Next Article in Journal
Improvement of Power Density and COD Removal in a Sediment Microbial Fuel Cell with α-FeOOH Nanoparticles
Previous Article in Journal
Fabrication of Platinum-Decorated NiCo-Layered Double Hydroxide Nanoflowers for Electrocatalytic Ammonia Oxidation Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advancements in Cobalt-Based Catalysts for Enhanced CO2 Hydrogenation: Mechanisms, Applications, and Future Directions: A Short Review

1
Department of Environmental Science and Engineering, Xi’an Jiaotong University, Xi’an 710049, China
2
Shaanxi Institute of Geological Survey Mineral Geological Survey Center, Xi’an 710068, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(9), 560; https://doi.org/10.3390/catal14090560
Submission received: 3 August 2024 / Revised: 21 August 2024 / Accepted: 23 August 2024 / Published: 25 August 2024

Abstract

:
In 2020, China put forward the national energy and economic development strategy goal of “carbon peak and carbon neutrality”; in this context, the hydrogenation of carbon dioxide into clean energy and high-value-added chemicals can effectively alleviate the current environmental pressure. This process represents a crucial avenue for the advancement of green energy and the realisation of a sustainable energy development strategy. Among the efficient catalysts designed for CO2 hydrogenation reactions, transition metal cobalt has garnered extensive attention from researchers due to its relatively abundant reserves and low economic cost. This paper first introduces the thermodynamic process of carbon dioxide hydrogenation and discusses methods to improve the efficiency of the catalytic reaction from a thermodynamic perspective. It then briefly describes the reaction mechanism of cobalt-based catalysts in the carbon dioxide hydrogenation reaction. Based on this understanding, this paper reviews recent research on the application of cobalt-based catalysts in the hydrogenation of carbon dioxide to produce methane, hydrocarbon chemicals, and alcohols. Finally, the methods to improve the catalytic efficiency of these catalysts are discussed, and future research directions are proposed.

1. Introduction

As the world’s population grows, the economy develops, and the standard of living improves, the demand for fossil energy is increasing rapidly. This heightened consumption of fossil energy has led to a significant rise in CO2 emissions into the atmosphere. The insulating effect of CO2 contributes to climate issues such as the greenhouse effect, sea level rise, and acid rain, posing threats to the survival of humans, wildlife, and plants [1]. Historically, nature maintained a balance between the production and consumption of atmospheric CO2 through the absorption and dissolution by land and oceans. However, increased human activities and environmental degradation have disrupted this balance, resulting in a steady rise in atmospheric CO2 concentrations [2,3,4]. Balancing CO2 emissions with absorption while meeting energy needs is a critical challenge. In response, scientists have proposed carbon capture, utilisation, and storage (CCUS) technology pathways, as shown in Figure 1.
CCUS, which encompasses the capture of emitted CO2 from industrial processes or directly from the atmosphere, its subsequent transportation to a suitable storage site, and its permanent underground storage or utilisation for enhanced oil recovery (EOR) or other industrial applications [5,6], represents one of the most effective contemporary strategies for achieving chemical resource utilisation of CO2. At present, it has been extensively documented as a means of retrofitting thermal power generation, petrochemicals, plastic recycling, blast furnace steelmaking, and numerous other industrial sectors [7,8,9]. At this juncture, the cost of carbon sequestration in China is approximately RMB 500/tonne. The low carbon price is insufficient to offset the expense of capture and sequestration, and the conventional CCS technology necessitates optimal geological conditions, which presents a challenge in terms of economic viability [10,11,12]. It is therefore more economically viable to utilise high-purity carbon captured through carbon capture technology as a raw material for the manufacture of industrial products, including plastics, oil, and natural gas [13]. One such avenue of research is the utilisation of CO2 hydrogenation for the production of fuel, which offers the dual benefit of carbon recycling and the alleviation of the global fossil energy crunch.
The CO2 hydrogenation reaction necessitates the availability of a substantial quantity of cost-effective and environmentally benign hydrogen sources. In recent years, the advent of novel energy generation technologies has led to a significant expansion in the utilisation of renewable energy sources, including hydro, geothermal, wind, and solar, for electricity generation. The resulting electricity is employed in the production of hydrogen through the electrolysis of water, thereby facilitating the greening of hydrogen sources. Hydrogen produced from renewable energy sources can be converted via heterogeneous catalysis into high-value-added chemicals (including gasoline, diesel, methane, methanol, ethanol, and other advanced alcohols). This process not only mitigates the ecological and environmental problems caused by excessive CO2 emissions from human activities but also provides a green liquid fuel. High-purity hydrocarbon fuels, upon combustion, produce CO2 and H2O, thus achieving a sustainable carbon cycle of “energy–CO2–energy”. This approach holds significant practical importance and promising application prospects for “green development”, combining environmental protection and energy savings [14].
To date, a large number of homogeneous and heterogeneous catalysts have been reported for CO2 hydrogenation [15]. Under relatively mild reaction conditions, homogeneous catalysts show high activity and selectivity in CO2 hydrogenation, and the reaction mechanism is easy to be studied because of the single active centre; the reaction products are mainly formic acid, methanol, methane, etc. [15,16,17]. However, they also have the disadvantages of difficult separation of catalyst and substrate and short service life. Compared with homogeneous catalysts, heterogeneous catalysts have the advantages of easy recycling, longer service life, and higher stability, which are favourable for industrial applications [15,18]. Therefore, heterogeneous phase catalysis has been more widely studied. Heterogeneous catalytic reactions for CO2 hydrogenation are mainly divided into electrocatalysis, photocatalysis, and thermal catalysis. The first two methods are challenging to achieve industrial production in the near future due to their low efficiency. Therefore, this article focuses exclusively on the thermal heterogeneous catalytic process for CO2 hydrogenation. The CO2 hydrogenation reaction catalysed by multiphase catalysts represents a current research focus, with metals including Cu, Ni, Co, In, and Fe having been widely reported as effective CO2 hydrogenation catalysts. In recent years, numerous reviews on CO2 hydrogenation reactions have been published [19,20,21]. Among them, Co has a better CO2 activation ability at low temperatures, has relatively high stability, and is relatively inexpensive, which has received extensive attention from researchers [22,23,24]. Due to the excessive hydrogenation capacity and lack of RWGS activity, Co catalysts have obvious advantages for methanation on Co catalysts and have better performance at low temperatures compared to Ni and Fe catalysts [23,25,26,27,28]. In recent years, we have found that Co-based catalysts have made a breakthrough in obtaining products such as low-carbon olefins, long-chain alkanes, and high-carbon alcohols through the addition of additives, the establishment of interactions with carriers, or the formation of alloy phases [22,29,30,31]. Meanwhile, the sintering resistance and catalyst lifetime were also improved by forming Co2C active sites or constructing tandem catalysts [32,33,34].
However, few authors have conducted independent reviews on the performance of Co-based catalysts in CO2 hydrogenation reactions. Accordingly, this paper presents a review of the research progress on thermally catalysed CO2 hydrogenation over Co-based multiphase catalysts. Firstly, the CO2 hydrogenation reaction pathways and potential Co-active sites of the catalysts are discussed. Subsequently, the research progress of cobalt-based catalysts for methane, hydrocarbons, and alcohols is summarised in conjunction with the design ideas of the catalysts in terms of additives addition, morphology modification, and size control, among other factors. Finally, this review addresses the challenges associated with the application of CO2 hydrogenation reactions using Co-based catalysts, as well as potential strategies for addressing these challenges. This review offers insights that can inform the design and synthesis of new, highly efficient catalysts.

2. Thermodynamic Considerations in CO2 Hydrogenation

As the highest oxidation state of carbon, CO2 has a high enthalpy of formation (~394 kJ·mol⁻1), making it thermodynamically unfavourable for chemical transformations. Therefore, H2, with a high Gibbs free energy of formation, is usually added to CO2 transformations to supply the necessary energy for the reaction and facilitate the process [35,36,37]. For instance, the enthalpy change for the reaction of CO2 to CO is significantly reduced when H2 is added, as shown in the following equations (Equations (1) and (2)):
CO2 → CO + 1/2O2
(ΔH°298K = 293.0 kJ/mol  ΔG°298K = 257.2 kJ/mol)
CO2 + H2 → CO + H2O
(ΔH°298K = 41.2 kJ/mol  ΔG°298K =28.6 kJ/mol)
CO2 hydrogenation was first proposed by Charles Ross Prichard and Paul Sabatier in the early 20th century. They identified that the high activation energy barrier of the carbon–oxygen double bond in CO2 was the main obstacle to the hydrogenation reaction. The researchers used catalyst design ideas such as optimising metal–carrier interaction (MSI), constructing symmetry-breaking active centres, and regulating the number of oxygen vacancies on the surface, which lowered the thermodynamic energy barriers to carbon dioxide activation, facilitated the carbon dioxide hydrogenation reaction, and regulated the distribution of reaction products [38,39,40]. Over the past century, researchers have explored and developed various catalyst systems to achieve high selectivity for a wide range of products such as methane, methanol, long-chain hydrocarbons, and higher alcohols [22,41], while also investigating possible reaction pathways.
CO2 emitted from industrial activities and H2 produced from clean energy sources follow different reaction pathways on various catalysts, resulting in distinct product compositions (Figure 2). Methane is the most readily obtained product of the reaction. Among the many methanation catalysts, Ni-based catalysts have been extensively studied in industry due to their low cost and availability [42]. However, at high reaction temperatures, Ni particles are prone to sintering or carbon deposition, leading to deactivation.
In the realm of methanol production, copper is the most extensively studied metal. Cu/ZnO/Al2O3 catalysts have been widely implemented in industrial production, yet they also suffer from issues such as sintering or carbon deposits, which shorten the catalysts’ lifespan [43,44]. Consequently, researchers have investigated alternative transition metal and noble metal catalysts for the efficient and stable production of C1 products. Among these, Co catalysts have emerged as a promising option due to their low cost, compatibility with mild reaction conditions, ability to catalyse reactions at low temperatures [23,45], and resistance to high-temperature sintering [26]. Additionally, they offer a longer catalytic lifetime, enhancing their potential for practical applications [33].
The significant potential of Co-based catalysts lies in their ability to promote carbon chain growth. This is particularly relevant in the synthesis of long-chain compounds, which often require the use of methanol or CO as platform molecules for conversion to higher carbon products via Fischer-Tropsch Synthesis (FTS) and methanol-to-hydrocarbon (MTH) processes [38,46]. It is possible to modify FTS catalysts (e.g., Co-based, Fe-based catalysts) or MTH catalysts to combine the reverse water–gas shift (RWGS) reaction with FTS and high-temperature methanol synthesis with MTH, thereby producing C2+ hydrocarbons through the direct hydrogenation of CO2.
The synthesis of high-carbon alcohols is more challenging than that of C2+ hydrocarbons. Since the 1980s, researchers have systematically investigated the catalysts used in the hydrogenation of CO2 to high-carbon alcohols, focusing on four principal metals: Rh, Cu, Mo, and Co [47].
On the technical side, scholars generally agree that the activation (hydrogenation) of carbon dioxide is more susceptible to kinetics rather than thermodynamics [48]. For example, thermodynamically, the number of molecules in a reverse water–gas reaction (RWGS) does not change during the course of the reaction, so increasing the pressure has no effect on the CO yield (Equation (2)).
In contrast, the hydrogenation of CO2 to methanol is a thermodynamically exothermic process (Equation (3)) and is thermodynamically favourable. Furthermore, since the reaction of CO2 hydrogenation to methanol involves a reduction in the number of molecules, high pressure favours the conversion of CO2 and the high selectivity of methanol. Considering the Gibbs free energy, the synthesis of ethanol from CO2 is thermodynamically more favourable, with a free energy change of less than zero (Equation (4)). This suggests that the reaction is more likely to proceed in a forward direction under identical conditions.
Additionally, studies have indicated that the ratio of CO2 to H2 in the feedstock gas significantly influences the reaction efficiency. For CO2 hydrogenation to C1 products, a higher H2 content in the feedstock gas correlates with a higher CO2 conversion rate and product selectivity [38].
CO2 + 3H2 → CH3OH + H2O
(ΔH°298K = −49.5 kJ/mol  ΔG°298K = 3.5 kJ/mol)
CO2 + 3H2 → 1/2 CH3OH + 3/2 H2O
(ΔH°298K = −86.7 kJ/mol  ΔG°298K = −32.4 kJ/mol)
In conclusion, the incorporation of catalysts or alterations to thermodynamic parameters can enhance the efficiency of CO2 hydrogenation. This paper presents a review of Co-based catalysts capable of efficiently converting carbon sources.

3. Mechanism of Reaction Co-Based Catalysts

3.1. Detailed Pathways of CO2 Hydrogenation and Catalyst Behaviour

The researchers have investigated the reaction mechanism based on the characterisation results obtained by in situ XRD, in situ DRIFTS, in situ FT–IR spectroscopy, and other techniques, combined with simulation calculations. They have proposed several potential reaction pathways for CO2 hydrogenation over Co-based catalysts. This paper outlines the most prevalent reaction pathways.
For C1 products such as methane and methanol, three principal pathways are observed [36,49,50,51], as illustrated in Figure 3. One pathway is the formate pathway, where CO2 hydrogenation produces *HCOO intermediates. These intermediates undergo further cleavage of the C–O and C–H bonds to produce *CHxO, which is subsequently hydrogenated to produce CH3OH. Alternatively, the hydrogenation of the C–O bond after it is broken again can produce *CHx, leading to the formation of CH4. Zhang et al. employed a phosphorylation strategy to modify the surface of a structurally stable layered double-hydroxide-derived Co-Al catalyst (CoAlLDH) for the hydrogenation of CO2 to generate methanol [32]. The results of in situ DRIFTS and DFT calculations (Figure 4) demonstrated that the reaction proceeded via the formic acid pathway. The phosphorylated surface oxygen vacancies underwent significant electron transfer, facilitating the direct hydrogenation of the key intermediate *H3CO to methanol by inhibiting the cleavage of the C–O bond in *H3CO.
The second pathway is the RWGS + CO-hydro pathway, where CO2 is first hydrogenated to produce *HOCO intermediates. The C–O bond in the intermediates is then broken to produce CO, which can either be directly desorbed as CO or further hydrogenated to form *CHxO intermediates. These intermediates continue to be hydrogenated, resulting in the formation of C1 products. Lu et al. investigated the kinetic and thermodynamic behaviour of Co/MoS2 single-atom catalysts in the CO2 hydrogenation reaction [52]. Theoretical studies were conducted to elucidate the kinetic and thermodynamic behaviour of CO2 reduction reaction on Co/MoS2 through the utilisation of DFT simulations, transition state calculations, and reaction rate constants for a multitude of potential pathways. The findings indicated that the optimal pathway for the reaction of CO2 with hydrogen is the combination of the RWGS and CO hydrotransformation. The entire reaction pathway can be summarised as *CO2 → *CO → *CHO → *CH2O → *CH2OH and *CH3O → CH3OH, as shown in Figure 5. The third pathway is the direct hydrogenation pathway. In this pathway, a C–O bond in CO2 breaks to form *CO, which either dissociates to CO or undergoes further hydrogenation to form methane.
For the synthesis of C2+ products, a large number of studies have found RWGS to be a prerequisite step in the reaction [53,54,55,56]. The key to the formation of C2+ products is the insertion of *CHx into other intermediates [50,57]. The intermediate *CHx is formed mainly by the conversion of *CO and *HCOO, and the insertion of one or more *CHx into another intermediate results in the formation of a C–C bond, which lengthens the carbon chain [58,59]. Co-based catalysts are excellent FTS catalysts, capable of promoting the growth of the product carbon chain in FTS, but have difficulty in promoting the RWGS reaction. This provides ideas for the design of efficient catalysts and it is common practice to develop tandem catalysts with the aim of producing C2+ hydrocarbons via the RWGS + FTS pathway and high-carbon alcohols via the RWGS + HAS pathway. Canio and colleagues prepared Pd-modified Co/TiO2 catalysts, observing that Pd positively affects RWGS activity. It also enhances the reducibility of cobalt through the spillover effect, thus promoting Co-based CO-FTS catalysts [53].

3.2. Synergistic Effects and Active Sites in CO2 Hydrogenation

In the case of homogeneous catalysts, all hydrogenation reactions are understood to follow a deprotonation mechanism. This involves the active centre of the metal catalyst undergoing redox cycling during the reaction. The reported mechanisms for Co-based homogeneous catalysts include the Co(-I)/Co(I) catalytic cycle, the Co(I)/Co(III) catalytic cycle, and the Co(-I)/Co(0)/Co(+I) catalytic cycle [56,60,61,62]. In contrast, in the case of heterogeneous catalysts, the reaction typically occurs without any alteration to the state of the active centre. The catalytic active sites on Co-based heterogeneous catalysts are complex, with Co0, Co2⁺, Coδ⁺, and Co2C, among others, all considered active sites for CO2 hydrogenation [53,63]. Different intermediate species form at different active sites, and these intermediates determine the reaction pathways that influence the efficiency and products of the CO2 hydrogenation reaction. Have and colleagues investigated the CO2 hydrogenation reaction mechanism of various Co species. The results of the Modulated Excitation (ME) Diffuse Reflectance Infrared Fourier Transform Spectroscopy (DRIFTS) are presented in Figure 6. The CoO active phase follows the hydrogen-assisted pathway, with major intermediates including *CO3, *HCOO, *CHO, and others [64]. In contrast, the metallic Co0 active phase follows a direct dissociation pathway, with *CO as the major intermediate. Kinetically, the direct dissociation pathway is faster, resulting in the production of C1 products, while the hydrogen-assisted pathway is more favourable for the production of C2+ hydrocarbons.
In conventional Fischer-Tropsch Synthesis (FTS) reactions, the formation of Co2C is typically regarded as a significant contributor to catalyst deactivation. However, recent studies have shown that Co2C can facilitate non-dissociative adsorption and CO insertion, thereby enhancing the selectivity of oxygenated compounds in the product [65]. Davis et al. prepared Na-promoted cobalt oxide catalysts (1%Na-20%Co/SiO2) using SiO2 as the carrier and investigated the impact of distinct pretreatment conditions on the Co-active species [66]. When hydrogen was used as the reducing gas fraction, cobalt (III) oxide was fully reduced to metallic cobalt following reduction at 350 °C, with methane identified as the primary product. When the temperature was reduced to 250 °C, Co3O4 was reduced to CoO, leading to a reduction in the hydrogenation ability of the catalyst and a corresponding decrease in methane selectivity. Following the reduction to CO, the active phases of CoO and Co2C were obtained, resulting in a further reduction in CH4 selectivity to 15.3%, while alcohol selectivity increased to 73.2%. Zhang et al. employed the in situ DRIFTS technique to analyse NaCo-Si3N4 catalysts hydrogenated in incoming CO2 and observed that the intermediates of the RWGS pathway, *CO and *CHx, and the in situ formation of the Co2C site promoted the coupling of *CO and *CHx, which improved ethanol selectivity [57]. It is therefore proposed that Co0 is more capable of hydrogenation than Co2⁺ active sites and that the Co2C active phase is key for inhibiting methane formation and enhancing alcohol selectivity.
Synergistic interactions between multiple active sites or alloy phases formed with other metals can significantly affect the efficiency and products of the reaction. Fu et al. explored the effect of metal–oxide interactions on the CO2 hydrogenation reaction by calcining mixed CoAl hydroxides at different temperatures [67]. Their findings indicated that selecting moderate calcination temperatures could effectively balance the Co–metal oxide interactions, thus enabling the optimal production of Co0-Co2⁺ synergism. This moderate Co0-Co2⁺ synergism was found to promote CO2 hydrogenation via the HCOO* pathway, with methane being the main product (with a selectivity of over 98%). The coexistence of Co0 and CoOx is now widely acknowledged as a pivotal characteristic of Co-based catalysts for ethanol generation [59]. Ding et al. conducted density functional theory (DFT) simulations on Co0(111) and CoO(200) surfaces with the objective of deepening their understanding of the mechanism underlying ethanol production over CoOx catalysts. The initial CO2* hydrogenation process on Co(111) was found to require a lower activation potential barrier and to be more active for CO2 conversion than CoO(200). The strong hydrogenation of the Co(0) site promoted the formation of CHx* species more readily, while the CoO site had a more favourable formation of HCOO* species. These species could be coupled with the HCOO* species on the CoOx site to inhibit methane formation and improve ethanol selectivity [68]. Wang and his team employed in situ infrared (in situ FT-IR) and 1H nuclear magnetic resonance (1H NMR) characterisation techniques to examine the CoAlOx catalysts [45]. Their findings indicated that the optimised Co-CoO phase enhanced the efficiency of the insertion of *CHx onto *HCOO intermediates, facilitating the formation of acetate intermediates for ethanol production. Further, it was demonstrated that the CoNi alloy phase on the CoNiAlOx catalyst facilitated the generation of *CHx intermediates. The hydrogenation of *CHx to ethanol was achieved by linking it to *HCOO on the CoO surface [58]. The following year, a team used in situ DRIFTS spectra and XPS to characterise the Co@Si0.95 catalyst. They showed that *CH3O on Co@Si is an intermediate in CH3OH formation. The irreducible oxygen on the surface of CoO prevents the breaking of the C–O bond in *CH3O, allowing for the hydrogenation of more *CH3O to form CH3OH rather than CH4 [69].
Given that Co0 plays a role in the dissociative activation of CO2 and H2, and Coδ⁺ facilitates the non-dissociative activation of C–O, the simultaneous presence of Co0 and Coδ⁺-active sites is considered a beneficial approach for inhibiting methanation and promoting carbon chain growth. This approach aims to improve the performance of catalytic synthesis of higher alcohol [70]. An et al. investigated the catalytic properties of CoGaxAl2−xO4/SiO2 (x = 0, 0.5, 1.0, 1.5, 2.0) catalysts for CO2 hydrogenation to ethanol [71]. Their results demonstrated that the formation of Co0-Coδ+-active sites on the catalyst surface, due to the strong interaction between Ga2O3 and Co, facilitated ethanol synthesis by coupling *CHx and CO intermediates. This led to an ethanol selectivity of 20.1% over the optimal catalyst (270 °C, 3.0 MPa).
In conclusion, the active sites of Co exhibit varying degrees of activation capacity for H2 and CO2, and are capable of forming distinct stable intermediates. The synergistic effect between different sites allows for the regulation of CO2 hydrogenation activity and product selectivity by constructing multi-site catalysts [72].

4. Optimisation of Co-Based Catalysts

Co is one of the most reactive Group VIII metals for hydrogenation, exhibiting excellent low-temperature hydrogenation capabilities. It has been extensively studied for applications in Fischer-Tropsch synthesis [46]. Co-based catalysts also have a strong adsorption activation capacity for CO2, with complex active centres that facilitate hydrogenation on Co0, CoO, Co2C, and Co2N species [64,73,74]. The component ratios and structures of the different Co species determine the CO2 conversion rate and the types of products formed [22].
The strong hydrogenation capacity of Co catalysts typically results in methane as the dominant product of CO2 hydrogenation. To achieve the selective formation of long-chain hydrocarbons or oxygenated products [75], it is necessary to weaken this hydrogenation capacity, control the degree of CO2 reduction, and promote C–C bond coupling [38]. According to the literature, Co catalysts can aid in carbon chain growth by forming alloys with other metals or interacting with supports [53,76,77]. Additionally, the incorporation of additives can alter the surface acidity, alkalinity, strength, and hardness of the catalysts, as well as the surface and electronic structures, thereby affecting the CO2 hydrogenation reaction rate and product distribution by adjusting the reaction pathways [22].
For example, in a K-Co0 catalyst, the addition of the alkali metal K as an additive increased the electron density around Co0 and promoted CO2 adsorption, significantly enhancing the CO2 conversion rate [71]. Using in situ characterisation techniques to analyse the microstructure and surface active species of the catalysts in their working state, it was found that CO2 on pure-phase Co catalysts was adsorbed and dissociated to CO*, which then either desorbed directly as CO or was hydrogenated to CH4. Conversely, some CO2 adsorbed on the K-Co0 surface was hydrogenated to form HCOO* intermediates, which were further reduced through a C–C coupling step to produce C2+ products.
Developments in materials chemistry have led to an interest in new catalysts such as metal–organic frameworks (MOFs) and bifunctional catalysts [78,79,80]. Among these, bifunctional catalysts, which possess two different kinds of active centres, can facilitate a variety of synergistic reactions simultaneously. They exhibit high stability and the ability to precisely regulate product selectivity [81,82], making them a focus of research and expanding new avenues in the study of CO2 hydrogenation catalysts.

4.1. Detailed Pathways of CO2 Hydrogenation and Catalyst Behaviour

Methane is the main component of natural gas, with a hydrogen content of up to 25%, making it an advanced gaseous fuel and an important raw material for the production of chemical products such as hydrogen energy, carbon black, carbon monoxide, acetylene, hydrocyanic acid, and formaldehyde. In addition, the Gibbs free energy of the carbon dioxide methanation reaction is less than zero (Equation (5)), so methane at room temperature is relatively stable. It can be used as a medium for storing hydrogen to effectively solve the problem of hydrogen being difficult to store, and can then transport it to the industrial development pain point [83,84]. The methanation reaction, also known as the Sabatier reaction [51,84], requires CO2 activation at high temperatures due to the structural stability of the linear, non-polar molecule. However, thermodynamically, methanation is a highly exothermic reaction with a decreasing gas volume (Equation (5)), which can occur spontaneously at room temperature [79,84,85].
CO2 + 3H2 → CH4 + H2O(g)
(ΔH°298K = −165 kJ/mol  ΔG°298K = −113.6 kJ/mol)
High temperatures promote the endothermic reverse water–gas shift (RWGS) reaction, producing CO as a by-product. Therefore, the reaction should preferably occur at low temperatures [25]. Among Group VIII metals, Co is the most active for hydrogenation and exhibits low-temperature activity [84], making it promising for methanation reactions. Tu et al. reported that Co-Zr0.1-B-O amorphous catalysts, with Zr added as an additive, could carry out CO2 methanation reactions as early as 140 °C. The CO2 conversion at 180 °C could reach 10.7 mmol CO2 gcat1 h⁻1, with a methane selectivity of 97.8%, comparable to noble metal catalysts under similar conditions [23].
The addition of carriers is a common means of catalyst optimisation, and metal–carrier interactions can effectively tune the performance of multiphase catalysts. Liang et al. loaded Co and Ni onto Al2O3 and found that Co/Al2O3 catalysts possessed better activity and stability, with a lower reaction onset temperature compared to Ni/Al2O3 (Figure 7a,b) [26]. The coordination effect between the metal Co and the support Al2O3 promotes the formation of reaction intermediates (bicarbonate, carbonate, and formate), resulting in higher methanation catalyst activity. After the reaction, analysis using TG, TPO-MS, and XRD (Figure 7e–h) revealed that less coke formed on the surface of Co/Al2O3 catalysts due to the stronger interaction between Co and Al2O3 compared to Ni. This strong interaction improves resistance to coking and reduces catalyst sintering, leading to better stability. Bogdan et al. compared the performance of Ni/CNT and Co/CNT catalysts for CO2 hydro-methanation under supercritical conditions and showed that Co/CNT catalysts were more active than Ni/CNT catalysts in low H2/CO2 feed gas ratios, with almost 100% methane selectivity and no by-product CO [27]. According to these studies, Co demonstrates better hydrogenation capacity and greater stability than Ni. CeO2 and ZrO2 are considered important Co-based catalyst carriers. Díez-Ramírez and his team investigated the CO2 methanation performance of cobalt-based catalysts with different metal carriers added (CeO2, ZrO2, Gd2O3, ZnO2) at low temperatures (200–300 °C) and atmospheric pressure [28]. Among them, the Co–Ce interaction on the Co/CeO2 catalyst enhanced the hydrogenation performance of the catalyst, and the methanation activity was superior to that of the other catalysts. The better the dispersion of metal nanoparticles on the carrier, the stronger the metal–carrier interaction. Li et al. found that the hydroxyl and carboxyl groups on citric acid, a low-cost complexing agent, favoured the dispersion of the metal on the support. The dispersion of Co on the surface of 2%Co-ZrO2 catalysts prepared with citric acid was better, increasing the strength of the Co–ZrO2 interaction [86]. This strong interaction between metal Co and ZrO2 supports contributed to forming more reduction-active sites and oxygen vacancies, promoting CO2 adsorption and significantly improving CO2 conversion (85%) and CH4 selectivity (99%). Suitable carrier size facilitates the dispersion of metal nanoparticles. Alexander et al. controlled the metal–carrier interaction strength on Co/CeZrO4 catalysts by varying the size of the carriers [87]. It was found that cobalt nanoparticles on CZ carriers in the range of 20–30 nm showed better dispersion and stronger metal–carrier interactions, and the cobalt nanoparticles were more stable during the reduction step and therefore had higher CO2 methanation activity compared to CZ carriers with smaller particle sizes. Meanwhile, the characterisation results demonstrated that the easy formation of oxygen vacancies and oxygen overflow in cerium dioxide is the key to high CO2 methanation activity. In addition, the use of novel MOF materials with large surface areas as carriers can achieve high dispersion of active sites and improve the catalytic performance [88]. In a study by Ampaiah et al., using cerium-zirconium oxide (CZ) as a carrier in the MOF stencilling method resulted in enhanced catalytic activity compared to the conventional co-precipitation method [70,89]. Incorporating CZ–MOF carriers increased the specific surface area and pore volume, facilitating the dispersion of Co3O4. This elevated the oxygen vacancies on the catalyst surface, enhancing CO2 adsorption capacity.
Constructing systematic synergies between metals by adding metal additives can affect the catalytic performance, and the researchers designed Ni-Co bimetallic catalysts to improve the methanation performance. Li et al. found that in the Ni-Co-MgO bimetallic catalyst, a strong synergistic interaction between Co and Ni facilitated the CO2 methanation reaction. In this process, the active site Coδ⁺ activated CO2 to form a monodentate carbonate intermediate, which then reacted with H* dissociated at Ni0 to form methane (Figure 8) [90]. The redox nature of Co in the Ni-Co/CeO2-ZrO2 catalysts reduced carbon deposition and sintering, to which Ni catalysts are prone, thus improving reaction efficiency and stability [91]. It was also observed that when CH4 was present in the feedstock, the methanation reaction was carried out simultaneously with the low-temperature reforming reaction, enhancing catalytic activity.
Mesoporous oxides, with their large specific surface area and sufficient pore space, have great application value as catalyst carriers. Zhou et al. developed Co-based catalysts for CO2 hydrotreating methanation using novel mesoporous supports. Both Co/KIT-6 (KIT-6 is a porous silica molecular sieve) and Co/meso-SiO2 catalysts were prepared with well-defined mesoporous structures and highly dispersed surface Co particles. The Co/KIT-6 catalysts, with a specific surface area of 368.9 m2/g and a highly ordered bicontinuous mesoporous structure, exhibited better CO2 hydrogenation activity and methane selectivity than Co/meso-SiO2 catalysts at high reaction temperatures [92].
Co carbides and nitrides are also very reactive. Yu et al. developed a Co2C/γ-Al2O3 catalyst for CO2 hydrogenation at 300 °C and 3 bar, demonstrating an 89% conversion and 99% selectivity to CH4 [73]. This performance indicated higher CO2 conversion and methane selectivity compared to CoOx/γ-Al2O3 catalysts. Azzaz and colleagues prepared Co4N/γ-Al2O3 catalysts for CO2 methanation using γ-Al2O3 as a carrier. Their results showed that Co4N exhibited higher CO2 methanation activity than Co. Co4N/γ-Al2O3 demonstrated reduced susceptibility to coke deposition and sintering of metal particles and good anti-deactivation performance. This was attributed to the introduction of nitride, which resulted in nitrogen vacancies, increasing the catalyst surface’s alkalinity and enhancing CO2 and H2 adsorption [93]. Table 1 summarises the catalytic performance of recently reported Co-based catalysts for CO2 methanation.
In summary, compared with Ni and noble metal catalysts, Co-based catalysts show advantages of low reaction temperature, relatively good stability, and low cost in CO2 hydrogenation to methane, holding good prospects for industrialisation. Researchers optimised the Co-active site by adjusting the chemical microenvironment to promote C–O bond cleavage and H2 dissociation, thereby improving methanation efficiency.

4.2. CO2 Hydrogenation to Hydrocarbon Chemicals and Fuels

CO2 hydrogenation can produce low-carbon hydrocarbons and liquid fuels, with low-carbon olefins (ethylene, propylene, etc.) being the most basic raw materials in the chemical industry. Liquid fuels, including gasoline (C5–C11), aviation fuel (C8–C16), and diesel fuel (C10–C20), are important energy sources today, derived from petroleum processing [84,94]. Given the declining availability of oil and the global pursuit of sustainable energy sources, converting carbon dioxide into hydrogen via non-oil pathways offers dual benefits: mitigating carbon emissions and reducing reliance on oil resources [75,95].
The Fischer-Tropsch Synthesis (FTS) process, which uses syngas as a feedstock to synthesise C2+ hydrocarbons, has a long history. Numerous highly effective FTS catalysts have also shown high activity in CO2 hydrogenation reactions [96]. Co-based catalysts are high-performing FTS catalysts with substantial carbon chain growth capacity. However, due to their low reverse water–gas shift (RWGS) activity, the primary CO2 hydrogenation product is CH4 [29,97]. To enhance the selectivity for C2+ hydrocarbons in CO2 hydrogenation products, researchers typically incorporate components with RWGS activity into Co catalysts [38,98].
Cu catalysts, extensively researched for WGS, are also employed in RWGS reactions. Consequently, Co-Cu bimetallic catalysts serve as highly effective catalysts for hydrogenating CO2 to long-chain hydrocarbons [24]. In 1950, Russell and Miller found that adding Cu to Co-based catalysts could lower the catalytic reaction temperature, reduce CH4 generation, and promote liquid hydrocarbon production [99]. Alkali metals (K, Na, Cs), characterised by high electron density, facilitate CO2 adsorption, inhibit methane production, and promote carbon chain growth in the product [54,100]. Shi et al. reported the synthesis of alkali metal-modified CoCu/TiO2 catalysts. Doping with alkali metals altered the structural properties and crystal structure of the catalysts, enhanced CO2 chemisorption, reduced H2 adsorption, effectively inhibited CH4 formation, and improved the selectivity of C5+ hydrocarbons, achieving a selectivity of up to 42.1% for C5+ hydrocarbons at 250 °C and 5.0 MPa [30,101].
Fe is an excellent catalyst for the FTS process, exhibiting high activity in the RWGS reaction and playing a pivotal role in CO2 hydrogenation to long-chain hydrocarbons [20,102]. Recent studies on Co-Fe bimetallic catalysts show that Co can advance the chain extension reaction and inhibit carbon deposit generation, thereby improving catalyst activity [53,103]. Wang et al. designed catalysts using carbon shells to achieve a controlled distribution of Co and Fe. The isolated active site of Fe-Co catalysts, separated by the inside and outside of the carbon shells, effectively promoted CO2 hydrogenation efficiency, suppressed methanogenesis, and improved the selectivity of C5+ hydrocarbons, increasing it from 19.8% to 39.7% [55], as Figure 9 shows. Kim et al. prepared Na-modified CoFe2O4/CNT catalysts for the CO2 hydrogenation reaction. Catalysts with 0.32 wt% Na exhibited a selectivity of 38.8% for C2–4 olefins and up to 40.9% for C5+ hydrocarbons [104]. Xu et al. synthesised a series of homogeneous K-containing spinel-type ZnCoFe2−xO4 nanoparticle catalysts via ZnCoFe-LDHs precursors, which exhibited excellent properties for the hydrogenation of CO2 to light olefins [105]. A combination of structural characterisation and reaction results demonstrated that an increase in the Co content facilitated the reduction of Co and Fe species in hydrogen, thereby promoting the formation of more CoFe alloy phases. The electron-rich Fe0 atoms in the CoFe alloys enhanced the dissociation of CO intermediates and significantly contributed to the formation of iron and cobalt carbons, thus improving the reactivity of the catalysts in the production of hydrocarbons and simultaneously suppressing the CO2 methanation and the secondary hydrogenation reaction of olefins. Hwang et al. induced the formation of FeCo alloys using N-coordinated Co atomic structures (Co-NC) as carriers. DFT calculations show that the FeCo alloy structure promotes FTS while inhibiting CH4 formation [31].
The introduction of other transition metals, such as Mo, Mn, Pd, etc., is also favourable for the formation and stabilisation of the active phase and improves the selectivity of long-chain hydrocarbons in the product. Owen et al. reported that a Na-Co-Mo/SiO2-TiO2 catalyst exhibited a 13.5% CO2 conversion, 37.3% selectivity for C2–4 olefins, and 23.1% selectivity for CH4 at 200 °C and 0.1 MPa. They determined that the surface area of the carriers and the metal–oxide carrier interactions influence Co particle size, which subsequently affects the activity and selectivity of the catalysts [106]. He et al. discovered that Co6/MnOx nanocatalysts could effectively catalyse CO2 hydrogenation, with a selectivity of up to 53.2% for liquid hydrocarbons (C5–26) in the total product and a very low CO selectivity of 0.4%. TPD data indicated that incorporating the Mn promoter enhanced CO2 adsorption, inhibited H2 adsorption over the Co catalysts, and drove the generation of long chains [30]. Canio et al. used Pd as an additive to enhance the CO2 hydrogenation performance of Co/TiO2 catalysts. Their findings revealed that the interactions between the two metals were not more pronounced or favourable, and that the mixed monometallic catalysts exhibited higher fuel selectivity than the bimetallic catalysts [107].
The spatial arrangement of catalysts synthesised by methods such as co-precipitation, impregnation, and physical mixing is not subject to precise control [31,108]. The development of well-designed and structurally well-defined bifunctional catalysts not only contributes to the creation of new catalysts but also elucidates reaction mechanisms. Xie et al. designed a bifunctional CeO2-Pt@mSiO2-Co core–shell catalyst for synthesising hydrocarbons with the target product C2–4 using two metal–oxide interfaces to catalyse the sequential reaction [49]. In this case, the higher RWGS activity of CeO2/Pt favoured CO formation, which underwent an F-T reaction to produce C2–4 hydrocarbons at the adjacent Co/mSiO2 interface (Figure 10). Heuntae et al. reported a Mn-promoted Co@CoOx/Co2C core–shell catalyst, where oxygen vacancies on the CoOx in the shell layer activated the RWGS reaction. The resulting CO was transported to the Co2C phase near the nucleus and the metal Co phase to continue the reaction, resulting in the production of long-chain hydrocarbons via the FTS reaction, with a selectivity of up to 14.9% of the C21+ hydrocarbons under optimal reaction conditions [109].
In conclusion, the researchers employed a strategy of modulating the interactions between the active sites of Co and other metals. This involved the introduction of alkali metals, transition metals, or the construction of structurally precise bifunctional catalysts. The objective was to inhibit methane generation, promote C–C chain growth, and improve C2+ hydrocarbon selectivity. Table 2 summarises the catalytic performance of recently reported Co-based catalysts for CO2 to hydrocarbon chemicals and fuels.

4.3. Innovations in CO2-to-Methanol Catalysis

George A. Olah, the 2006 Nobel Prize winner in chemistry, first introduced the concept of the “methanol economy” in his book “Beyond the Oil and Gas Era—The Methanol Economy” (Figure 11). This model represents a green cycle of C1 chemistry and offers a potential solution to the problem of fossil fuel resource scarcity [111]. Alcohols possess several advantageous properties, including high energy density, low vapor pressure, and affinity for water. These characteristics render them excellent fuels, solvents, and chemical raw materials, with a vast and growing market demand [112,113].
The conversion of carbon dioxide (CO2) into methanol is a thermodynamically exothermic process (Equation (3)). Furthermore, it is a reaction process in which the number of molecules is reduced. Consequently, low temperatures and high pressures are conducive to CO2 conversion and high methanol selectivity. To promote the reaction, researchers added metal additives with low-temperature hydrogenation activity, as reported by Zeng et al. His team prepared a Pt3Co catalyst utilising the noble metal Pt as an additive [114]. FT-IR results confirmed that the higher negative charge density of the Pt atoms at the apex of the Pt3Co octopod promotes the activation of CO2, thereby improving catalytic activity and increasing the yield of methanol from CO2 hydrogenation.
The use of oxide carriers allows for modification of the size, shape, and chemical state of the catalyst, which can significantly affect the reaction and interfacial activity [115]. In a recent study, Li et al. reported a MnOx/m-Co3O4 catalyst with a MnOx/CoO interface that exhibited enhanced catalytic activity and methanol selectivity compared to MnO nanoparticles and mesoporous Co3O4 carriers alone [116]. Wang et al. used a method of loading Co nanoparticles onto amorphous SiO2, resulting in the formation of Co–O–SiOₙ bonds on the Co@Si catalyst, which inhibited the formation of CO and CH4. Additionally, this catalyst formed a stable intermediate, CH3O*, improving methanol selectivity. At 320 °C, the catalyst demonstrated a selectivity of 70.5% for methanol and an CO2 conversion of 8.6% [69].
The synthesis of catalysts with the aid of stencil agents can create more active sites and reaction spaces, thereby improving reaction efficiency. Alexey et al. employed gradual pyrolytic oxidative decomposition of indium-impregnated ZIF-67(Co) MOFs to form In2O3@Co3O4 reticulated shell composites [117]. This process yielded a maximum methanol yield of 0.65 gMeOH gcat1 h⁻1 over 100 h, with methanol selectivity as high as 87%, superior to that of commercial Cu/ZnO/Al2O3 catalysts. Screening tests revealed that lower temperatures effectively inhibited side reactions (methanation and RWGS), reducing CO and CH4 selectivities to 11% and 2%, respectively. Lian et al. synthesised Co/C-N materials using ZIF-67 calcined under a N2 atmosphere. They then prepared Co@Co3O4/C-N catalysts by partially oxidising Co/C-N in air under different conditions [118]. It was observed that hydrogen dissociated from the surface of the metal cobalt, and partial oxidation of Co/C-N reduced the cobalt content and hydrogen dissociation ability, decreasing CO2 conversion while increasing methanol selectivity. The catalyst exhibited the highest methanol yield of 2.0 mmol g cat⁻1 h⁻1 at 220 °C.
Co-based catalysts demonstrate efficacy in CO2 hydrogenation to methanol and show considerable potential in synthesising higher-carbon alcohols [22]. The addition of metal additives or oxide supports modulates the Co-active sites on the catalyst surface, altering its ability to dissociate H2 and break C–O bonds [53,76]. Wang et al. reported a CoAlOx catalyst for selective hydrogenation of CO2 to synthesise ethanol, adjusting the catalyst’s activity composition by using different pre-reduction temperatures [45]. The catalyst was optimised for the surface Co-CoO phase following reduction at 600 °C, enhancing acetate intermediate generation and ethanol selectivity in the product, with ethanol selectivity reaching up to 92.1% (Figure 12). It is crucial to adjust the oxidation state of Co in order to form the Co-Coδ⁺-active phase, with the aim of improving alcohol selectivity for Co-based catalysts [65]. Zheng et al. prepared a series of LaCo1xGaxO3 chalcogenide catalysts for ethanol production using the citric acid complexation method [119]. The interaction between Co and Ga resulted in the presence of Coδ⁺ at the reduced interface. The synergistic effect of Co0 and Coδ⁺ moderately weakened the hydrogenation capacity of Co0, inhibited CO2 methanation, and promoted ethanol formation. The conversion over LaCo0.7Ga0.3O3 of CO2 was 9.8%, with 74.7% selectivity to alcohols and 88.1% ethanol content in the alcohol mixture.
The structural morphology of catalysts significantly impacts the conversion and selectivity of reactions. As shown in Figure 13, Ouyang et al. reported using Pt/Co3O4-r (nanorods) and Pt/Co3O4-p (nanoplates) catalysts with Pt additives for the hydrogenation of CO2 to generate C2+ alcohols [120]. The combined effect of Pt/Co on the catalysts after reduction at 200 °C and the partially reduced Co3O4−x surface oxygen vacancies enhanced the adsorption of CO2 and H2, facilitating the reaction. The selectivity of the Pt/Co3O4-r catalyst for C2+ alcohols was found to be 14.8%, while the C2+ alcohol selectivity of the Pt/Co3O4-p catalyst was determined to be 19.2%. The team then synthesised Co3O4-m (mesoporous) using mesoporous silica (KIT-6) as a template [121]. The ordered mesoporous structure on Co3O4-m promoted carbon chain growth and reduced methane selectivity. To further reduce methane content as a by-product, platinum, a noble metal with strong RWGS capability, was introduced into the Pt/Co3O4-m catalyst. This demonstrated a decrease in methane selectivity and an increase in CO and alcohol selectivity.
Cu is the most widely studied catalyst for CO2 hydrogenation to alcohols. The introduction of Cu as an additive can promote the formation of high-carbon alcohols. Muhammad et al. prepared bimetallic Cu-Co tandem catalysts by decorating Co nanoparticles on a dendritic Cu substrate [33]. The Cu-Co interface facilitated the generation of acetaldehyde intermediates for the direct and selective production of n-butanol-rich C3+ alcohols. Additionally, the Cu-Co catalysts demonstrated remarkable stability over 1000 h, effectively inhibiting re-oxidation and carbon deposition. Incorporating non-metallic elements, such as carbon (C) and phosphorus (P), into the catalyst structure can inhibit CO* intermediate dissociation and promote alcohol synthesis [32,34]. Zhang et al. constructed a Co-Co2C catalyst with dual active sites, which exhibited an electron-loading effect on carbon species [34]. This increased the CO2 conversion rate and facilitated non-dissociation of COx intermediates, resulting in increased methanol selectivity while maintaining high ethanol selectivity. Introducing carbon elements effectively inhibited re-oxidation and carbon deposition, stabilising the Cu-Co catalyst. Furthermore, carbon elements stabilised the microstructure of the active components, providing the catalyst with good anti-sintering properties and improved catalytic stability.
The preceding work indicates that Co-based catalysts are not readily capable of forming stable oxygenated intermediates due to their pronounced hydrogenation activity. The principal product is methane, and the selectivity for alcohols is typically poor. Researchers have constructed synergistic active sites, such as Co0-Co2⁺, Co0-Coδ⁺, Co-Co2C, etc., to inhibit the methanation reaction and improve alcohol selectivity by adding additives, carriers, or adjusting the catalyst structure. Table 3 summarises the catalytic performance of recently reported Co-based catalysts for CO2 to hydrocarbon chemicals and fuels.

5. Limitations and Future Perspectives

5.1. Limitations

Co-based catalysts play a pivotal role in the hydrogenation of CO2 to methane, hydrocarbon chemicals, and alcohols. This is due to the strong hydrogen activation ability of Co, which increases the likelihood of CH4 formation. This paper discusses the reaction mechanisms of CO2 hydrogenation with Co-based catalysts, introduces optimised Co-based catalysts for CO2 methanation, and explores the generation of hydrocarbon chemicals and alcohols through the addition of auxiliaries and carrier modifications. Despite the development of various strategies to optimise the nature and structure of Co-based catalysts, the results remain unsatisfactory. Firstly, the potential of Co-based catalysts lies more in the synthesis of long-chain carbon hydrocarbons than in CO2 methanation. However, current CO2 conversion and selectivity for long-chain products do not meet industrial standards, preventing commercial applications. Secondly, achieving precise selectivity for higher alcohols in CO2 hydrogenation with Co-based catalysts remains a significant challenge due to numerous by-products, including CO, methane, and other C2₊ hydrocarbons. Thirdly, the understanding of the reaction mechanism remains incomplete. Although numerous researchers have discussed various active sites on Co catalysts and proposed potential reaction pathways, current characterisation methods do not allow for sufficiently clear and direct observation of complex molecular and atomic motions during the reaction. Furthermore, the complex synergistic effects between multiple active sites hinder a deeper exploration of the reaction mechanism. The lack of clarity in the reaction mechanism constrains researchers’ ability to design high-performance catalysts. Finally, the issue of a sustainable source of hydrogen in the feedstock gas is one of the limiting conditions for the industrialisation of CO2 hydrogenation, and if the source of hydrogen cannot be guaranteed to be clean, the idea of CO2 hydrogenation reactions to produce green fuels will be dashed. Currently, the main technologies for hydrogen production from renewable sources are hydrolysis, biomass thermochemical conversion, and biomass bioconversion [122,123]. Most of these technologies are still in the laboratory stage, and electrolysis of water, which has been used for industrial production, suffers from high costs [123]. How to obtain affordable and sustainable hydrogen is still a direction that researchers have to strive to explore.

5.2. Future Perspectives

In light of these considerations, the following strategies are proposed to address these issues:
(1) Design of Catalysts for Synergistic Interaction of Multiple Active Sites: Based on the current understanding of single active-site catalysts, constructing catalysts that achieve synthetic goals through synergistic interactions is promising; for example, enhancing the efficacy of Co-based catalysts that follow the RGWS–FTS pathway by introducing a low-temperature, high-activity RWGS active site. Alternatively, we can learn from syngas reactions to design HAS multifunctional catalysts that utilise the limiting or moderating effect of zeolites on reaction centres to enhance selectivity for higher carbon alcohols in CO2 hydrogenation reactions.
(2) Development and Popularisation of Advanced Characterisation Techniques: Developing novel techniques based on existing in situ characterisation methods, such as in situ XRD, DRIFTS, and FT-IR, to enhance instrument sensitivity and refine the experimental design. Introducing cutting-edge technologies, such as attosecond lasers, can record instantaneous electronic motion processes, providing insights into molecular and atomistic-scale interactions and dynamics. This enables precise identification and continuous monitoring of active components and key intermediates during CO2 hydrogenation on catalysts.
(3) Introduction of Novel Computational Techniques: Theoretical computation is a valuable tool for elucidating reaction pathways. Recent novel computational techniques, including quantum computing and reaction modelling, integrated with machine learning algorithms, can aid in theoretical interpretation and prediction of catalysts. Purposefully modelling the microenvironment surrounding different Co active sites and analysing free energy barriers to identify the decisive step in the CO2 hydrogenation reaction rate can provide theoretical assistance in constructing catalysts with high selectivity and stability for CO2 hydrogenation.
(4) Development of Reactors to Enhance Catalytic Efficiency: Beyond investigating novel and more efficient catalysts, efforts should be directed towards the design and optimisation of reactors. For instance, membrane reactors, which combine reaction and separation steps, can circumvent equilibrium limitations in converting CO2 to liquid fuels by continuously and selectively removing water as a by-product. This results in a notable enhancement in CO2 conversion and liquid fuel yield. Membrane reactors are currently used in CO2 hydrogenation to methanol [124,125,126].
These strategies aim to address the current challenges in optimising Co-based catalysts for CO2 hydrogenation, paving the way for improved catalytic performance and potential commercial applications.

Author Contributions

X.H.: Conceptualisation, Methodology, Formal Analysis, Writing—Original Draft, Writing—Review and Editing, Visualisation. X.W.: Formal Analysis, Writing—Review and Editing, Visualisation. H.X.: Conceptualisation, Methodology, Resources, Writing—Review and Editing, Supervision, Project Administration, Funding Acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 52270078.

Data Availability Statement

No new data were created or analysed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Liu, C.; Kang, J.; Huang, Z.Q.; Song, Y.H.; Xiao, Y.S.; Song, J.; He, J.X.; Chang, C.R.; Ge, H.Q.; Wang, Y.; et al. Gallium nitride catalyzed the direct hydrogenation of carbon dioxide to dimethyl ether as primary product. Nat. Commun. 2021, 12, 2305. [Google Scholar] [CrossRef] [PubMed]
  2. Le Quéré, C.; Andres, R.J.; Boden, T.; Conway, T.; Houghton, R.A.; House, J.I.; Marland, G.; Peters, G.P.; van der Werf, G.R.; Ahlström, A.; et al. The global carbon budget 1959–2011. Earth Syst. Sci. Data 2013, 5, 165–185. [Google Scholar] [CrossRef]
  3. Lovelock, C.E.; Duarte, C.M. Dimensions of Blue Carbon and emerging perspectives. Biol. Lett. 2019, 15, 20180781. [Google Scholar] [CrossRef] [PubMed]
  4. Mac Dowell, N.; Fennell, P.S.; Shah, N.; Maitland, G.C. The role of CO2 capture and utilization in mitigating climate change. Nat. Clim. Chang. 2017, 7, 243–249. [Google Scholar] [CrossRef]
  5. Chen, S.; Liu, J.; Zhang, Q.; Teng, F.; McLellan, B.C. A critical review on deployment planning and risk analysis of carbon capture, utilization, and storage (CCUS) toward carbon neutrality. Renew. Sustain. Energy Rev. 2022, 167, 112537. [Google Scholar] [CrossRef]
  6. Truong, T.H.; Lin, B.-W.; Lo, C.-H.; Tung, C.-P.; Chao, C.-W. Possible pathways for low carbon transitions: Investigating the efforts of oil companies in CCUS technologies. Energy Strategy Rev. 2024, 54, 101421. [Google Scholar] [CrossRef]
  7. Gu, C.; Li, K.; Gao, S.; Li, J.; Mao, Y. CO2 abatement feasibility for blast furnace CCUS retrofits in BF-BOF steel plants in China. Energy 2024, 294, 130756. [Google Scholar] [CrossRef]
  8. Huang, Z.; Deng, S.; Zhang, Q.; Zhao, R.; Li, S.; Veselovskaya, J.; Kozlov, D.; Wang, J. Closing plastic loop with CCUS: Life cycle assessment of a novel strategy for plastic sustainable transition and negative emissions. Resour. Conserv. Recycl. 2024, 208, 107738. [Google Scholar] [CrossRef]
  9. Sun, B.; Fan, B.; Wu, C.; Xie, J. Exploring incentive mechanisms for the CCUS project in China’s coal-fired power plants: An option-game approach. Energy 2024, 288, 129694. [Google Scholar] [CrossRef]
  10. Lu, S.; Yu, Z.; Zhang, Y.; Xu, T. Review of non-isothermal processes in CCUS from a geomechanical perspective. Earth Sci. Rev. 2024, 255, 104848. [Google Scholar] [CrossRef]
  11. Azar, C.; Lindgren, K.; Obersteiner, M.; Riahi, K.; van Vuuren, D.P.; den Elzen, K.M.G.; Möllersten, K.; Larson, E.D. The feasibility of low CO2 concentration targets and the role of bio-energy with carbon capture and storage (BECCS). Clim. Chang. 2010, 100, 195–202. [Google Scholar] [CrossRef]
  12. Heuberger, C.F.; Staffell, I.; Shah, N.; Mac Dowell, N. Quantifying the value of CCS for the future electricity system. Energy Environ. Sci. 2016, 9, 2497–2510. [Google Scholar] [CrossRef]
  13. Nath, F.; Mahmood, M.N.; Yousuf, N. Recent advances in CCUS: A critical review on technologies, regulatory aspects and economics. Geoenergy Sci. Eng. 2024, 238, 212726. [Google Scholar] [CrossRef]
  14. De, S.; Dokania, A.; Ramirez, A.; Gascon, J. Advances in the design of heterogeneous catalysts and thermocatalytic processes for CO2 utilization. ACS Catal. 2020, 10, 14147–14185. [Google Scholar] [CrossRef]
  15. Bai, S.-T.; De Smet, G.; Liao, Y.; Sun, R.; Zhou, C.; Beller, M.; Maes, B.U.W.; Sels, B.F. Homogeneous and heterogeneous catalysts for hydrogenation of CO2 to methanol under mild conditions. Chem. Soc. Rev. 2021, 50, 4259–4298. [Google Scholar] [CrossRef]
  16. Liu, W.; Sahoo, B.; Junge, K.; Beller, M. Cobalt complexes as an emerging class of catalysts for homogeneous hydrogenations. Acc. Chem. Res. 2018, 51, 1858–1869. [Google Scholar] [CrossRef]
  17. Sen, R.; Goeppert, A.; Surya Prakash, G.K. Homogeneous hydrogenation of CO2 and CO to methanol: The renaissance of low-temperature catalysis in the context of the methanol economy. Angew. Chem. Int. Ed Engl. 2022, 61, e202207278. [Google Scholar] [CrossRef] [PubMed]
  18. Onishi, N.; Himeda, Y. Homogeneous catalysts for CO2 hydrogenation to methanol and methanol dehydrogenation to hydrogen generation. Coord. Chem. Rev. 2022, 472, 214767. [Google Scholar] [CrossRef]
  19. Furimsky, E. CO2 hydrogenation to methanol and methane over carbon-supported catalysts. Ind. Eng. Chem. Res. 2020, 59, 15393–15423. [Google Scholar] [CrossRef]
  20. Yang, S.; Zhang, L.; Wang, Z. Advances in the preparation of light alkene from carbon dioxide by hydrogenation. Fuel 2022, 324, 124503. [Google Scholar] [CrossRef]
  21. Xu, Y.; Wang, L.; Zhou, Q.; Li, Y.; Liu, L.; Nie, W.; Xu, R.; Zhang, J.; Cheng, Z.; Wang, H. Carbon dioxide enabled hydrogen storage by methanol: Highly selective and efficient catalysis with well-defined heterogeneous catalysts. Coord. Chem. Rev. 2024, 508, 215775. [Google Scholar] [CrossRef]
  22. Liu, S.; He, Y.; Fu, W.; Chen, J.; Ren, J.; Liao, L.; Sun, R.; Tang, Z.; Mebrahtu, C.; Zeng, F. Hetero-site cobalt catalysts for higher alcohols synthesis by CO2 hydrogenation: A review. J. CO2 Util. 2023, 67, 102322. [Google Scholar] [CrossRef]
  23. Tu, J.; Wu, H.; Qian, Q.; Han, S.; Chu, M.; Jia, S.; Feng, R.; Zhai, J.; He, M.; Han, B. Low temperature methanation of CO2 over an amorphous cobalt-based catalyst. Chem. Sci. 2021, 12, 3937–3943. [Google Scholar] [CrossRef]
  24. 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] [PubMed]
  25. Zhang, Z.; Shen, C.; Sun, K.; Jia, X.; Ye, J.; Liu, C.-J. Advances in studies of the structural effects of supported Ni catalysts for CO2 hydrogenation: From nanoparticle to single atom catalyst. J. Mater. Chem. A Mater. Energy Sustain. 2022, 10, 5792–5812. [Google Scholar] [CrossRef]
  26. Liang, C.; Tian, H.; Gao, G.; Zhang, S.; Liu, Q.; Dong, D.; Hu, X. Methanation of CO2 over alumina supported nickel or cobalt catalysts: Effects of the coordination between metal and support on formation of the reaction intermediates. Int. J. Hydrogen Energy 2020, 45, 531–543. [Google Scholar] [CrossRef]
  27. Bogdan, V.I.; Pokusaeva, Y.A.; Koklin, A.E.; Savilov, S.V.; Chernyak, S.A.; Lunin, V.V.; Kustov, L.M. Carbon dioxide reduction with hydrogen on carbon-nanotube-supported catalysts under supercritical conditions. Energy Technol. 2019, 7, 1900174. [Google Scholar] [CrossRef]
  28. Díez-Ramírez, J.; Sánchez, P.; Kyriakou, V.; Zafeiratos, S.; Marnellos, G.E.; Konsolakis, M.; Dorado, F. Effect of support nature on the cobalt-catalyzed CO2 hydrogenation. J. CO2 Util. 2017, 21, 562–571. [Google Scholar] [CrossRef]
  29. Owen, R.E.; O’Byrne, J.P.; Mattia, D.; Plucinski, P.; Pascu, S.I.; Jones, M.D. Cobalt catalysts for the conversion of CO2 to light hydrocarbons at atmospheric pressure. Chem. Commun. 2013, 49, 11683–11685. [Google Scholar] [CrossRef]
  30. He, Z.; Cui, M.; Qian, Q.; Zhang, J.; Liu, H.; Han, B. Synthesis of liquid fuel via direct hydrogenation of CO2. Proc. Natl. Acad. Sci. USA 2019, 116, 12654–12659. [Google Scholar] [CrossRef]
  31. Hwang, S.-M.; Han, S.J.; Park, H.-G.; Lee, H.; An, K.; Jun, K.-W.; Kim, S.K. Atomically alloyed Fe–co catalyst derived from a N-coordinated co single-atom structure for CO2 hydrogenation. ACS Catal. 2021, 11, 2267–2278. [Google Scholar] [CrossRef]
  32. Zhang, H.; Mao, D.; Zhang, J.; Wu, D. Tuning the CO2 hydrogenation path by moderately phosphating the Co-Al catalyst toward methanol synthesis. Appl. Catal. B 2024, 340, 123257. [Google Scholar] [CrossRef]
  33. Irshad, M.; Chun, H.-J.; Khan, M.K.; Jo, H.; Kim, S.K.; Kim, J. Synthesis of n-butanol-rich C3+ alcohols by direct CO2 hydrogenation over a stable Cu–Co tandem catalyst. Appl. Catal. B 2024, 340, 123201. [Google Scholar] [CrossRef]
  34. Zhang, S.; An, K.; Xin, J.; Jiang, Y.; Niu, M.; Song, P.; Wu, C.; Wang, H.; Liu, Y. Carbon modified active pairs of Co-Co2C for CO2 hydrogenation to alcohols. Chem. Eng. J. 2024, 486, 150334. [Google Scholar] [CrossRef]
  35. Jia, C.; Gao, J.; Dai, Y.; Zhang, J.; Yang, Y. The thermodynamics analysis and experimental validation for complicated systems in CO2 hydrogenation process. J. Energy Chem. 2016, 25, 1027–1037. [Google Scholar] [CrossRef]
  36. Kattel, S.; Liu, P.; Chen, J.G. Tuning selectivity of CO2 hydrogenation reactions at the metal/oxide interface. J. Am. Chem. Soc. 2017, 139, 9739–9754. [Google Scholar] [CrossRef]
  37. 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. Carbon dioxide methanation on heterogeneous catalysts: A review. Environ. Chem. Lett. 2022, 20, 3613–3630. [Google Scholar] [CrossRef]
  38. Gao, P.; Zhang, L.; Li, S.; Zhou, Z.; Sun, Y. Novel heterogeneous catalysts for CO2 hydrogenation to liquid fuels. ACS Cent. Sci. 2020, 6, 1657–1670. [Google Scholar] [CrossRef] [PubMed]
  39. Jiang, F.; Wang, S.; Liu, B.; Liu, J.; Wang, L.; Xiao, Y.; Xu, Y.; Liu, X. Insights into the influence of CeO2 crystal facet on CO2 hydrogenation to methanol over pd/CeO2 catalysts. ACS Catal. 2020, 10, 11493–11509. [Google Scholar] [CrossRef]
  40. Li, H.; Zhao, J.; Luo, L.; Du, J.; Zeng, J. Symmetry-breaking sites for activating linear carbon dioxide molecules. Acc. Chem. Res. 2021, 54, 1454–1464. [Google Scholar] [CrossRef]
  41. 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, 356, 471–489. [Google Scholar] [CrossRef]
  42. Su, X.; Xu, J.; Liang, B.; Duan, H.; Hou, B.; Huang, Y. Catalytic carbon dioxide hydrogenation to methane: A review of recent studies. J. Energy Chem. 2016, 25, 553–565. [Google Scholar] [CrossRef]
  43. Studt, F.; Behrens, M.; Kunkes, E.L.; Thomas, N.; Zander, S.; Tarasov, A.; Schumann, J.; Frei, E.; Varley, J.B.; Abild-Pedersen, F.; et al. The mechanism of CO and CO2 hydrogenation to methanol over Cu-based catalysts. ChemCatChem 2015, 7, 1105–1111. [Google Scholar] [CrossRef]
  44. Kunkes, E.L.; Studt, F.; Abild-Pedersen, F.; Schlögl, R.; Behrens, M. Hydrogenation of CO2 to methanol and CO on Cu/ZnO/Al2O3: Is there a common intermediate or not? J. Catal. 2015, 328, 43–48. [Google Scholar] [CrossRef]
  45. Wang, L.; Wang, L.; Zhang, J.; Liu, X.; Wang, H.; Zhang, W.; Yang, Q.; Ma, J.; Dong, X.; Yoo, S.J.; et al. Selective hydrogenation of CO2 to ethanol over cobalt catalysts. Angew. Chem. Int. Ed Engl. 2018, 57, 6104–6108. [Google Scholar] [CrossRef] [PubMed]
  46. Khodakov, A.Y.; Chu, W.; Fongarland, P. Advances in the development of novel cobalt Fischer-Tropsch catalysts for synthesis of long-chain hydrocarbons and clean fuels. Chem. Rev. 2007, 107, 1692–1744. [Google Scholar] [CrossRef]
  47. Zeng, F.; Mebrahtu, C.; Xi, X.; Liao, L.; Ren, J.; Xie, J.; Heeres, H.J.; Palkovits, R. Catalysts design for higher alcohols synthesis by CO2 hydrogenation: Trends and future perspectives. Appl. Catal. B 2021, 291, 120073. [Google Scholar] [CrossRef]
  48. Ren, M.; Zhang, Y.; Wang, X.; Qiu, H. Catalytic hydrogenation of CO2 to methanol: A review. Catalysts 2022, 12, 403. [Google Scholar] [CrossRef]
  49. Xie, C.; Chen, C.; Yu, Y.; Su, J.; Li, Y.; Somorjai, G.A.; Yang, P. Tandem catalysis for CO2 hydrogenation to C2-C4 hydrocarbons. Nano Lett. 2017, 17, 3798–3802. [Google Scholar] [CrossRef]
  50. Roy, S.; Cherevotan, A.; Peter, S.C. Thermochemical CO2 hydrogenation to single carbon products: Scientific and technological challenges. ACS Energy Lett. 2018, 3, 1938–1966. [Google Scholar] [CrossRef]
  51. Vogt, C.; Monai, M.; Kramer, G.J.; Weckhuysen, B.M. The renaissance of the Sabatier reaction and its applications on Earth and in space. Nat. Catal. 2019, 2, 188–197. [Google Scholar] [CrossRef]
  52. Lu, Z.; Cheng, Y.; Li, S.; Yang, Z.; Wu, R. CO2 thermoreduction to methanol on the MoS2 supported single Co atom catalyst: A DFT study. Appl. Surf. Sci. 2020, 528, 147047. [Google Scholar] [CrossRef]
  53. Gnanamani, M.K.; Jacobs, G.; Hamdeh, H.H.; Shafer, W.D.; Liu, F.; Hopps, S.D.; Thomas, G.A.; Davis, B.H. Hydrogenation of carbon dioxide over Co–Fe bimetallic catalysts. ACS Catal. 2016, 6, 913–927. [Google Scholar] [CrossRef]
  54. Liang, J.; Liu, J.; Guo, L.; Wang, W.; Wang, C.; Gao, W.; Guo, X.; He, Y.; Yang, G.; Yasuda, S.; et al. CO2 hydrogenation over Fe-Co bimetallic catalysts with tunable selectivity through a graphene fencing approach. Nat. Commun. 2024, 15, 512. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, X.; Guo, L.; Ai, P.; Wu, H.; Lu, Z.; Huang, J.; Tong, J.; Zheng, L.; Sun, S. Isolated dual-active Fe-Co sites efficiently promote CO2 hydrogenation upgrading. Fuel 2024, 364, 131054. [Google Scholar] [CrossRef]
  56. Raghav, H.; Pendem, C.; Tripathi, S.; Kumar, S.; Sarkar, B. Enhanced light olefin production from CO2 over potassium promoted Fe–Co bimetallic ZrO2 supported catalysts. Fuel 2024, 368, 131645. [Google Scholar] [CrossRef]
  57. Wang, D.; Xie, Z.; Porosoff, M.D.; Chen, J.G. Recent advances in carbon dioxide hydrogenation to produce olefins and aromatics. Chem 2021, 7, 2277–2311. [Google Scholar] [CrossRef]
  58. Wang, L.; He, S.; Wang, L.; Lei, Y.; Meng, X.; Xiao, F.-S. Cobalt–nickel catalysts for selective hydrogenation of carbon dioxide into ethanol. ACS Catal. 2019, 9, 11335–11340. [Google Scholar] [CrossRef]
  59. Xu, D.; Wang, Y.; Ding, M.; Hong, X.; Liu, G.; Tsang, S.C.E. Advances in higher alcohol synthesis from CO2 hydrogenation. Chem 2021, 7, 849–881. [Google Scholar] [CrossRef]
  60. Yao, W.; Olajide, G.; Boudreaux, C.M.; Wysocki, M.M.; Ahmed, M.K.; Qu, F.; Szilvási, T.; Papish, E.T. Cobalt(I) pincer complexes as catalysts for CO2 hydrogenation to formate. Organometallics 2024, 43, 1447–1458. [Google Scholar] [CrossRef]
  61. Wu, S.-B.; Zhang, T.; Chung, L.W.; Wu, Y.-D. A missing piece of the mechanism in metal-catalyzed hydrogenation: Co (−I)/Co (0)/Co (+I) catalytic cycle for Co (−I)-catalyzed hydrogenation. Org. Lett. 2019, 21, 360–364. [Google Scholar] [CrossRef]
  62. Yan, X.; Ge, H.; Yang, X. Hydrogenation of CO2 to methanol catalyzed by cp*co complexes: Mechanistic insights and ligand design. Inorg. Chem. 2019, 58, 5494–5502. [Google Scholar] [CrossRef]
  63. Zhang, S.; Liu, X.; Shao, Z.; Wang, H.; Sun, Y. Direct CO2 hydrogenation to ethanol over supported Co2C catalysts: Studies on support effects and mechanism. J. Catal. 2020, 382, 86–96. [Google Scholar] [CrossRef]
  64. Have, I.C.T.; Kromwijk, J.J.G.; Monai, M.; Ferri, D.; Sterk, E.B.; Meirer, F.; Weckhuysen, B.M. Uncovering the reaction mechanism behind CoO as active phase for CO2 hydrogenation. Nat. Commun. 2022, 13, 324. [Google Scholar] [CrossRef]
  65. Chen, Y.; Liu, J.; Chen, X.; Gu, S.; Wei, Y.; Wang, L.; Wan, H.; Guan, G. Development of multifunctional catalysts for the direct hydrogenation of carbon dioxide to higher alcohols. Molecules 2024, 29, 2666. [Google Scholar] [CrossRef] [PubMed]
  66. Gnanamani, M.K.; Jacobs, G.; Keogh, R.A.; Shafer, W.D.; Sparks, D.E.; Hopps, S.D.; Thomas, G.A.; Davis, B.H. Fischer-Tropsch synthesis: Effect of pretreatment conditions of cobalt on activity and selectivity for hydrogenation of carbon dioxide. Appl. Catal. A Gen. 2015, 499, 39–46. [Google Scholar] [CrossRef]
  67. Fu, W.; Liu, S.; He, Y.; Chen, J.; Ren, J.; Chen, H.; Sun, R.; Tang, Z.; Mebrahtu, C.; Zeng, F. CO2 hydrogenation over CoAl based catalysts: Effects of cobalt-metal oxide interaction. Appl. Catal. A Gen. 2024, 678, 119720. [Google Scholar] [CrossRef]
  68. Ding, X.; Fu, J.; Lyu, Y.; Ma, L.; Xu, Y.; Liu, X. Active sites of the cobalt catalysts controlled by surface silanols of silicalite-1 in CO2 hydrogenation to ethanol. Chem. Eng. J. 2024, 494, 152923. [Google Scholar] [CrossRef]
  69. Wang, L.; Guan, E.; Wang, Y.; Wang, L.; Gong, Z.; Cui, Y.; Meng, X.; Gates, B.C.; Xiao, F.-S. Silica accelerates the selective hydrogenation of CO2 to methanol on cobalt catalysts. Nat. Commun. 2020, 11, 1033. [Google Scholar] [CrossRef]
  70. Cui, G.; Lou, Y.; Zhou, M.; Li, Y.; Jiang, G.; Xu, C. Review of mechanism investigations and catalyst developments for CO2 hydrogenation to alcohols. Catalysts 2024, 14, 232. [Google Scholar] [CrossRef]
  71. Li, W.; Nie, X.; Yang, H.; Wang, X.; Polo-Garzon, F.; Wu, Z.; Zhu, J.; Wang, J.; Liu, Y.; Shi, C.; et al. Crystallographic dependence of CO2 hydrogenation pathways over HCP-Co and FCC-Co catalysts. Appl. Catal. B 2022, 315, 121529. [Google Scholar] [CrossRef]
  72. Fan, T.; Liu, H.; Shao, S.; Gong, Y.; Li, G.; Tang, Z. Cobalt catalysts enable selective hydrogenation of CO2 toward diverse products: Recent progress and Perspective. J. Phys. Chem. Lett. 2021, 12, 10486–10496. [Google Scholar] [CrossRef]
  73. Yu, Y.; Mottaghi-Tabar, S.; Iqbal, M.W.; Yu, A.; Simakov, D.S.A. CO2 methanation over alumina-supported cobalt oxide and carbide synthesized by reverse microemulsion method. Catal. Today 2020, 379, 250–261. [Google Scholar] [CrossRef]
  74. Jacobs, G.; Das, T.K.; Zhang, Y.; Li, J.; Racoillet, G.; Davis, B.H. Fischer–Tropsch synthesis: Support, loading, and promoter effects on the reducibility of cobalt catalysts. Appl. Catal. A Gen. 2002, 233, 263–281. [Google Scholar] [CrossRef]
  75. Ra, E.C.; Kim, K.Y.; Kim, E.H.; Lee, H.; An, K.; Lee, J.S. Recycling carbon dioxide through catalytic hydrogenation: Recent key developments and perspectives. ACS Catal. 2020, 10, 11318–11345. [Google Scholar] [CrossRef]
  76. Deng, K.; Lin, L.; Rui, N.; Vovchok, D.; Zhang, F.; Zhang, S.; Senanayake, S.D.; Kim, T.; Rodriguez, J.A. Studies of CO2 hydrogenation over cobalt/ceria catalysts with in situ characterization: The effect of cobalt loading and metal–support interactions on the catalytic activity. Catal. Sci. Technol. 2020, 10, 6468–6482. [Google Scholar] [CrossRef]
  77. Liu, S.; Yang, C.; Zha, S.; Sharapa, D.; Studt, F.; Zhao, Z.-J.; Gong, J. Moderate surface segregation promotes selective ethanol production in CO2 hydrogenation reaction over CoCu catalysts. Angew. Chem. Int. Ed Engl. 2022, 61, e202109027. [Google Scholar] [CrossRef]
  78. Fan, W.K.; Tahir, M. Recent trends in developments of active metals and heterogenous materials for catalytic CO2 hydrogenation to renewable methane: A review. J. Environ. Chem. Eng. 2021, 9, 105460. [Google Scholar] [CrossRef]
  79. 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]
  80. Fayisa, B.A.; Yang, Y.; Zhen, Z.; Wang, M.-Y.; Lv, J.; Wang, Y.; Ma, X. Engineered chemical utilization of CO2 to methanol via direct and indirect hydrogenation pathways: A review. Ind. Eng. Chem. Res. 2022, 61, 10319–10335. [Google Scholar] [CrossRef]
  81. Wei, J.; Ge, Q.; Yao, R.; Wen, Z.; Fang, C.; Guo, L.; Xu, H.; Sun, J. Directly converting CO2 into a gasoline fuel. Nat. Commun. 2017, 8, 15174. [Google Scholar] [CrossRef]
  82. Gao, P.; Li, S.; Bu, X.; Dang, S.; Liu, Z.; Wang, H.; Zhong, L.; Qiu, M.; Yang, C.; Cai, J.; et al. Direct conversion of CO2 into liquid fuels with high selectivity over a bifunctional catalyst. Nat. Chem. 2017, 9, 1019–1024. [Google Scholar] [CrossRef]
  83. Frontera, P.; Macario, A.; Ferraro, M.; Antonucci, P. Supported catalysts for CO2 methanation: A review. Catalysts 2017, 7, 59. [Google Scholar] [CrossRef]
  84. Zhou, W.; Cheng, K.; Kang, J.; Zhou, C.; Subramanian, V.; Zhang, Q.; Wang, Y. New horizon in C1 chemistry: Breaking the selectivity limitation in transformation of syngas and hydrogenation of CO2 into hydrocarbon chemicals and fuels. Chem. Soc. Rev. 2019, 48, 3193–3228. [Google Scholar] [CrossRef]
  85. Sahebdelfar, S.; Takht Ravanchi, M. Carbon dioxide utilization for methane production: A thermodynamic analysis. J. Pet. Sci. Eng. 2015, 134, 14–22. [Google Scholar] [CrossRef]
  86. 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]
  87. Parastaev, A.; Muravev, V.; Huertas Osta, E.; van Hoof, A.J.F.; Kimpel, T.F.; Kosinov, N.; Hensen, E.J.M. Boosting CO2 hydrogenation via size-dependent metal–support interactions in cobalt/ceria-based catalysts. Nat. Catal. 2020, 3, 526–533. [Google Scholar] [CrossRef]
  88. 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]
  89. Jampaiah, D.; Damma, D.; Chalkidis, A.; Venkataswamy, P.; Bhargava, S.K.; Reddy, B.M. MOF-derived ceria-zirconia supported Co3O4 catalysts with enhanced activity in CO2 methanation. Catal. Today 2020, 356, 519–526. [Google Scholar] [CrossRef]
  90. Li, Z.; Chen, J.; Xie, Y.; Wen, J.; Weng, H.; Wang, M.; Zhang, J.; Cao, J.; Tian, G.; Zhang, Q.; et al. Zonal activation of molecular carbon dioxide and hydrogen over dual sites Ni-Co-MgO catalyst for CO2 methanation: Synergistic catalysis of Ni and Co species. J. Energy Chem. 2024, 91, 213–225. [Google Scholar] [CrossRef]
  91. 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]
  92. 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]
  93. Razzaq, R.; Li, C.; Usman, M.; Suzuki, K.; Zhang, S. A highly active and stable Co4N/γ-Al2O3 catalyst for CO and CO2 methanation to produce synthetic natural gas (SNG). Chem. Eng. J. 2015, 262, 1090–1098. [Google Scholar] [CrossRef]
  94. Jiang, Y.; Wang, K.; Wang, Y.; Liu, Z.; Gao, X.; Zhang, J.; Ma, Q.; Fan, S.; Zhao, T.-S.; Yao, M. Recent advances in thermocatalytic hydrogenation of carbon dioxide to light olefins and liquid fuels via modified Fischer-Tropsch pathway. J. CO2 Util. 2023, 67, 102321. [Google Scholar] [CrossRef]
  95. Ghosh, I.K.; Iqbal, Z.; van Heerden, T.; van Steen, E.; Bordoloi, A. Insights into the unusual role of chlorine in product selectivity for direct hydrogenation of CO/CO2 to short-chain olefins. Chem. Eng. J. 2021, 413, 127424. [Google Scholar] [CrossRef]
  96. Cheng, K.; Kang, J.; King, D.L.; Subramanian, V.; Zhou, C.; Zhang, Q.; Wang, Y. Advances in catalysis for syngas conversion to hydrocarbons. In Advances in Catalysis; Elsevier: Amsterdam, The Netherlands, 2017; pp. 125–208. [Google Scholar] [CrossRef]
  97. Guo, L.; Sun, J.; Ji, X.; Wei, J.; Wen, Z.; Yao, R.; Xu, H.; Ge, Q. Directly converting carbon dioxide to linear α-olefins on bio-promoted catalysts. Commun. Chem. 2018, 1, 11. [Google Scholar] [CrossRef]
  98. Li, W.; Zhang, A.; Jiang, X.; Janik, M.J.; Qiu, J.; Liu, Z.; Guo, X.; Song, C. The anti-sintering catalysts: Fe–Co–Zr polymetallic fibers for CO2 hydrogenation to C2= –C4= –rich hydrocarbons. J. CO2 Util. 2018, 23, 219–225. [Google Scholar] [CrossRef]
  99. Russell, W.W.; Miller, G.H. Catalytic hydrogenation of carbon dioxide to higher hydrocarbons1. J. Am. Chem. Soc. 1950, 72, 2446–2454. [Google Scholar] [CrossRef]
  100. Li, W.; Zhang, G.; Jiang, X.; Liu, Y.; Zhu, J.; Ding, F.; Liu, Z.; Guo, X.; Song, C. CO2 hydrogenation on unpromoted and M-promoted co/TiO2 catalysts (M = Zr, K, cs): Effects of crystal phase of supports and metal–support interaction on tuning product distribution. ACS Catal. 2019, 9, 2739–2751. [Google Scholar] [CrossRef]
  101. Shi, Z.; Yang, H.; Gao, P.; Chen, X.; Liu, H.; Zhong, L.; Wang, H.; Wei, W.; Sun, Y. Effect of alkali metals on the performance of CoCu/TiO2 catalysts for CO2 hydrogenation to long-chain hydrocarbons. Chin. J. Catalysis 2018, 39, 1294–1302. [Google Scholar] [CrossRef]
  102. Zhang, Y.-X.; Guo, X.-Y.; Liu, B.; Zhang, J.-L.; Gao, X.-H.; Ma, Q.-X.; Fan, S.-B.; Zhao, T.-S. Cellulose modified iron catalysts for enhanced light olefins and linear C5+ α-olefins from CO hydrogenation. Fuel 2021, 294, 120504. [Google Scholar] [CrossRef]
  103. Numpilai, T.; Chanlek, N.; Poo-Arporn, Y.; Wannapaiboon, S.; Cheng, C.K.; Siri-Nguan, N.; Sornchamni, T.; Kongkachuichay, P.; Chareonpanich, M.; Rupprechter, G.; et al. Pore size effects on physicochemical properties of Fe-Co/K-Al2O3 catalysts and their catalytic activity in CO2 hydrogenation to light olefins. Appl. Surf. Sci. 2019, 483, 581–592. [Google Scholar] [CrossRef]
  104. Kim, K.Y.; Lee, H.; Noh, W.Y.; Shin, J.; Han, S.J.; Kim, S.K.; An, K.; Lee, J.S. Cobalt ferrite nanoparticles to form a catalytic co–Fe alloy carbide phase for selective CO2 hydrogenation to light olefins. ACS Catal. 2020, 10, 8660–8671. [Google Scholar] [CrossRef]
  105. Xu, Q.; Xu, X.; Fan, G.; Yang, L.; Li, F. Unveiling the roles of Fe-Co interactions over ternary spinel-type ZnCoxFe2-xO4 catalysts for highly efficient CO2 hydrogenation to produce light olefins. J. Catal. 2021, 400, 355–366. [Google Scholar] [CrossRef]
  106. Owen, R.E.; Plucinski, P.; Mattia, D.; Torrente-Murciano, L.; Ting, V.P.; Jones, M.D. Effect of support of Co-Na-Mo catalysts on the direct conversion of CO2 to hydrocarbons. J. CO2 Util. 2016, 16, 97–103. [Google Scholar] [CrossRef]
  107. Scarfiello, C.; Durupt, A.; Tison, Y.; Minh, D.P.; Soulantica, K.; Serp, P. Effects of Pd and Co intimacy in Pd-modified Co/TiO2 catalysts for direct CO2 hydrogenation to fuels: The closer not the better. Catal. Sci. Technol. 2024, 14, 2896–2907. [Google Scholar] [CrossRef]
  108. Riedel, T.; Schulz, H.; Schaub, G.; Jun, K.-W.; Hwang, J.-S.; Lee, K.-W. Fischer–tropsch on iron with H2/CO and H2/CO2as synthesis gases: The episodes of formation of the Fischer–tropsch regime and construction of the catalyst. Top. Catal. 2003, 26, 41–54. [Google Scholar] [CrossRef]
  109. Jo, H.; Khan, M.K.; Irshad, M.; Arshad, M.W.; Kim, S.K.; Kim, J. Unraveling the role of cobalt in the direct conversion of CO2 to high-yield liquid fuels and lube base oil. Appl. Catal. B 2022, 305, 121041. [Google Scholar] [CrossRef]
  110. Shi, Z.; Yang, H.; Gao, P.; Li, X.; Zhong, L.; Wang, H.; Liu, H.; Wei, W.; Sun, Y. Direct conversion of CO2 to long-chain hydrocarbon fuels over K–promoted CoCu/TiO2 catalysts. Catal. Today 2018, 311, 65–73. [Google Scholar] [CrossRef]
  111. Olah, G.A.; Goeppert, A.; Prakash, G.K.S. Beyond Oil and Gas: The Methanol Economy; Wiley: Hoboken, NJ, USA, 2009; ISBN 978352732422. [Google Scholar] [CrossRef]
  112. Dalena, F.; Senatore, A.; Basile, M.; Knani, S.; Basile, A.; Iulianelli, A. Advances in methanol production and utilization, with particular emphasis toward hydrogen generation via membrane reactor technology. Membranes 2018, 8, 98. [Google Scholar] [CrossRef]
  113. Christensen, E.; Yanowitz, J.; Ratcliff, M.; McCormick, R.L. Renewable oxygenate blending effects on gasoline properties. Energy Fuels 2011, 25, 4723–4733. [Google Scholar] [CrossRef]
  114. Khan, M.U.; Wang, L.; Liu, Z.; Gao, Z.; Wang, S.; Li, H.; Zhang, W.; Wang, M.; Wang, Z.; Ma, C.; et al. Pt3Co octapods as superior catalysts of CO2 hydrogenation. Angew. Chem. Int. Ed Engl. 2016, 55, 9548–9552. [Google Scholar] [CrossRef] [PubMed]
  115. Konsolakis, M.; Lykaki, M. Recent advances on the rational design of non-precious metal oxide catalysts exemplified by CuOx/CeO2 binary system: Implications of size, shape and electronic effects on intrinsic reactivity and metal-support interactions. Catalysts 2020, 10, 160. [Google Scholar] [CrossRef]
  116. Li, C.-S.; Melaet, G.; Ralston, W.T.; An, K.; Brooks, C.; Ye, Y.; Liu, Y.-S.; Zhu, J.; Guo, J.; Alayoglu, S.; et al. High-performance hybrid oxide catalyst of manganese and cobalt for low-pressure methanol synthesis. Nat. Commun. 2015, 6, 6538. [Google Scholar] [CrossRef]
  117. Pustovarenko, A.; Dikhtiarenko, A.; Bavykina, A.; Gevers, L.; Ramírez, A.; Russkikh, A.; Telalovic, S.; Aguilar, A.; Hazemann, J.-L.; Ould-Chikh, S.; et al. Metal–organic framework-derived synthesis of cobalt indium catalysts for the hydrogenation of CO2 to methanol. ACS Catal. 2020, 10, 5064–5076. [Google Scholar] [CrossRef]
  118. Lian, Y.; Fang, T.; Zhang, Y.; Liu, B.; Li, J. Hydrogenation of CO2 to alcohol species over Co@Co3O4/C-N catalysts. J. Catal. 2019, 379, 46–51. [Google Scholar] [CrossRef]
  119. Zheng, J.-N.; Kang, A.; Wang, J.-M.; Jing, L.; Yuan, L. Effect of support nature on the cobalt-catalyzed CO2 hydrogenation Co/La-Ga-O composite oxide catalyst. J. Fuel Chem. Technol. 2019, 47, 697–708. [Google Scholar] [CrossRef]
  120. Ouyang, B.; Xiong, S.; Zhang, Y.; Liu, B.; Li, J. The study of morphology effect of Pt/Co3O4 catalysts for higher alcohol synthesis from CO2 hydrogenation. Appl. Catal. A Gen. 2017, 543, 189–195. [Google Scholar] [CrossRef]
  121. Liu, B.; Ouyang, B.; Zhang, Y.; Lv, K.; Li, Q.; Ding, Y.; Li, J. Effects of mesoporous structure and Pt promoter on the activity of Co-based catalysts in low-temperature CO2 hydrogenation for higher alcohol synthesis. J. Catal. 2018, 366, 91–97. [Google Scholar] [CrossRef]
  122. Andrews, J.W. Hydrogen production and carbon sequestration by steam methane reforming and fracking with carbon dioxide. Int. J. Hydrogen Energy 2020, 45, 9279–9284. [Google Scholar] [CrossRef]
  123. Sarmah, M.K.; Singh, T.P.; Kalita, P.; Dewan, A. Sustainable hydrogen generation and storage—A review. RSC Adv. 2023, 13, 25253–25275. [Google Scholar] [CrossRef]
  124. Li, H.; Qiu, C.; Ren, S.; Dong, Q.; Zhang, S.; Zhou, F.; Liang, X.; Wang, J.; Li, S.; Yu, M. Na+-gated water-conducting nanochannels for boosting CO2 conversion to liquid fuels. Science 2020, 367, 667–671. [Google Scholar] [CrossRef]
  125. Saeidi, S.; Amin, N.A.S.; Rahimpour, M.R. Hydrogenation of CO2 to value-added products—A review and potential future developments. J. CO2 Util. 2014, 5, 66–81. [Google Scholar] [CrossRef]
  126. Samimi, F.; Karimipourfard, D.; Rahimpour, M.R. Green methanol synthesis process from carbon dioxide via reverse water gas shift reaction in a membrane reactor. Chem. Eng. Res. Des. 2018, 140, 44–67. [Google Scholar] [CrossRef]
Figure 1. Schematic of carbon capture, utilisation, and storage (CCUS) technology.
Figure 1. Schematic of carbon capture, utilisation, and storage (CCUS) technology.
Catalysts 14 00560 g001
Figure 2. Catalytic hydrogenation of carbon dioxide into high-value fuels and chemicals.
Figure 2. Catalytic hydrogenation of carbon dioxide into high-value fuels and chemicals.
Catalysts 14 00560 g002
Figure 3. Mechanism of catalytic carbon dioxide hydrogenation over Co-based catalysts.
Figure 3. Mechanism of catalytic carbon dioxide hydrogenation over Co-based catalysts.
Catalysts 14 00560 g003
Figure 4. (a) Reaction energy diagram for CO2 hydrogenation on the CoAlLDH-P2-R model catalyst. Reaction energy diagram for the hydrogenation routes of H3CO* over the (b) CoAlLDH-P2-R and (c) CoAlLDH-R models. (d) Partial density of states (PDOS) of the atoms near the oxygen vacancy on the surface of the CoAlLDH-P2-R and CoAlLDH-R models. (e) Mechanism illustration and CO2 hydrogenation pathways on CoAlLDH-P2-R. Reproduced from Ref. [32] with permission from Elsevier, copyright 2024.
Figure 4. (a) Reaction energy diagram for CO2 hydrogenation on the CoAlLDH-P2-R model catalyst. Reaction energy diagram for the hydrogenation routes of H3CO* over the (b) CoAlLDH-P2-R and (c) CoAlLDH-R models. (d) Partial density of states (PDOS) of the atoms near the oxygen vacancy on the surface of the CoAlLDH-P2-R and CoAlLDH-R models. (e) Mechanism illustration and CO2 hydrogenation pathways on CoAlLDH-P2-R. Reproduced from Ref. [32] with permission from Elsevier, copyright 2024.
Catalysts 14 00560 g004
Figure 5. The geometry structures, the barrier energy (black fonts), and the reaction energy (red fonts) of the reversed water–gas shift reaction (RWGS) and the CO-hydrogenation. Reproduced from Ref. [52] with permission from Elsevier, copyright 2020.
Figure 5. The geometry structures, the barrier energy (black fonts), and the reaction energy (red fonts) of the reversed water–gas shift reaction (RWGS) and the CO-hydrogenation. Reproduced from Ref. [52] with permission from Elsevier, copyright 2020.
Catalysts 14 00560 g005
Figure 6. Modulated excitation diffuse reflectance infrared Fourier transform spectroscopy (ME DRIFTS) during CO2 hydrogenation with the cobaltbased catalysts. (a) Averaged time-resolved DRIFT spectra and (b) phase-resolved amplitude spectra of CoO (suffix: -ox) and metallic Co (suffix: -red) supported catalysts (T = 250 °C, P = 1 bar, H2/CO2 = 3). Adsorbed surface species with characteristic vibrational energies: (c). CO, (d). formyl, (e). formate, (f). carbonate, and (g). bicarbonate. (h). Simplified reaction pathways for cobalt-catalyzed CO2 hydrogenation to hydrocarbons. Reproduced from Ref. [64] with permission from Nature Communications, copyright 2022.
Figure 6. Modulated excitation diffuse reflectance infrared Fourier transform spectroscopy (ME DRIFTS) during CO2 hydrogenation with the cobaltbased catalysts. (a) Averaged time-resolved DRIFT spectra and (b) phase-resolved amplitude spectra of CoO (suffix: -ox) and metallic Co (suffix: -red) supported catalysts (T = 250 °C, P = 1 bar, H2/CO2 = 3). Adsorbed surface species with characteristic vibrational energies: (c). CO, (d). formyl, (e). formate, (f). carbonate, and (g). bicarbonate. (h). Simplified reaction pathways for cobalt-catalyzed CO2 hydrogenation to hydrocarbons. Reproduced from Ref. [64] with permission from Nature Communications, copyright 2022.
Catalysts 14 00560 g006
Figure 7. Prolonged reaction time tests over the catalysis: (a) 15%Ni/Al2O3 and (b) 15%Co/ Al2O3. TG (c,d) and TPO-MS (eh) characterisations of the 15%Ni/Al2O3 and 15%Co/Al2O3 after tests. Reproduced from Ref. [26] with permission from Elsevier, copyright 2020.
Figure 7. Prolonged reaction time tests over the catalysis: (a) 15%Ni/Al2O3 and (b) 15%Co/ Al2O3. TG (c,d) and TPO-MS (eh) characterisations of the 15%Ni/Al2O3 and 15%Co/Al2O3 after tests. Reproduced from Ref. [26] with permission from Elsevier, copyright 2020.
Catalysts 14 00560 g007
Figure 8. Schematic diagram of the (a) reaction pathway for CO2 methanation; (b) possible reaction pathway over Ni-Co-MgO. Reproduced from Ref. [90] with permission from Elsevier, copyright 2024.
Figure 8. Schematic diagram of the (a) reaction pathway for CO2 methanation; (b) possible reaction pathway over Ni-Co-MgO. Reproduced from Ref. [90] with permission from Elsevier, copyright 2024.
Catalysts 14 00560 g008
Figure 9. Detailed hydrocarbon distribution over different catalysts. (a) Catalytic performances over different catalysts. (b) Effects of cobalt contents on catalyst performance. (c) Reaction scheme over catalysts for the synthesis of hydrocarbons from CO2 hydrogenation. Reproduced from Ref. [55] with permission from Elsevier, copyright 2024.
Figure 9. Detailed hydrocarbon distribution over different catalysts. (a) Catalytic performances over different catalysts. (b) Effects of cobalt contents on catalyst performance. (c) Reaction scheme over catalysts for the synthesis of hydrocarbons from CO2 hydrogenation. Reproduced from Ref. [55] with permission from Elsevier, copyright 2024.
Catalysts 14 00560 g009
Figure 10. (a) Catalytic performance of single-interface catalysts CeO2-Pt@mSiO2 and CeO2@mSiO2-Co, physical mixture catalyst, and tandem catalyst CeO2-Pt@mSiO2-Co (H2/CO2 ratio is 3, reaction temperature is 250 °C). (b) CO2 conversion and hydrocarbon distribution at different H2/CO2 ratios over the tandem catalyst at 250 °C. (c) Schematic of synthetic process. Reproduced from Ref. [49] with permission from ACS, copyright 2017.
Figure 10. (a) Catalytic performance of single-interface catalysts CeO2-Pt@mSiO2 and CeO2@mSiO2-Co, physical mixture catalyst, and tandem catalyst CeO2-Pt@mSiO2-Co (H2/CO2 ratio is 3, reaction temperature is 250 °C). (b) CO2 conversion and hydrocarbon distribution at different H2/CO2 ratios over the tandem catalyst at 250 °C. (c) Schematic of synthetic process. Reproduced from Ref. [49] with permission from ACS, copyright 2017.
Catalysts 14 00560 g010
Figure 11. Schematic diagram of the green methanol cycle.
Figure 11. Schematic diagram of the green methanol cycle.
Catalysts 14 00560 g011
Figure 12. (a,b) The performance of various catalysts in CO2 hydrogenation. Reaction conditions: catalyst (20 mg), H2O (2 mL), initial pressure 4.0 MPa (H2/CO2 = 3:1), 15 h, 140 °C or 200 °C. The yields and selectivity are based on the number of moles of carbon. (c) XRD patterns of CoAl-LDH, Co-Al oxides, and CoAlOx catalysts were reduced at different temperatures. Reproduced from Ref. [45] with permission from GDCh, copyright 2018.
Figure 12. (a,b) The performance of various catalysts in CO2 hydrogenation. Reaction conditions: catalyst (20 mg), H2O (2 mL), initial pressure 4.0 MPa (H2/CO2 = 3:1), 15 h, 140 °C or 200 °C. The yields and selectivity are based on the number of moles of carbon. (c) XRD patterns of CoAl-LDH, Co-Al oxides, and CoAlOx catalysts were reduced at different temperatures. Reproduced from Ref. [45] with permission from GDCh, copyright 2018.
Catalysts 14 00560 g012
Figure 13. TEM (a,b) and SEM (c,d) images of Co3O4-r and Co3O4-p. (e) Effect of temperature on alcohol selectivity over Pt/Co3O4-r and Pt/Co3O4-p catalysts. Reproduced from Ref. [120] with permission from Elsevier, copyright 2017.
Figure 13. TEM (a,b) and SEM (c,d) images of Co3O4-r and Co3O4-p. (e) Effect of temperature on alcohol selectivity over Pt/Co3O4-r and Pt/Co3O4-p catalysts. Reproduced from Ref. [120] with permission from Elsevier, copyright 2017.
Catalysts 14 00560 g013
Table 1. CO2 hydrogenation to methane over Co-based catalysts.
Table 1. CO2 hydrogenation to methane over Co-based catalysts.
EntryCatalystT
(°C)
P
(MPa)
GHSV bWHSV cSV bCO2 Conv. (%)CH4 Sel. (%)H2/CO2Ref.
1Co-Zr0.1-B-O a1808.0----97.81.0[23]
215 wt% Co/Al2O34001 atm 16,000--82804.0[26]
325 wt% Ni/Al2O34001 atm 16,000--75694.0[25]
4Co(10)/CNT2508.0---241001.0[27]
5Ni(10)/CNT2508.0---41001.0[27]
6Ni-Co-MgO290-20,000--80.599.64.0[90]
7Ni-Co/CeO2-ZrO2350--12,000-61974.0[91]
82%Co-ZrO24003.0 --720085994.0[86]
9Co/KIT-6280-22,000--48.91004.6[92]
10Co/meso-SiO2300-22,000--2868.14.6[89]
1120 wt%Co/CZ-MOF3201.5--150081.21003.0[73]
12Co2C/γ-Al2O33000.360,000--89994.0[93]
1320Co4N/γ-Al2O33001.55000--98984.0[93]
14Co/CeO23000.1---97.0≈1009.0[28]
1510CoCZ7002250.1---50≈464.0[87]
a. Kettle reactor b. The unit is mL·g−1·h−1 c. The unit is mL·g−1·hcat−1.
Table 2. CO2 Hydrogenation to hydrocarbon chemicals and fuels over Co-based catalysts.
Table 2. CO2 Hydrogenation to hydrocarbon chemicals and fuels over Co-based catalysts.
EntryCatalystT
(°C)
P
(MPa)
CO2 Conv. (%)CO Sel. (%)Product Selectivity (%)H2/
CO2
Ref.
CH4C=2–4C2–4C5+
1C/Co3@C/Fe3202.033.423.814.436.89.139.73.0[55]
2Na-CoCu/TiO22505.018.430.226.131.842.13.0[101]
3K-CoCu/TiO22505.013.035.134.130.835.13.0[110]
4Na-Co-Mo/SiO2-TiO22000.113.566.723.137.315.424.23.0[106]
5Na-CoFe2O4/CNT3401.034.418.614.85.538.840.93.0[104]
6ZnCo0.5Fe1.5O43202.549.65.818.96.236.138.73.0[105]
7FeK/Co-NC3002.551.72.822.234.243.63.0[31]
8Co6/MnOx a2008.015.30.446.653.41.0[30]
910Co/TiO2* + 1Pd/TiO2*2202.019.40.866.412.720.12.0[107]
10CeO2-Pt@mSiO2-Co2500.622474604003.0[49]
11Mn-Co@CoOx/Co2C2704.064.70.244.22.017.836.03.0[109]
a. Kettle reactor.
Table 3. CO2 Hydrogenation to hydrocarbon chemicals and fuels over Co-based catalysts.
Table 3. CO2 Hydrogenation to hydrocarbon chemicals and fuels over Co-based catalysts.
EntryCatalystT
(°C)
P
(MPa)
CO2 Conv. (%)Product Selectivity (%)H2/CO2Ref.
COCH4CH3OHC2+OH
1Pt3Co a1503.2-Produced methanol = 17.3mmol(5h)3.0[114]
2MnOx/m-Co3O42500.4-SMethanol = 30%, SEthylene = 10%, SDimeyhyi = 5%3.2[116]
3Co@Si0.953202.08.6821.570.5-3.0[69]
4In2O3@Co3O42505.0-112.087-4.0[117]
5Co@Co3O4/C-N a2202.018.6079.518.31.23.0[118]
6CoAlOx-600 a1404.0---->92.13.0[45]
7Pt/Co3O4-r2002.027.8060.014.83.0[120]
8Pt/Co3O4-p2002.022.4077.819.23.0[120]
9Pt/Co3O4-m2002.010.728.314.947.23.0[120]
10LaCo0.7Ga0.3O32403.09.8023.174.73.0[119]
11Na-CuCo-93304.022.120.518.04.1SC3+OH = 27.4%1.0[121]
12600-CDM
(CoGa1.0Al1.0O4/SiO2)
2703.03.7--29.516.9 C-mol%-[34]
a. Kettle reactor.
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

He, X.; Wang, X.; Xu, H. Advancements in Cobalt-Based Catalysts for Enhanced CO2 Hydrogenation: Mechanisms, Applications, and Future Directions: A Short Review. Catalysts 2024, 14, 560. https://doi.org/10.3390/catal14090560

AMA Style

He X, Wang X, Xu H. Advancements in Cobalt-Based Catalysts for Enhanced CO2 Hydrogenation: Mechanisms, Applications, and Future Directions: A Short Review. Catalysts. 2024; 14(9):560. https://doi.org/10.3390/catal14090560

Chicago/Turabian Style

He, Xixue, Xinyu Wang, and Hao Xu. 2024. "Advancements in Cobalt-Based Catalysts for Enhanced CO2 Hydrogenation: Mechanisms, Applications, and Future Directions: A Short Review" Catalysts 14, no. 9: 560. https://doi.org/10.3390/catal14090560

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