good activator for methanation on Ni/γ-Al2O3 is manganese. Zhao et al. [125] reported maximum activation of 15 wt% Ni/γ-Al2O3 by adding 1.7 wt% Mn. These authors attributed the activation effect to the increased number of CO2 adsorption sites, the improved Ni dispersion, and the weakened interactions between the nickel species and the support. In good agreement, Vrijburg et al. [126] reported significantly higher methanation rates and CH4 selectivities for Mn-promoted Ni/γ-Al2O3. The optimal Mn/Ni atomic ratio was found to be 0.25, showing metallic Ni particles and small oxidic Mn2+ species. According to these authors, Mn addition improves Ni dispersion and enhances Ni2+ reducibility by weakening the interaction between the Ni-oxide precursor and the support. Wu et al. [127] reported on the excellent methanation activity of a catalyst with 20 wt% Ni/Al2O3-ZrO2 doped with 6.7% Mn. All these studies conclude that Mn oxide species are dispersed on the support surface, favoring nickel dispersion.

3.3.2. Effect of a Second Oxide Component on Ni/CeO2

A number of studies reported on the further promotion of Ni/CeO2 for CO2 methanation. The use of CeO2-ZrO2 solid solutions is a common way to stabilize ceria. Pan et al. [128] found Ni/Ce0.5Zr0.5O2 to definitely be more active than Ni/γ-Al2O3 and attributed to the stronger basicity of the Ce0.5Zr0.5O2 support, having more and stronger CO2 adsorption sites than γ-Al2O3. Looking at different CeO2-ZrO2 support compositions, Ullah et al. [129] found maximum activity for 5 wt% Ni catalysts when the support composition is Ce0.7Ni0.3O2. A number of studies reported the promoting effect of rare earth ions on the activity of Ni/CeO2, such as Eu3+ [130], Sm3+ [131], Y3+ [132], the couples La3+ -Sm3+, La3+-Pr3+, and La3+ -Mg2+ [133], etc. Looking at the effect of alkali earth metals, Liu et al. [134] found an increase in the methanation activity of Ni/CeO2 by adding Ca2+ and Mg2+, with a more limited effect for Sr2+ and a deactivation effect for Ba2+.

3.4. Effect of Alloying

Several studies report bimetallic Ni-based catalysts being more active than pure nickel ones in CO2 methanation. This subject has been reviewed recently by Tsiotsias et al. [26]. The addition of small amounts of noble metals (Pt, Ru, Rh, Pd, and Re) to Ni/Al2O3 provides several advantages, such as excellent activity at low temperatures, high reducibility, and stability once oxidized [135], producing even new sites or by forming an alloy with nickel as evidenced for Ru and Pt or Pd, respectively. However, the use of very expensive noble metals to activate nickel may make sense only if they are added in extremely small amounts, due to their activity even at small loadings over oxide support. According to Pan et al. [136] the addition of 0.2 wt% Ru causes a 3-fold rate enhancement at low temperature for selective methanation of CO2 relative to the monometallic 20 wt% Ni/CeO2 and to 0.2 wt% Ru/CeO2.
The use of less expensive transition metal to increase Ni activity seems to be more interesting. As already cited, according to Pandey [31], the moderate activity of pure nickel is significantly increased by the addition of small amounts of iron. For example, the addition of 1 wt% cobalt was reported to significantly increase the low-temperature activity of 15 wt% Ni/MgO-Al2O3 catalysts [137]. Similarly, the addition of 4.7 wt% iron significantly increases the low-temperature activity of a 20 wt% Ni/MgO-Al2O3 catalyst [138]. According to Xu et al. [139], the additions of Fe, Co, and, to a lower extent, Mn on Ni/Ce0.8Zr0.2 enhance activity, while copper addition decreases both activity and selectivity to CH4. Over low-loading Ni/La2O3, copper addition increases rWGS activity and selectivity to CO [102].

4. Discussion

The data summarized above show that, essentially, all Ni-containing catalysts found activity in CO2 hydrogenation. In Table 1, the most relevant criteria for selecting optimal Ni-based CO2 methanation catalysts are proposed. Catalysts where most of the nickel species remain unreduced, because they strongly interact with the oxide support, usually have predominant rWGS activity to CO, while those where sufficient nickel can be reduced to the metallic state have at least good initial activity to methanation. It seems that good and stable methanation activity is associated mainly with very small nickel metal particles associated with very high nickel dispersion, which can be stabilized mainly on ionic oxides. In fact, for large particle Ni catalysts, fast deactivation by carbon deposition is usually observed, which may also result in switching toward higher CO selectivity by rWGS activity, likely due to carbon or nickel carbide species.
Methanation occurring on unsupported Ni particles and Ni/SiO2 should fully occur at the nickel metal surface, where carbide or carbon intermediates are reasonably formed and active in methanation [140]. However, mechanisms via oxygenate intermediates (carboxylate, carbonate, formate, formyl, and methoxy) can also occur at nickel surfaces [141]. On ionic oxides, where CO2 adsorbs and forms different types of carbonate and bicarbonate species [23], a more important role of the support in the methanation mechanism is plausible [142].
Very high low-temperature methanation activity is obtained on small Ni metal particles supported on semiconducting and slightly reducible metal oxides characterized by medium surface basicity, such as, in particular, for Ni/CeO2. On the other hand, it seems also quite evident that the addition of basic, semiconducting, and moderately reducible components to alumina (a non-conducting, non-reducible, and mainly acidic support) is sufficient to significantly increase the methanation activity at 473–523 K, up to values comparable to those obtained with (doped) Ni/CeO2, and with Ru/Al2O3 as well. Such components to be added to alumina support include, e.g., cerium or manganese oxides.
The activating effect of such an intermediate level of basicity is usually attributed to the resulting ability of the support to adsorb CO2 in the form of carbonate species that, thanks to the spillover of hydrogen from nickel metal particles, is at the origin of a oxygenate intermediate-based hydrogenation mechanism [143]. For example, the activating effect of the addition of La2O3 to Ni/γ-Al2O3 has been attributed in part to the increased strength in the adsorption of CO2 [59]; as shown in Figure 4, the adsorbed species of CO2 are fully modified.
On γ-Al2O3, two slightly different forms of bicarbonates are observed (Figure 4, bands at 1644 cm−1; O-C-O asymmetric stretching, 1484, 1442 cm−1, two different O-C-O symmetric stretchings, 1236 cm−1, COH deformation mode), whose band decreased by the addition of La2O3 until almost disappearing, while bands due to two forms of carbonates are formed (COx stretchings at 1680 and 1380 cm−1, likely bridging carbonates, and at 1540 and 1435 cm−1, likely monodentate carbonates). However, slight effects in modifying the electronic properties of Ni metal particles by increasing support basicity as well as semiconducting or redox species cannot be excluded [59].
As shown above, the addition of small amounts of another metal, e.g., cobalt or ruthenium, can also produce a significant increase in activity at low temperatures.
On the other hand, to develop industrial catalysts, besides (initial) activity, other factors must be considered and are even more relevant. In their work concerning high-temperature methanation of CO-rich syngases, Nguyen et al. [13] reported that catalysts must be stable against metal particles and support sintering, formation of carbon and gum residues, as well as the formation of nickel tetracarbonyl. Such a volatile compound can favor Ni particle growth through its formation and decomposition. The need to avoid Ni(CO)4 formation drives the limiting of nickel loading and the use of sufficiently high reaction temperatures, e.g., >573 K, where such gaseous compounds cannot form. Thus, for this reason, the very low-temperature activity of the catalyst cannot be so relevant.
It seems that the limitation of the formation of carbon residue and gum can be obtained using catalysts allowing the production of small to very small nickel metal particles in reaction conditions. This is apparently obtained using high-dispersing ionic oxide supports and a moderate loading of nickel.
Another point is certainly the solid state stability of the support and its reactivity towards nickel species. In fact, most metal oxides allow solubility of Ni2+, forming mixed oxides. The penetration of nickel in the bulk of the support, with the help of the slight oxidizing properties of the reactant CO2 and of the product H2O, would result in the consumption of the active phase (metallic nickel) and in the deterioration of the morphological properties of the support. From this point of view, γ-Al2O3 seems to be optimal, taking into account that it tends to be covered by a NiAl2O4-like surface layer upon the preparation of the catalyst that protects the bulk from further Ni penetration. On the other hand, for decades, the use of Ni/Al2O3 catalysts in several other hydrogenation processes in relevant industrial conditions has been a confirmation of the stability of this system. In fact, as reported recently by Spennati et al., commercial Ni catalysts for COx methanation seem to be exclusively based on Ni/Al2O3 [98]. This might be due to its cheapness and well-known chemistry of alumina, coupled with its commercial availability, stability in mild conditions, and versatility, which make it preferred by the industry, e.g., as hydrotreating catalysts, as reported years ago [144]. It must also be taken into consideration that even though coprecipitated Ni/Al2O3-based catalysts may be a little more active than impregnated ones at least at initial steps, the less complete use of nickel in coprecipitated samples and the likely simpler preparation of catalysts by co-impregnation of commercial alumina with nickel and dopants can have advantages (also from the environmental point of view) with respect to more complex coprecipitation procedures.

5. Conclusions

A detailed analysis of the scientific literature data concerning catalysts for CO2 methanation allowed us to evidence the main features needed for such an application. It seems that the necessary features of the best-performing oxide-supported nickel catalyst for CO2 methanation are as follows:
  • The use of an ionic support that allows good nickel dispersion, resulting in very small nickel metal particles. Such small metal particles would be very active toward methanation and would not produce many carbon residues and gums;
  • The use of a support and/or surface additives that produce a medium basicity, allowing a medium-strong adsorption of CO2;
  • A medium nickel loading, which allows the free support’s geometric surface to be covered with optimal density by small nickel metal particles, without the production of larger Ni crystals and allows sufficient active sites for CO2 adsorption on the support surface;
  • Working at temperatures where Ni(CO)4 formation is impossible (e.g., >573 K).
Taking into account the already significant use of Ni/Al2O3 systems for catalytic hydrogenations, its ascertained long-term stability, and the common use of γ-alumina as a catalyst support, the application of 15–20 wt% nickel over a typical γ-Al2O3 (200 m2/g) with small amounts of slightly reducible and basic promoters, such as Ce4+ and Mnn+, or/and of simply mild basic components, such as Mg2+, and maybe small amounts of an alloying metal, such as cobalt or iron, can be recommended. The most likely alternatives are based on Ni/CeO2 with additives, whose long-term stability, however, should be verified.

Author Contributions

Conceptualization, methodology, and writing—review and editing, G.B., E.S., P.R. and G.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded under the National Recovery and Resilience Plan (NRRP), Mission 4 Component 2 Investment 1.3—Call for tender No. 1561 of 11.10.2022 of Ministero dell’Uni-versità e della Ricerca (MUR); funded by the European Union—NextGenerationEU Project title “Network 4 Energy Sustainable Transition—NEST” (Project code PE0000021).

Data Availability Statement

No new data were created or analyzed in this review study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Götz, M.; Lefebvre, J.; Mörs, F.; Koch, A.M.; Graf, F.; Bajohr, S.; Reimert, R.; Kolb, R. Renewable Power-to-Gas: A Technological and Economic Review. Renew. Energy 2016, 85, 1371–1390. [Google Scholar] [CrossRef] [Green Version]
  2. Vega Puga, E.; Moumin, G.; Neumann, N.C.; Roeb, M.; Ardone, A.; Sattler, C. Holistic View on Synthetic Natural Gas Production: A Technical, Economic and Environmental Analysis. Energies 2022, 15, 1608. [Google Scholar] [CrossRef]
  3. Tripodi, A.; Conte, F.; Rossetti, I. Carbon Dioxide Methanation: Design of a Fully Integrated Plant. Energy Fuels 2020, 34, 7242–7256. [Google Scholar] [CrossRef]
  4. Bolt, A.; Dincer, I.; Agelin-Chaab, M. A critical review of synthetic natural gas production techniques and technologies. J. Nat. Gas Sci. Eng. 2020, 84, 103670. [Google Scholar] [CrossRef]
  5. Mazheika, A.; Wang, Y.G.; Valero, R.; Viñes, F.; Illas, F.; Ghiringhelli, L.M.; Levchenko, S.V.; Scheffler, M. Artificial-intelligence-driven discovery of catalyst genes with application to CO2 activation on semiconductor oxides. Nat. Commun. 2022, 13, 419. [Google Scholar] [CrossRef] [PubMed]
  6. 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, 2358–2368. [Google Scholar] [CrossRef]
  7. Stangeland, K.; Kalai, D.; Li, H.; Yu, Z. CO2 methanation: The effect of catalysts and reaction conditions. Energy Proced. 2017, 105, 2022–2027. [Google Scholar] [CrossRef]
  8. Bertucco, A.; Fermeglia, M. 50 years of Soave Equation of State (SRK): A source of inspiration for chemical engineers. Fluid Phase Equilibria 2023, 566, 113678. [Google Scholar] [CrossRef]
  9. Torrente-Murciano, L.; Mattia, D.; Jones, M.D.; Plucinski, P.K. Formation of hydrocarbons via CO2 hydrogenation—A thermodynamic Study. J. CO2 Util. 2014, 6, 34–39. [Google Scholar] [CrossRef]
  10. Schmider, D.; Maier, L.; Deutschmann, O. Reaction Kinetics of CO and CO2 Methanation over Nickel. Ind. Eng. Chem. Res. 2021, 60, 5792–5805. [Google Scholar] [CrossRef]
  11. Liao, L.; Chen, L.; Ye, R.P.; Tang, X.; Liu, J. Robust nickel silicate catalysts with high Ni loading for CO2 methanation. Chem. Asian J. 2021, 16, 678–689. [Google Scholar] [CrossRef] [PubMed]
  12. Busca, G. Metal catalysts in hydrogenation and dehydrogenation reactions. In Heterogeneous Catalytic Materials; Busca, G., Ed.; Elsevier: Amsterdam, The Netherlands, 2014; pp. 297–343. [Google Scholar]
  13. Nguyen, T.T.M.; Wissing, L.; Skjøth-Rasmussen, M.S. High temperature methanation: Catalyst considerations. Catal. Today 2013, 215, 233–238. [Google Scholar] [CrossRef]
  14. Frontera, P.; Macario, A.; Ferraro, M.; Antonucci, P.L. Supported Catalysts for CO2 Methanation: A Review. Catalysts 2017, 7, 59. [Google Scholar] [CrossRef]
  15. Ashok, J.; Pati, P.; Hongmanorom, P.; Tianxi, Z.; Junmei, C.; Kawi, S. A review of recent catalyst advances in CO2 methanation processes. Catal. Today 2020, 356, 471–489. [Google Scholar] [CrossRef]
  16. Tan, C.H.; Nomanbhay, S.; Shamsuddin, A.H.; Park, Y.K.; Hernández-Cocoletzi, H.; Show, P.L. Current Developments in Catalytic Methanation of Carbon Dioxide—A Review. Front. Energy Res. 2022, 9, 795423. [Google Scholar] [CrossRef]
  17. Pham, C.Q.; Bahari, M.B.; Kumar, P.S.; Ahmed, S.F.; Xiao, L.; Kumar, S.; Qazaq, A.S.; Siang, T.J.; Tran, H.T.; Islam, A.; et al. Carbon dioxide methanation on heterogeneous catalysts: A review. Environ. Chem. Lett. 2022, 20, 3613–3630. [Google Scholar] [CrossRef]
  18. Abild-Pedersen, F.; Andersson, M.P. CO adsorption energies on metals with correction for high coordination adsorption sites—A density functional study. Surf. Sci. 2007, 601, 1747–1753. [Google Scholar] [CrossRef]
  19. Goldberger, W.M.; Othmer, D.F. Kinetics of Nickel Carbonyl Formation. Ind. Eng. Chem. Process Des. Dev. 1963, 2, 202–209. [Google Scholar] [CrossRef]
  20. Calatayud, M.; Markovits, A.; Menetrey, M.; Mguig, B.; Minot, C. Adsorption on perfect and reduced surfaces of metal oxides. Catal. Today 2003, 85, 125–143. [Google Scholar] [CrossRef]
  21. Burghaus, U. Surface chemistry of CO2—Adsorption of carbon dioxide on clean surfaces at ultrahigh vacuum. Progr. Surf. Sci. 2014, 89, 161–217. [Google Scholar] [CrossRef]
  22. Monachino, E.; Greiner, M.; Knop-Gericke, A.; Schlögl, R.; Dri, C.; Vesselli, E.; Comelli, G. Reactivity of Carbon Dioxide on Nickel: Role of CO in the Competing Interplay between Oxygen and Graphene. J. Phys. Chem. Lett. 2014, 5, 1929–1934. [Google Scholar] [CrossRef] [PubMed]
  23. Busca, G.; Lorenzelli, V. Infrared spectroscopic identification of species arising from reactive adsorption of carbon oxides on metal oxide surfaces. Mater. Chem. 1982, 7, 89–126. [Google Scholar] [CrossRef]
  24. Ramis, G.; Busca, G.; Lorenzelli, V. Low-temperature CO2 adsorption on metal oxides: Spectroscopic characterization of some weakly adsorbed species. Mater. Chem. 1991, 29, 425–435. [Google Scholar] [CrossRef]
  25. Strucks, P.; Failing, L.; Kaluza, S. A Short Review on Ni-Catalyzed Methanation of CO2: Reaction Mechanism, Catalyst Deactivation, Dynamic Operation. Chem. Ing. Tech. 2021, 93, 1526–1536. [Google Scholar] [CrossRef]
  26. Tsiotsias, A.I.; Charisiou, N.D.; Yentekakis, I.V.; Goula, M.A. Bimetallic Ni-Based Catalysts for CO2 Methanation: A Review. Nanomaterials 2021, 11, 28. [Google Scholar] [CrossRef] [PubMed]
  27. Abboud, M.; Alnefaie, R.; Alhanash, A. Unsupported and silica-supported nickel nanoparticles: Synthesis and application in catalysis. J. Nanopart. Res. 2022, 24, 21. [Google Scholar] [CrossRef]
  28. Teng, C.; He, J.; Zhu, L.; Ren, L.; Chen, J.; Hong, M.; Wang, Y. Fabrication and Characterization of Monodisperse Magnetic Porous Nickel Microspheres as Novel Catalysts Nanoscale. Res. Lett. 2015, 10, 384. [Google Scholar] [CrossRef] [Green Version]
  29. Lucchini, M.A.; Testino, A.; Ludwig, C.; Kambolis, A.; El-Kazzi, A.; Cervellino, A.; Riani, P.; Canepa, F. Continuous synthesis of nickel nanopowders: Characterization, process optimization, and catalytic properties. Appl. Catal. B Environ. 2014, 156–157, 404–415. [Google Scholar] [CrossRef]
  30. 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]
  31. Pandey, D.; Ray, K.; Bhardwaj, R.; Bojja, S.; Chary, K.V.R.; Deo, G. Promotion of unsupported nickel catalyst using iron for CO2 methanation. Int. J. Hydrogen Energy 2018, 43, 4987–5000. [Google Scholar] [CrossRef]
  32. Varvoutis, G.; Lykaki, M.; Stefa, S.; Binas, V.; Marnellos, G.E.; Konsolakis, M. Deciphering the role of Ni particle size and nickel-ceria interfacial perimeter in the low-temperature CO2 methanation reaction over remarkably active Ni/CeO2 nanorods. Appl. Catal. B Environ. 2021, 29, 120401. [Google Scholar] [CrossRef]
  33. Kristiani, A.; Takeishi, K. CO2 methanation over nickel-based catalyst supported on yttria-stabilized zirconia. Catal. Comm. 2022, 165, 106435. [Google Scholar] [CrossRef]
  34. Lee, G.D.; Moon, M.J.; Park, J.H.; Park, S.S.; Hong, S.S. Korean, Raney Ni catalysts derived from different alloy precursors Part II. CO and CO2 methanation activity. Korean J. Chem. Eng. 2005, 22, 541–546. [Google Scholar] [CrossRef]
  35. Hao, Z.; Shen, J.; Lin, S.; Han, X.; Chang, X.; Liu, J.; Li, M.; Ma, X. Decoupling the effect of Ni particle size and surface oxygen deficiencies in CO2 methanation over ceria supported Ni. Appl. Catal. B 2021, 286, 119922. [Google Scholar] [CrossRef]
  36. 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 Chem. 2014, 383–384, 10–16. [Google Scholar] [CrossRef]
  37. Riani, P.; Garbarino, G.; Infantes-Molina, A.; Rodríguez-Castellón, E.; Canepa, F.; Busca, G. Hydrogen from steam reforming of ethanol over cobalt nanoparticles: Effect of boron, impurities. Appl. Catal. A Gen. 2016, 518, 67–77. [Google Scholar] [CrossRef]
  38. Lan, P.-W.; Wang, C.C.; Chen, C.Y. Enhancing the formation of nickel catalysts (111) crystal plane and CO2 methanation reactivity by external magnetic field. J. Taiwan Inst. Chem. Eng. 2020, 116, 188–196. [Google Scholar] [CrossRef]
  39. Muroyama, H.; Tsuda, Y.; Asakoshi, T.; Masitah, H.; Okanishi, T.; Matsui, T.; Eguchi, K. Carbon Dioxide Methanation over Ni Catalysts Supported on Various Metal Oxides. J. Catal. 2016, 343, 178–184. [Google Scholar] [CrossRef] [Green Version]
  40. Hatzisymeon, M.; Petala, A.; Panagiotopoulou, P. Carbon Dioxide Hydrogenation over Supported Ni and Ru Catalysts. Catal. Lett. 2021, 151, 888–900. [Google Scholar] [CrossRef]
  41. Ma, Y.; Liu, J.; Chu, M.; Yue, J.; Cui, Y.; Xu, G. Cooperation Between Active Metal and Basic Support in Ni-Based Catalyst for Low-Temperature CO2 Methanation. Catal. Lett. 2020, 150, 1418–1426. [Google Scholar] [CrossRef]
  42. Gac, W.; Zawadzki, W.; Rotko, M.; Greluk, M.; Słowik, G.; Kolb, G. Effects of support composition on the performance of nickel catalysts in CO2 methanation reaction. Catal. Today 2020, 357, 468–482. [Google Scholar] [CrossRef]
  43. Lee, Y.H.; Ahn, J.Y.; Nguyen, D.D.; Chang, S.W.; Kim, S.S.; Lee, S.M. Role of oxide support in Ni based catalysts for CO2 methanation. RSC Adv. 2021, 11, 17648. [Google Scholar] [CrossRef] [PubMed]
  44. Li, M.; Amari, H.; van Veen, A.C. Metal-oxide interaction enhanced CO2 activation in methanation over ceria supported nickel nanocrystallites. Appl. Catal. B Environ. 2018, 239, 27–35. [Google Scholar] [CrossRef]
  45. Busca, G. The surface acidity of solid oxides and its characterization by IR spectroscopic methods. An attempt at systematization. Phys. Chem. Chem. Phys. 1999, 1, 723–736. [Google Scholar] [CrossRef]
  46. Wu, Y.; Wang, X.; Liu, L.; Zhang, Z.; Shen, J. Alumina-Doped Silica Aerogels for High-Temperature Thermal Insulation. Gels 2021, 7, 122. [Google Scholar] [CrossRef] [PubMed]
  47. Feng, J.; Yan, Y.; Chen, D.; Ni, W.; Yang, J.; Ma, S.; Mo, W. Study of thermal stability of fumed silica based thermal insulating composites at high temperatures. Compos. Part B 2011, 42, 1821–1825. [Google Scholar] [CrossRef]
  48. Busca, G. Location and Accessibility of Hydroxy Groups in Silico-aluminate Porous Materials as Studied by IR Spectroscopy. Curr. Phys. Chem. 2012, 2, 136–150. [Google Scholar] [CrossRef]
  49. Busca, G. Catalytic materials based on silica and alumina: Structural features and generation of surface acidity. Progr. Mat. Sci. 2019, 104, 215–249. [Google Scholar] [CrossRef]
  50. Mile, B.; Stirling, D.; Zammitt, M.A.; Lovell, A.; Webb, M. The location of nickel oxide and nickel in silica-supported catalysts: Two forms of “NiO” and the assignment of temperature-programmed reduction profiles. J. Catal. 1988, 114, 217–229. [Google Scholar] [CrossRef]
  51. Bonneviot, L.; Legendre, O.; Kermarec, M.; Olivier, D.; Che, M. Characterization by UV-vis-NIR Reflectance Spectroscopy of the Exchange Sites of Nickel on Silica. J. Coll. Interf. Sci. 1990, 134, 534–547. [Google Scholar] [CrossRef]
  52. Riani, P.; Spennati, E.; Villa Garcia, M.; Sanchez Escribano, V.; Busca, G.; Garbarino, G. Ni/Al2O3 catalysts for CO2 methanation: Effect of silica and nickel loading. Int. J. Hydrogen Energy 2023, in press. [Google Scholar] [CrossRef]
  53. Xu, Y.; Du, X.; Li, J.; Wang, P.; Zhu, J.; Ge, F.; Zhou, J.; Song, M.; Zhu, W. A comparison of Al2O3 and SiO2 supported Ni-based catalysts in their performance for the dry reforming of Methane. J. Fuel. Chem. Technol. 2019, 47, 199–208. [Google Scholar] [CrossRef]
  54. Lucchini, M.A.; Testino, A.; Kambolis, A.; Proff, C.; Ludwig, C. Sintering and coking resistant core–shell microporous silica–nickel nanoparticles for CO methanation: Towards advanced catalysts production. Appl. Catal. B Envir. 2016, 182, 94–101. [Google Scholar] [CrossRef]
  55. Gac, W.; Zawadzki, W.; Słowik, G.; Sienkiewicz, A.; Kierys, A. Nickel catalysts supported on silica microspheres for CO2 methanation. Micropor. Mesopor. Mat. 2018, 272, 79–91. [Google Scholar] [CrossRef]
  56. Wu, H.C.; Chang, Y.C.; Wu, J.H.; Lin, J.H.; Linc, 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]
  57. Galhardo, T.S.; Braga, A.H.; Arpini, B.H.; Szanyi, J.; Gonçalves, R.V.; Zornio, B.F.; Miranda, C.R.; Rossi, L.M. Optimizing Active Sites for High CO Selectivity during CO2 Hydrogenation over Supported Nickel Catalysts. J. Am. Chem. Soc. 2021, 143, 4268–4280. [Google Scholar] [CrossRef]
  58. Chen, C.S.; Budi, C.S.; Wu, H.C.; Saikia, S.; Kao, H.M. Size-Tunable Ni Nanoparticles Supported on Surface-Modified, Cage-Type Mesoporous Silica as Highly Active Catalysts for CO2 Hydrogenation. ACS Catal. 2017, 7, 8367–8381. [Google Scholar] [CrossRef]
  59. Garbarino, G.; Wang, C.; Cavattoni, T.; Finocchio, E.; Riani, P.; Flytzani-Stephanopoulos, M.; Busca, G. A study of Ni/La-Al2O3 catalysts: A competitive system for CO2 Methanation. Appl. Catal. B Environ. 2019, 248, 286–297. [Google Scholar] [CrossRef]
  60. 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 Materials. Energy Fuels 2018, 32, 3681–3689. [Google Scholar] [CrossRef]
  61. 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]
  62. Aziz, A.A.; Jalil, A.A.; Triwahyono, S.; Mukti, R.R.; Taufiq-Yap, Y.H.; Sazegar, M.R. Highly active Ni-promoted mesostruc-tured silica nanoparticles for CO2 methanation. Appl. Catal. B Environ. 2014, 147, 359–368. [Google Scholar] [CrossRef]
  63. Yang, M.; Zhu, L.; Zhu, X.; Reubroycharoen, P.; Wang, S. CO2 methanation over nickel-based catalysts supported on MCM-41 with in situ doping of zirconium. J. CO2 Util. 2020, 42, 101304. [Google Scholar] [CrossRef]
  64. Vogt, C.; Groeneveld, E.; Kamsma, G.; Nachtegaal, M.; Lu, L.; Kiely, C.J.; Berben, P.H.; Meirer, F.; Weckhuysen, B.M. Unravelling structure sensitivity in CO2 hydrogenation over nickel. Nat. Catal. 2018, 1, 127–134. [Google Scholar] [CrossRef] [Green Version]
  65. Zhu, P.; Chen, Q.; Yoneyama, Y.; Tsubaki, N. Nanoparticle modified Ni-based bimodal pore catalysts for enhanced CO2 methanation. RSC Adv. 2014, 4, 64617. [Google Scholar] [CrossRef]
  66. 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]
  67. Busca, G. The surface of transition aluminas. A critical Review. Catal. Today 2014, 226, 2–13. [Google Scholar] [CrossRef]
  68. Busca, G. Structural, surface and catalytic properties of aluminas. Adv. Catal. 2014, 57, 319–404. [Google Scholar] [CrossRef]
  69. Garbarino, G.; Chitsazan, S.; Phung, T.K.; Riani, P.; Busca, G. Preparation of supported catalysts: A study of the effect of small amounts of silica on Ni/Al2O3 catalysts. Appl. Catal. A Gen. 2015, 505, 86–97. [Google Scholar] [CrossRef]
  70. 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]
  71. Li, C.Y.; Zhang, H.J.; Chen, Z.Q. Reaction between NiO and Al2O3 in NiO/-Al2O3 catalysts probed by positronium. Atom. Appl. Surf. Sci. 2013, 266, 17–21. [Google Scholar] [CrossRef]
  72. Li, M.; Fang, S.; Hu, Y.H. Self-stabilization of Ni/Al2O3 Catalyst with a NiAl2O4 Isolation Layer in Dry Reforming of Methane. Catal. Lett. 2022, 152, 2852–2859. [Google Scholar] [CrossRef]
  73. García-Diéguez, M.; Finocchio, E.; Larrubia, M.A.; Alemany, L.J.; Busca, G. Characterization of alumina-supported Pt, Ni and PtNi alloy catalysts for the dry reforming of methane. J. Catal. 2010, 274, 11–20. [Google Scholar] [CrossRef]
  74. Van Herwijnen, T.; Van Doesburg, H.; De Jong, W.A. Kinetics of the methanation of CO and CO2 on a nickel catalyst. J. Catal. 1973, 28, 391–402. [Google Scholar] [CrossRef]
  75. Aksoylu, A.E.; Onsan, Z.L. Hydrogenation of carbon oxides using coprecipitated and impregnated Ni/Al2O3 catalysts. Appl. Catal. A Gen. 1997, 164, 1–11. [Google Scholar] [CrossRef]
  76. Rahmani, S.; Rezaei, M.; Meshkani, F. Preparation of highly active nickel catalysts supported on mesoporous nanocrystalline γ-Al2O3 for CO2 methanation. J. Ind. Eng. Chem. 2014, 20, 1346–1352. [Google Scholar] [CrossRef]
  77. Beierlein, D.; Häussermann, D.; Pfeifer, M.; Schwarz, T.; Stöwe, K.; Traa, Y.; Klemm, E. Is the CO2 methanation on highly loaded Ni-Al2O3 catalysts really structure sensitive? Appl. Catal. B Environ. 2019, 247, 200–219. [Google Scholar] [CrossRef]
  78. Lei, H.; Lin, Q.; 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]
  79. Marocco, P.; Morosanu, E.A.; Giglio, E.; Ferrero, D.; Mebrahtu, C.; Lanzini, A.; Abate, S.; Bensaid, S.; Perathoner, S.; Santarelli, M.; et al. CO2 methanation over Ni/Al hydrotalcite-derived catalyst: Experimental characterization and kinetic study. Fuel 2018, 225, 230–242. [Google Scholar] [CrossRef]
  80. Bian, Z.; Chan, Y.M.; Yu, Y.; Kawi, S. Morphology dependence of catalytic properties of Ni/CeO2 for CO2 methanation: A kinetic and mechanism study. Catal. Today 2020, 347, 31–38. [Google Scholar] [CrossRef]
  81. Wang, L.; Zhang, S.; Liu, Y. Reverse water gas shift reaction over Co-precipitated Ni-CeO2catalysts. J. Rare Earths 2008, 26, 66–70. [Google Scholar] [CrossRef]
  82. Ye, R.-P.; Li, Q.; Gong, W.; Wang, T.; Razink, J.J.; Lin, L.; Qin, Y.-Y.; Zhou, Z.; Adidharma, H.; Tang, J.; et al. High-performance of nanostructured Ni/CeO2 catalyst on CO2 methanation. Appl. Catal. B Environ. 2019, 268, 118474. [Google Scholar] [CrossRef]
  83. Martin, N.M.; Hemmingsson, F.; Schaefer, A.; Ek, M.; Merte, L.R.; Hejral, U.; Gustafson, J.; Skoglundh, M.; Dippel, A.C.; Gutowski, O.; et al. Structure–function relationship for CO2 methanation over ceria supported Rh and Ni catalysts under atmospheric pressure conditions. Catal. Sci. Technol. 2019, 9, 1644. [Google Scholar] [CrossRef] [Green Version]
  84. Li, Z.; Werner, K.; Chen, L.; Jia, A.; Qian, K.; Zhong, J.Q.; You, R.; Wu, L.; Zhang, L.; Pan, H.; et al. Interaction of Hydrogen with Ceria: Hydroxylation, Reduction, and Hydride Formation on the Surface and in the Bulk. Chem. Eur. J. 2021, 27, 5268–5276. [Google Scholar] [CrossRef] [PubMed]
  85. Deraz, N.M. Effects of heat treatment on physicochemical properties of cerium based nickel system. J. Anal. Appl. Pyrol. 2012, 95, 56–60. [Google Scholar] [CrossRef]
  86. Padi, S.P.; Shelly, L.; Komarala, E.P.; Schweke, D.; Hayun, S.; Rosen, B.A. Coke-free methane dry reforming over nano-sized NiO-CeO2 solid solution after exsolution. Catal. Commun. 2020, 138, 105951. [Google Scholar] [CrossRef]
  87. Wrobel, G.; Sohier, M.P.; D’Huysser, A.; Bonnelle, J.P.; Marcq, J.P. Hydrogenation catalysts based on nickel and rare earth oxides. Part II: XRD, electron microscopy and XPS studies of the cerium-nickel-oxygen-hydrogen system. Appl. Catal. A Gen. 1993, 101, 73–93. [Google Scholar] [CrossRef]
  88. Ebrahimi, P.; Kumar, A.; Khraisheh, M. A review of recent advances in water-gas shift catalysis for hydrogen production. Emerg. Mat. 2020, 3, 881–917. [Google Scholar] [CrossRef]
  89. Xie, Y.; Wen, J.; Li, Z.; Chen, J.; Zhang, Q.; Ning, P.; Chen, Y.; Hao, J. Progress in reaction mechanisms and catalyst development of ceria-based catalysts for low-temperature CO2 methanation. Green Chem. 2023, 25, 130. [Google Scholar] [CrossRef]
  90. Min, C.; Zanglong, G.; Jian, Z.; Fangli, J.; Wei, C. CO2 selective hydrogenation to synthetic natural gas (SNG) over four nano-sized Ni/ZrO2 samples: ZrO2 crystalline phase & treatment impact. J. Energy Chem. 2016, 25, 1070–1077. [Google Scholar]
  91. 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 Environ. 2019, 244, 159–169. [Google Scholar] [CrossRef]
  92. 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]
  93. Vargas, E.; Romero-Sáez, M.; Denardin, J.C.; Gracia, F. The ultrasound-assisted synthesis of effective monodisperse nickel nanoparticles: Magnetic characterization and its catalytic activity in CO2 methanation. New J. Chem. 2016, 40, 7307–7310. [Google Scholar] [CrossRef]
  94. Ma, L.; Ye, R.; Huang, Y.; Ramirez Reina, T.; Wang, X.; Li, C.; Li, X.; Zhang, L.; Fan, M.; Zhang, R.; et al. Enhanced low-temperature CO2 methanation performance of Ni/ZrO2 catalysts via a phase engineering strategy. Chem. Eng. J. 2022, 446, 137031. [Google Scholar] [CrossRef]
  95. 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]
  96. Chandra Bose, A.; Ramamoorthy, R.; Ramasamy, S. Formability of metastable tetragonal solid solution in nanocrystalline NiO–ZrO2 powders. Mater. Lett. 2000, 44, 203–207. [Google Scholar] [CrossRef]
  97. Kondo, H.; Sekino, T.; Kusunose, T.; Nakayama, T.; Yamamoto, Y.; Niihara, K. Phase stability and electrical property of NiO-doped yttria-stabilized zirconia. Mater. Lett. 2003, 57, 1624–1628. [Google Scholar] [CrossRef]
  98. Spennati, E.; Riani, P.; Garbarino, G. A perspective of lanthanide promoted Ni-catalysts for CO2 hydrogenation to methane: Catalytic activity and open challenges. Catal. Today 2023, in press. [Google Scholar] [CrossRef]
  99. Li, Y.; Men, Y.; Liu, S.; Wang, J.; Wang, K.; Tang, Y.; An, W.; Pan, X.; Li, L. Remarkably efficient and stable Ni/Y2O3 catalysts for CO2 methanation: Effect of citric acid addition. Appl. Catal. B Environ. 2021, 293, 120206. [Google Scholar] [CrossRef]
  100. Ilsemann, J.; Sonström, A.; Gesing, T.M.; Anwander, R.; Bäumer, M. Highly Active Sm2O3-Ni Xerogel Catalysts for CO2 Methanation. ChemCatChem 2019, 11, 1732–1741. [Google Scholar] [CrossRef]
  101. Tang, G.; Gong, D.; Liu, H.; Wang, L. Highly loaded mesoporous Ni–La2O3 catalyst prepared by colloidal solution combustion method for CO2 methanation. Catalysts 2019, 9, 442. [Google Scholar] [CrossRef] [Green Version]
  102. Ebrahimi, P.; Kumar, A.; Khraisheh, M. Combustion synthesis of lanthanum oxide supported Cu, Ni, and CuNi nanoparticles for CO2 conversion reaction. Int. J. Hydrogen Energy 2022, in press. [Google Scholar] [CrossRef]
  103. Valsamakis, I.; Garbarino, G. Lanthanum-based catalysts for (bio)ethanol conversion: Effect of preparation method on catalytic performance—hard templating versus hydrolysis. J. Chem. Technol. Biotechnol. 2021, 96, 1116–1124. [Google Scholar] [CrossRef]
  104. Dai, Y.; Xu, M.; Wang, Q.; Huang, R.; Jin, Y.; Bian, B.; Tumurbaatar, C.; Ishtsog, B.; Bold, T.; Yang, Y. Enhanced activity and stability of Ni/La2O2CO3 catalyst for CO2 methanation by metal-carbonate interaction. Appl. Catal. B Environ. 2020, 277, 119271. [Google Scholar] [CrossRef]
  105. Unwiset, P.; Chanapattharapol, K.C.; Kidkhunthod, P.; Poo-arporn, Y.; Ohtani, B. Catalytic activities of titania-supported nickel for carbon-dioxide methanation. Chem. Energy Sci. 2020, 228, 115955. [Google Scholar] [CrossRef]
  106. Hanaor, D.A.H.; Sorrel, C.C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2011, 46, 855–874. [Google Scholar] [CrossRef] [Green Version]
  107. Santamaria, L.; Lopez, G.; Arregi, A.; Amutio, M.; Artetxe, M.; Bilbao, J.; Olazar, M. Stability of different Ni supported catalysts in the in-line steam reforming of biomass fast pyrolysis volatiles. Appl. Catal. B Environ. 2019, 242, 109–120. [Google Scholar] [CrossRef]
  108. Messou, D.; Bernardin, V.; Meunier, F.; Borges Ordoño, M.; Urakawa, A.; Machado, B.F.; Collière, V.; Philippe, R.; Serp, P.; Le Berre, C. Origin of the synergistic effect between TiO2 crystalline phases in the Ni/TiO2-catalyzed CO2 methanation reaction. J. Catal. 2021, 398, 14–28. [Google Scholar] [CrossRef]
  109. Nakayama, T.; Ichikuni, N.; Sato, S.; Nozaki, F. Ni/Mgo catalyst prepared using citric acid for hydrogenation of carbon dioxide. Appl. Catal. A Gen. 1997, 158, 185–199. [Google Scholar] [CrossRef]
  110. Loder, M.A.; Siebenhofer, S.L. The reaction kinetics of CO2 methanation on a bifunctional Ni/MgO catalyst. J. Ind. Eng. Chem. 2020, 85, 196–207. [Google Scholar] [CrossRef]
  111. Aneggi, E.; Boaro, M.; Colussi, S.; de Leitenburg, C.; Trovarelli, A. Chapter 289—Ceria-Based Materials in Catalysis: Historical Perspective and Future Trends. In Handbook on the Physics and Chemistry of Rare Earths; Elsevier: Amsterdam, The Netherlands, 2016; Volume 50, pp. 209–242. [Google Scholar]
  112. Busca, G. Silica-alumina catalytic materials: A critical review. Catal. Today 2020, 357, 621–629. [Google Scholar] [CrossRef]
  113. Liang, C.; Ye, Z.; Dong, D.; Zhang, S.; Liu, Q.; Chen, G.; Li, C.; Wang, Y.; Hu, X. Methanation of CO2: Impacts of modifying nickel catalysts with variable valence additives on reaction mechanism. Fuel 2019, 254, 115654. [Google Scholar] [CrossRef]
  114. 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]
  115. Ahmad, W.; Younis, M.N.; Shawabkeh, R.; Ahmed, S. Synthesis of lanthanide series (La, Ce, Pr, Eu & Gd) promoted Ni/γ-Al2O3 catalysts for methanation of CO2 at low temperature under atmospheric pressure. Catal. Commun. 2017, 100, 121–126. [Google Scholar] [CrossRef]
  116. Guilera, J.; del Valle, J.; Alarcon, A.; Díaz, J.A.; Andreu, T. Metal-oxide promoted Ni/Al2O3 as CO2 methanation micro-size catalysts. J. CO2 Util. 2019, 30, 11–17. [Google Scholar] [CrossRef]
  117. Kim, M.J.; Youn, J.R.; Kim, H.J.; Seo, M.W.; Lee, D.; Go, K.S.; Lee, K.B.; Jeon, S.G. Effect of surface properties controlled by Ce addition on CO2 methanation over Ni/Ce/Al2O3 catalyst. Int. J. Hydrogen Energy 2020, 45, 24595–24603. [Google Scholar] [CrossRef]
  118. Riani, P.; Valsamakis, I.; Cavattoni, T.; Sanchez Escribano, V.; Busca, G.; Garbarino, G. Ni/SiO2-Al2O3 catalysts for CO2 methanation: Effect of La2O3 addition. Appl. Catal. B Envir. 2021, 284, 119697. [Google Scholar] [CrossRef]
  119. Gandara-Loe, J.; Portillo, E.; Odriozola, J.A.; Reina, T.R.; Pastor-Pérez, L. K-Promoted Ni-Based Catalysts for Gas-Phase CO2 Conversion: Catalysts Design and Process Modelling Validation. Front. Chem. 2021, 9, 785571. [Google Scholar] [CrossRef] [PubMed]
  120. Liang, C.; Hu, X.; Wei, T.; Jia, P.; Zhang, Z.; Dong, D.; Zhang, S.; Liu, Q.; Hu, G. Methanation of CO2 over Ni/Al2O3 modified with alkaline earth metals: Impacts of oxygen vacancies on catalytic activity. Int. J. Hydrogen Energy 2019, 44, 8197–8213. [Google Scholar] [CrossRef]
  121. Garbarino, G.; Pugliese, F.; Cavattoni, T.; Busca, G.; Costamagna, P. A Study on CO2 Methanation and Steam Methane Reforming over Commercial Ni/Calcium Aluminate Catalysts. Energies 2020, 13, 2792. [Google Scholar] [CrossRef]
  122. Garbarino, G.; Kowalik, P.; Riani, P.; Antoniak-Jurak, K.; Pieta, P.; Lewalska-Graczyk, A.; Lisowski, W.; Nowakowski, R.; Busca, G.; Pieta, I.S. Improvement of Ni/Al2O3 catalysts for low-temperature CO2 methanation by vanadium and calcium oxide addition. Ind. Eng. Chem. Res. 2021, 60, 655–664. [Google Scholar] [CrossRef]
  123. Liu, J.; Bing, W.; Xue, X.; Wang, F.; Wang, B.; He, S.; Zhang, Y.; Wei, M. Alkaline-assisted Ni nanocatalysts with largely enhanced low-temperature activity toward CO2 methanation. Catal. Sci. Technol. 2016, 6, 3976. [Google Scholar] [CrossRef]
  124. Fasolini, A.; Spennati, E.; Atakoohi, S.E.; Percivale, M.; Busca, G.; Basile, F.; Garbarino, G. A study of CO2 hydrogenation over Ni-MgAlOx catalysts derived from hydrotalcite precursors. Catal. Today 2023, 423, 114271. [Google Scholar] [CrossRef]
  125. Zhao, K.; Li, Z.; Bian, L. CO2 methanation and co-methanation of CO and CO2 over Mn-promoted Ni/Al2O3 catalysts. Front. Chem. Sci. Eng. 2016, 10, 273–280. [Google Scholar] [CrossRef]
  126. Vrijburg, W.L.; Garbarino, G.; Chen, W.; Parastaev, A.; Longo, A.; Pidko, E.A.; Hensen, E.J.M. Ni-Mn catalysts on silica-modified alumina for CO2 methanation. J. Catal. 2020, 382, 358–371. [Google Scholar] [CrossRef]
  127. Wu, Y.; Lin, J.; Xu, Y.; Ma, G.; Wang, J.; Ding, M. Transition Metals Modified Ni−M (M = Fe, Co, Cr and Mn) Catalysts Supported on Al2O3−ZrO2 for Low-Temperature CO2 Methanation. ChemCatChem 2020, 12, 3553–3559. [Google Scholar] [CrossRef]
  128. Pan, Q.; Peng, J.; Sun, T.; Wang, S.; Wang, S. Insight into the reaction route of CO2 methanation: Promotion effect of medium basic sites. Catal. Commun. 2014, 45, 74–78. [Google Scholar] [CrossRef]
  129. Ullah, N.; Tang, R.; Li, Z. Nanostructured Ceria-zirconia Supported Ni Catalysts for High Performance CO2 Methanation: Phase and morphology effect on activity. J. Taiwan Inst. Chem. Eng. 2022, 134, 104317. [Google Scholar] [CrossRef]
  130. Zhang, Z.; Yu, Z.; Feng, K.; Yan, B. Eu3+ doping-promoted Ni-CeO2 interaction for efficient low-temperature CO2 methanation. Appl. Catal. B Environ. 2022, 317, 121800. [Google Scholar] [CrossRef]
  131. Liu, F.; Park, Y.S.; Diercks, D.; Kazempoor, P.; Duan, C. Enhanced CO2 methanation activity of Sm 0.25Ce0.75O2-δ-Ni by modulating the chelating agents-to-metal cation ratio and tuning metal-support interaction. ACS Appl. Mater. Interfaces 2022, 14, 13295–13304. [Google Scholar] [CrossRef]
  132. Sun, C.; Beaunier, P.; La Parola, V.; Liotta, L.F.; Da Costa, P. Ni/CeO2 nanoparticles promoted by yttrium doping as catalysts for CO2 methanation. ACS Appl. Nano Mater. 2020, 3, 12355–12368. [Google Scholar] [CrossRef]
  133. Siakavelas, G.I.; Charisiou, N.D.; Alkhoori, A.; Alkhoori, S.; Sebastian, V.; Hinder, S.J.; Baker, M.A.; Yentekakis, I.V.; Polychronopoulou, K.; Goula, M.A. Highly selective and stable Ni/La-M (M=Sm, Pr, and Mg)-CeO2 catalysts for CO2 methanation. J. CO2 Util. 2021, 51, 101618. [Google Scholar] [CrossRef]
  134. Liu, K.; Xu, X.; Xu, J.; Fang, X.; Liu, L.; Wang, X. The distributions of alkaline earth metal oxides and their promotional effects on Ni/CeO2 for CO2 methanation. J. CO2 Util. 2020, 38, 113–124. [Google Scholar] [CrossRef]
  135. Mihet, M.; Lazar, M.D. Methanation of CO2 on Ni/γ-Al2O3: Influence of Pt, Pd or Rh promotion. Catal. Today 2018, 306, 294–299. [Google Scholar] [CrossRef]
  136. Pan, Y.; Han, X.; Chang, X.; Zhang, H.; Zi, X.; Hao, Z.; Chen, J.; Lin, Z.; Li, M.; Ma, X. Enhanced Low-Temperature CO2 Methanation over Bimetallic Ni–Ru Catalysts. Ind. Eng. Chem. Res. 2023, 62, 4344–4355. [Google Scholar] [CrossRef]
  137. Summa, P.; Świrk Da Costa, K.; Wang, Y.; Samojeden, B.; Rønning, M.; Hu, C.; Motak, M.; Da Costa, P. Effect of cobalt promotion on hydrotalcite-derived nickel catalyst for CO2 methanation. Appl. Mater. Today 2021, 25, 101211. [Google Scholar] [CrossRef]
  138. Huynh, H.L.; Zhu, J.; Zhang, Y.; Shen, Y.; Mekonnen Tucho, W.; Ding, Y.; Yu, Z. Promoting effect of Fe on supported Ni catalysts in CO2 methanation by in situ DRIFTS and DFT study. J. Catal. 2020, 392, 266–277. [Google Scholar] [CrossRef]
  139. Xu, L.; Cui, Y.; Chen, M.; Wen, X.; Lv, C.; Wu, X.; Wu, C.; Miao, Z.; Hu, X. Screening Transition Metals (Mn, Fe, Co, and Cu) Promoted Ni-Based CO2 Methanation Bimetal Catalysts with Advanced Low Temperature Activities. Ind. Eng. Chem. Res. 2021, 60, 8056–8072. [Google Scholar] [CrossRef]
  140. Heine, C.; Lechner, B.A.J.; Bluhm, H.; Salmeron, M. Recycling of CO2: Probing the Chemical State of The Ni(111) Surface During the Methanation Reaction with Ambient-pressure X-ray Photoelectron Spectroscopy. J. Am. Chem. Soc. 2016, 138, 13246–13252. [Google Scholar] [CrossRef]
  141. Huang, M.-X.; Liu, F.; He, C.-C.; Yang, S.-Q.; Chen, W.Y.; Ouyang, L.; Zhao, Y.-J. Interface promoted CO2 methanation: A theoretical study of Ni/La2O3. Chem. Phys. Lett. 2021, 768, 138396. [Google Scholar] [CrossRef]
  142. Sanchez-Escribano, V.; Larrubia Vargas, M.A.; Finocchio, E.; Busca, G. On the mechanisms and the selectivity determining steps in syngas conversion over supported metal catalysts: An IR study. Appl. Catal. A Gen. 2007, 316, 68–74. [Google Scholar] [CrossRef]
  143. Solis-Garcia, A.; Louvier-Hernandez, J.F.; Almendarez-Camarillo, A.; Fierro-Gonzalez, J.C. Participation of surface bicarbonate, formate and methoxy species in the carbon dioxide methanation catalyzed by ZrO2-supported Ni. Appl. Catal. B Environ. 2017, 218, 611–620. [Google Scholar] [CrossRef]
  144. Breysse, M.; Afanasiev, P.; Geantet, C.; Vrinat, M. Overview of support effects in hydrotreating catalysts. Catal. Today 2003, 86, 5–16. [Google Scholar] [CrossRef]
Figure 1. Equilibrium CO2 conversion and CH4 and CO yields (A): at 0.1 MPa; (B): at 2.03 MPa; and (C): at 20.3 MPa. Starting molar ratio H2/CO2 = 5. Calculated using a Gibbs reactor and Soave–Redlich–Kwong equation of state [8] and allowing a calculated phase for the reaction.
Figure 1. Equilibrium CO2 conversion and CH4 and CO yields (A): at 0.1 MPa; (B): at 2.03 MPa; and (C): at 20.3 MPa. Starting molar ratio H2/CO2 = 5. Calculated using a Gibbs reactor and Soave–Redlich–Kwong equation of state [8] and allowing a calculated phase for the reaction.
Energies 16 05304 g001
Figure 2. FE-SEM micrograph of spent unsupported Ni catalyst after methanation experiments showing Ni particles embedded in carbon residues. Modified from ref. [30].
Figure 2. FE-SEM micrograph of spent unsupported Ni catalyst after methanation experiments showing Ni particles embedded in carbon residues. Modified from ref. [30].
Energies 16 05304 g002
Figure 3. Methane yield and CO yield upon CO2 hydrogenation over 17 wt% NiO (fresh catalyst) on pure amorphous silica (17Ni-SiO2), on pure γ-Al2O3 (17Ni-Al2O3), and on 17 wt% La2O3-83 wt% γ-Al2O3. Feed 6 vol% CO2, 30 vol% H2, N2 balance, and GHSV 55,000 h−1. Modified from refs. [52,59].
Figure 3. Methane yield and CO yield upon CO2 hydrogenation over 17 wt% NiO (fresh catalyst) on pure amorphous silica (17Ni-SiO2), on pure γ-Al2O3 (17Ni-Al2O3), and on 17 wt% La2O3-83 wt% γ-Al2O3. Feed 6 vol% CO2, 30 vol% H2, N2 balance, and GHSV 55,000 h−1. Modified from refs. [52,59].
Energies 16 05304 g003
Figure 4. IR spectra of the adsorbed species after CO2 adsorption on γ-Al2O3, 4 wt% La2O3-Al2O3, and 14 wt% La2O3-Al2O3, modified from Ref. [59].
Figure 4. IR spectra of the adsorbed species after CO2 adsorption on γ-Al2O3, 4 wt% La2O3-Al2O3, and 14 wt% La2O3-Al2O3, modified from Ref. [59].
Energies 16 05304 g004
Table 1. Proposed criteria for selecting optimal Ni-based CO2 methanation catalysts.
Table 1. Proposed criteria for selecting optimal Ni-based CO2 methanation catalysts.
Nickel statereduced
Nickel particle sizevery small
Support typeionic, dispersing
not dissolving Ni
Nickel loadingnear a “monolayer” on the support surface
Support surface chemistrycontaining some basicity
Additive(s)favoring nickel dispersion
producing medium basicity
activating nickel by electronic effects
alloying nickel
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Busca, G.; Spennati, E.; Riani, P.; Garbarino, G. Looking for an Optimal Composition of Nickel-Based Catalysts for CO2 Methanation. Energies 2023, 16, 5304. https://doi.org/10.3390/en16145304

AMA Style

Busca G, Spennati E, Riani P, Garbarino G. Looking for an Optimal Composition of Nickel-Based Catalysts for CO2 Methanation. Energies. 2023; 16(14):5304. https://doi.org/10.3390/en16145304

Chicago/Turabian Style

Busca, Guido, Elena Spennati, Paola Riani, and Gabriella Garbarino. 2023. "Looking for an Optimal Composition of Nickel-Based Catalysts for CO2 Methanation" Energies 16, no. 14: 5304. https://doi.org/10.3390/en16145304

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