Next Article in Journal
Flower-like Titanium Dioxide/Cellulose Acetate Nanofibers for Catalytic Decomposition of Organic Pollutants Including Particulate Matter Removal
Previous Article in Journal
Catalytic Conversion of Glycerol to Lactic Acid Over Cu-Based Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of Mechanism Investigations and Catalyst Developments for CO2 Hydrogenation to Alcohols

by
Guoqing Cui
*,
Yingjie Lou
,
Mingxia Zhou
,
Yuming Li
,
Guiyuan Jiang
and
Chunming Xu
State Key Laboratory of Heavy Oil Processing, China University of Petroleum, Beijing 102249, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(4), 232; https://doi.org/10.3390/catal14040232
Submission received: 20 March 2024 / Revised: 27 March 2024 / Accepted: 29 March 2024 / Published: 31 March 2024
(This article belongs to the Section Industrial Catalysis)

Abstract

:
Heterogeneous thermal-catalytic CO2 hydrogenation to alcohols using renewable energy is a highly attractive approach for recycling greenhouse gases into high-value chemicals and fuels, thereby reducing the dependence on fossil fuels, while simultaneously mitigating the CO2 emission and environmental problems. Currently, great advances have been made on the heterogeneous catalysts, but an in-depth and more comprehensive understanding to further promote this reaction process is still lacking. Herein, we highlight the thermodynamic and kinetic analysis of CO2 hydrogenation reaction firstly. Then, various reaction pathways for CO2 hydrogenation to methanol and higher alcohols (C2+ alcohols) have been discussed in detail, respectively, by combining the experimental studies and density functional theory calculations. On this basis, the key factors influencing the reaction performance, such as metal dispersion, support modification, promoter addition and their structural optimization, are summarized on the metal-based and metal-oxide-based catalysts. In addition, the catalytic performance of CO2 hydrogenation to alcohols and the relationship between structure and properties are mainly summarized and analyzed in the past five years. To conclude, the current challenges and potential strategies in catalyst design, structural characterization and reaction mechanisms are presented for CO2 hydrogenation to alcohols.

Graphical Abstract

1. Introduction

Massive anthropogenic carbon dioxide (CO2) emissions lead to major environmental challenges, such as global warming and ocean acidification, predominantly caused by the continuing consumption of fossil fuels worldwide for rapid industry development. It is urgent to take highly efficient measures in alleviating CO2 emissions to protect the environment and build a sustainable society. Recently, some CO2 mitigation approaches have focused on carbon capture, utilization and storage (CCUS) technology, because CO2 is a renewable carbon resource with the characteristics of safety, non-toxicity and abundant reserves. From a chemical point of view, converting CO2 into some high-value-added chemicals and fuels has been considered as a promising method, which is of great significance for achieving the balance of economy, energy and environment [1]. There are various chemical processes for effective CO2 utilization, for example, thermal catalysis, electrocatalysis, photocatalysis and organic catalysis synthesis, etc. [2,3,4,5], which can obtain many valuable products, e.g., CO, methane, olefins, aromatics, gasoline, urea, alcohols and so on [6,7,8,9,10,11,12]. However, chemical conversion of CO2 is usually formidably challenging in both scientific and technical areas, due to it being thermodynamically stable and chemically inert [13]. When using hydrogen with a higher Gibbs free energy molecule as a reactant, the conversion reaction between CO2 and H2 is more favorable in thermodynamics. In addition, hydrogen can be obtained by some clean and green routes, such as the water splitting by electrolysis using renewable energy sources and biomass steam reforming. Therefore, converting CO2 to valuable products has become one of the most practical carbon dioxide utilization approaches, especially using hydrogen as reactant.
Currently, many methods have been developed to achieve effective CO2 hydroconversion, such as thermochemical, photochemical and electrochemical routes, etc. Among them, the thermochemical CO2 hydrogenation process has attracted widespread attention, because of its many unique advantages, including high catalytic performance and controllable reaction conditions. For instance, the CO, CH4 and methanol syntheses from CO2 hydrogenation reactions have gained great progress in large-scale applications in the recent decades. Among the various products, alcohols, including methanol, ethanol and other higher alcohols, are considerably desirable, owing to their broad range of applications. The methanol usually serves as one of the most important chemical raw materials, with high worldwide annual production of about 65 million tons, which is applied in pesticide, medicine, chemical, building materials, automotive and other industries. Moreover, methanol can be used as an intermediate and chemical building block for the synthesis of other commodities and specialty products, such as light olefins, formaldehyde, acetic acid, dimethyl ether, methyl tert-butyl ether and gasoline [14,15,16]. In addition, methanol is a good energy carrier for fuel and additive for gasoline, owing to its high volumetric energy density and octane rating, respectively. As the energy storage materials in the concept of “Methanol Economy”, renewable methanol will promote good development prospects in green and clean energy utilization in the future. As for the higher alcohols, such as ethanol, they display similar merits as methanol above. Importantly, the higher alcohols have a higher energy density and lower toxicity compared with methanol, leading to more suitable applications in fuel additives and pure fuels [17]. As a result, the direct synthesis of alcohols from CO2 hydrogenation has attracted wide investigation interest as a promising strategy to mitigate the prominent environmental issues.
In fact, to achieve effective CO2 hydrogenation, high temperatures (200–300 °C), elevated pressures (3–10 MPa) and catalysts are industrially necessary to overcoming the high energy barrier. In the CO2 hydrogenation to alcohols process, the major byproducts (CH4 and CO) are more easily generated via the methanation and reverse water–gas shift reaction (RWGS), respectively. In addition, the uncontrollable C–C bond coupling path is another key influencing factor in the synthesis of higher alcohols. Thus, the precise and highly efficient catalysis of alcohol product synthesis from CO2 hydrogenation is a constant challenge. In the past few decades, numerous advances have been made for CO2 hydrogenation to alcohols, including the development of efficient homo- and heterogeneous catalysts, in situ characterization methods and theoretical calculations to study intermediate species and reactor design. The catalyst design and preparation play an important role in catalytic performance, in which the various precious (Pt, Pd, and Rh) and non-precious metals (Cu, Co and Fe), supports (ZnO, Al2O3 and ZrO2) and promoters (K, Ga and Na) have been used for this reaction. Nevertheless, the key factors in determining the catalytic performance are difficult to figure out, because there are many affecting factors. In order to improve the process of CO2 hydrogenation to alcohols, it is necessary to provide an in-depth and more comprehensive understanding, from the reaction mechanisms to the effective preparation catalysts and their systematic investigations.
Herein, we firstly summarize the reaction mechanisms by the thermodynamic and kinetic analyses of CO2 hydrogenation reactions. The discussion is next extended to the structural characteristics of the metal-based and metal-oxide-based catalysts for CO2 hydrogenation to alcohols, in detail and the structure–activity relationship is mainly discussed. Finally, we put forward the current crucial challenges and potential strategies for the future design and characterization of new and more efficient heterogeneous catalysts, as well as the reaction mechanism, so as to further encourage future investigation for this promising field.

2. Reaction Mechanism

2.1. Thermodynamic and Kinetic Analysis

Thermodynamic analysis of CO2 hydrogenation to alcohols is important as it can not only be used to understand chemical reaction mechanisms, but also to optimize reaction conditions. Thermodynamically, CO2 is a quite stable molecule, owing to its high oxidation state and standard Gibbs free energy (−394.38 kJ·mol−1). The introduction of other reactants with higher Gibbs free energy (such as H2) can promote CO2 conversion smoothly [18,19]. In terms of the hydrogenation of CO2 to alcohols, the reaction process steps need to be discussed one by one, owing to the many products, including CO, CH4 and CH3OH and C2+ products (e.g., C2+ alcohols, C2+ alkanes). As shown in Table 1, the direct hydrogenation of CO2 to methanol reaction is an exothermic and volume-reduction process, leading to it being more favorable for reactions in a lower temperature and higher pressure. However, there are some side reactions at the same time, for example, reverse water–gas shift reaction in Entry 2. Although the CO can be hydrogenated to methanol in Entry 3, it needs some certain reaction conditions. From Entry 4, the thermodynamic characteristics of CO2 hydrogenation to C2+ alcohol are similar to the methanol synthesis reaction, which is suitable in a low temperature and high pressure. In addition, the equilibrium constant of alkane synthesis reaction in Entry 5 and 6 is higher than that of synthesis of higher alcohols from CO2 hydrogenation, resulting in the byproduct alkanes [20]. Methanol may also be another byproduct in CO2 hydrogenation to C2+ alcohol reaction.
In the actual catalysis reaction system, the performance of CO2 hydrogenation may be mainly affected by the kinetics process. For instance, although the low reaction temperature is thermodynamically beneficial to methanol synthesis, it will decrease the reaction rate. When raising reaction temperature, the CO by-product will increase via reverse water–gas shift reaction. Given the intricate thermodynamics and kinetics results, there are various possible reaction paths for CO2 hydrogenation to alcohols. Therefore, combing the recent theoretical and experimental advances in mechanistic studies [21,22,23], the main proposed reaction pathways are summarized and proposed as follows in Figure 1, the understanding of which can give important guidance for the rational design of highly efficient catalysts.

2.2. Methanol Synthesis Reaction Pathway

Despite numerous efforts, the reaction mechanism of CO2 hydrogenation to methanol is rather complicated and is yet to be fully understood. For example, when activating CO2 with hydrogen, the O or C atom in CO2 can be hydrogenated into COOH* or HCOO* intermediates, respectively. This results in many different pathways and has aroused great controversy. Based the current research progress, four main routes and corresponding intermediates are discussed, as shown in Figure 1.

2.2.1. HCOO* Pathway

The formate (HCOO*, * denotes the adsorption state) is generated by the reaction between preabsorbed surface H species and the carbon atom in the CO2 molecule at the beginning via a typical Eley–Rideal or Langmuir–Hinshelwood mechanism. Then, the further continuous hydrogenation of HCOO* can form dioxymethylene (H2COO*), formaldehyde (H2CO*), methoxy (H3CO*) intermediate and methanol product, respectively (Figure 1 pathway A). The HCOO* pathway above has been proposed in extensive catalysts, such as partial Cu-based, Pd-based and metal oxide-based catalysts, which are supported by the density functional theory (DFT) calculations and experiments.
Chen et al. [24] proposed that methanol formation follows the formate pathway on the CuZnZr/CuBr2 catalyst, as verified by in situ diffuse reflection infrared Fourier spectrum (DRIFTS) experiments. Lam et al. [25] reported CH3OH formation via formate intermediates on the metal–oxide interface with some specific Lewis acid sites of Cu/Al2O3 catalysts, by means of the DFT calculation and operando spectroscopy experiment. As shown in Figure 2a, Chang et al. [26] also observed the band of HCOO* (1578 cm−1) and H3CO* (2845 cm−1) by the infrared measurement on the Cu-ZrO2 catalyst for CO2 hydrogenation to methanol. Dang et al. [27] found the rate-determining step (RDS) for CH3OH formation on In2O3 is the hydrogenation process of HCOO* to H2COO* by virtue of the DFT calculation. (Figure 2b). Tezsevin et al. [28] investigated the hydrogenation of CO2 to methanol on the surfaces of holmium (Ho)-doped Cu (211) by DFT calculations. The relative energies of H2COO* formation are lower than that of the HCOOH*. This is because the doping of Ho could stabilize H2COO* intermediate instead of HCOOH* intermediate.

2.2.2. R-HCOO* Pathway

In addition to the hydrogenation of HCOO* to H2COO* above, the HCOO* can be hydrogenated to HCOOH*, hydroxymethoxy (H2COO*), following breakage of the C–O bond to H2CO*, to then form H3CO* and methanol, denoted as the r-HCOO* pathway (Figure 1 pathway B). Grabow et al. [29] found that the activation barrier of the HCOO* hydrogenation to H2COO* (1.59 eV) is higher than that of the HCOOH* (0.91 eV), and the formation of CH2O* is inclined to the HCOOH* rather than the H2COO* dissociation on the Cu (111) by DFT results. Cui et al. [30] found the HCOOH* path displays a lower energy barrier of 1.32 eV, which is much lower than that of the H2COO* path on the Cu13/γ-Al2O3 catalyst by DFT calculation. Cai et al. [31] thought the hydrogenation of HCOOH* thermodynamically is favorable on Pd/In2O3 catalysts due to the low Gibbs free energy of (−0.49 eV) the HCOOH* hydrogenation in comparison to that of CO* hydrogenation to CHO* (1.00 eV) in Figure 3. In addition to theoretical calculations, Wang et al. [32] found that the HCOOH hydrogenation to H2CO* on the Cd/TiO2 acts as a rate-determining step for CO2 hydrogenation to methanol by in situ diffuse reflectance infrared Fourier transform spectroscopy and DFT calculations.

2.2.3. RWGS + CO-Hydro Pathway

Although there is the by-product CO of the RWGS reaction in Table 1, the CO can also be hydrogenated to methanol. The RWGS+CO-Hydro pathway is proposed. Specifically, CO2 may be hydrogenated into carboxyl species (COOH*) following the breakage of the C–O bond to CO*, or the CO2 could be directly dissociated to CO*. The methanol can be obtained by continuous CO* hydrogenation (Figure 1 pathway C). Michiels et al. [33] found the CO* intermediate would be converted into HCO*, HCOO* or H2CO*, instead of the formate pathway by the microkinetic model on Cu(111). It has been shown in Figure 4 that Kattel et al. [34] reported the RWGS + CO-Hydro pathway is the main reaction path on Ti3O6H6/Cu(111), because the stable combination of the formate species and surface active sites may poison the active site. Shen et al. [35] found the energy barrier of the CO* hydrogenation to HCO*, as the rate-determining step (RDS) of the RWGS + CO-Hydro pathway, is 1.24 eV, which is much lower than that of the HCOO pathway from the DFT study on Ni4/In2O3 catalysts.

2.2.4. Trans-COOH Pathway

The trans-COOH pathway refers to the CO2 hydrogenation to carboxyl (COOH*), dihydroxycarbene (COHOH*), followed by the breakage of the C–O bond to COH* and then the continuous hydrogenation to CH3OH (Figure 1 pathway D). Tang et al. [36] found the HCOOH intermediate in the HCOO pathway is unstable, and the energy barrier of HCOO* hydrogenation to HCOOH* (1.34 eV) and CO* hydrogenation to HCO* (1.52 eV) are high. However, the energy barrier of the formation of COH* via three isomers of COHOH (0.57 eV) is much lower on Ga3Ni5 (221) surfaces by DFT calculations. Zhao et al. [37] found the CO2 hydrogenation to hydrocarboxyl (trans-COOH) is more favorable than the formate path in the presence of H2O, due to the low energy barrier in methanol synthesis on Cu (111) by DFT.

2.3. Higher Alcohol Synthesis Reaction Pathway

The synthesis of higher alcohols from CO2 hydrogenation is a complex process due to the coexistence of the C–O activation and C–C coupling processes, with difficulty in determining the reaction intermediates. Three widely accepted mechanisms are proposed, including the CO-mediated, the formate-mediated, and the methanol-mediated pathways.

2.3.1. CO Mediated Pathway

The CO-mediated pathway refers to the CO* firstly generated by the RWGS reaction, and then hydrogenated to CHx* and the remaining non-hydrogenated CO* reacts with CHx* species (including CO* insertion into alkyl or alkyl insertion into CO*) for the high alcohols (Figure 1 pathway E). An et al. [38] reported the Co0-Co2+ site on Co/La4Ga2O9 catalyst promoted the CO*, formed by RWGS reaction, dissociation and hydrogenation to CHx*, as well as the insertion of CO* and CHx* to ethanol. Xu et al. [39] recently revealed the Cu-ZnO dual sites were responsible for the non-dissociative activation of C–O to form CO*, and iron carbide improved the dissociation of the C–O bond to form alkyl, while the addition of Cs regulated the hydrogenation ability of the catalyst and promoted the insertion of CO* on the Cs-modified CuFeZn catalyst. According to the in situ DRIFTS analysis, they found that the important intermediates are CH3CHO* (2754 cm−1) and CH3CH2O (29,662,876 cm−1), and the formation of CH3CO follows the CO or CHO insertion mechanism via the CO-mediated pathway (Figure 5).

2.3.2. Formate-Mediated Pathway

The formate mediated pathway, which is different from the CO-mediated pathway, refers to the CHx* reacting with HCOO* rather than the CO* (Figure 1 pathway F). Wang et al. [40] reported the hydrogenation of CO2 to ethanol over a non-noble cobalt catalyst. An unsaturated C–H (CHx*) signal was observed at 2904 cm−1 and formate species at 1373 cm−1 and 1594 cm−1 by operando FT-IR spectroscopy (Figure 6). The bands belonging to CHx* and HCOO* decrease, and the saturated C–H (acetic acid) bands appear. These changes indicate that ethanol is formed by the insertion of CHx* and HCOO*, resulting in the high selectivity of ethanol. Moreover, Wang et al. [41] found the Cu-Cs-ZnO interface promoted the dissociation of HCOOH into CHO*, the C–C coupling of CHO* to ethanol on Cs/Cu/ZnO(000 1 ¯ ), combining DFT calculations and Kinetic Monte Carlo (KMC) simulations.

2.3.3. Methanol-Mediated Pathway

The methanol-mediated pathway refers to the methanol dehydroxylation to the CHx* intermediate, which then reacts with CO* or HCOO* intermediates to form C–C bonds. This process is repeated to grow carbon chains to form the higher alcohols (Figure 1 pathway G). He et al. [42] used 13CH3OH labeling experiments to investigate the reaction path of CO2 hydrogenation to higher alcohols on Pt/Co3O4 catalyst. When adding 13CH3OH, the peak (m/z = 47) attributed to either 13CH3CH2OH or CH313CH2OH was found, indicating the carbon atom of ethanol might be formed from methanol. Not only that, the authors found that water could promote the generation of the higher alcohols. The protonated methanol formed by water could dissociate to CH3*, OH* and H*, in which the CH3* could be further coupled with CO* to CH3CO*, followed by further hydrogenation to ethanol. Yang et al. [43] proposed a similar reaction path for the RhFeLi/TiO2 catalyst, in which the protonate methanol formed by the surface hydroxyls promotes the breakage of the C–O bond and the formation of CH3* by in situ DRIFTS. The CO* obtained by RWGS was inserted into CH3* to form ethanol (Figure 7).

3. Metal-Based Catalyst

Since CO2 is a symmetric linear molecule and a strong electron acceptor, the metal, with the unpaired d electrons in the outermost electron orbitals of metal, can provide electrons for CO2 activation. Additionally, the metal has a unique metallic bond with the d-band holes, which can accept electrons from the hydrogen with an electron donor, so its activation requires the acceptance of electrons. Therefore, metal-based catalysts have been widely applied to the CO2 hydrogenation reaction, especially in the synthesis of alcohols, obtaining good catalytic performance from fundamental research to industrial applications. Metal-based catalysts are mainly divided into the non-noble metal-based catalysts and noble metal-based catalysts.

3.1. Non-Noble Metal-Based Catalysts

The supported non-noble metal catalyst has the advantages of high economic benefit, abundant resources, being easy to obtain and having many preparation methods. Compared with precious metals, non-noble metal catalysts with lower hydrogenation activity may be more conducive to improving selectivity. However, their low activity and poor anti-toxicity are still puzzling researchers. In the process of CO2 hydrogenation to alcohols, non-noble metals, such as Cu and Co, have received extensive attention because of their excellent hydrogen dissociation and C–O bond dissociation ability. However, it is susceptible to sintering and the ability to promote C–C coupling is poor. Therefore, researchers improve it by loading different carriers and doping, so as to achieve high selectivity and excellent stability.

3.1.1. Cu-Based

Generally, copper (Cu), as a common transition metal in nature, has advantages of a low price and good electromagnetic properties (3d104s1). When used for CO2 hydrogenation to alcohols, the Cu usually displays a good ability in the hydrogen dissociate and CO2 activation [44]. However, there are some challenges in Cu-based catalysts, such as easy aggregation and sintering of Cu via Ostwald maturation in harsh reaction conditions [45]. In order to further enhance the catalytic performance, many strategies for the regulation of specific active sites have been carried out, including different levels of copper loading and support, as well as the addition of other metals. Therefore, the metal dispersion, metal alloy and metal–support interaction around the main active sites in the past five years are summarized.

Cu-Based Catalyst for Methanol Synthesis

The preparation method, as an important factor, can usually affect the catalytic active sites and reaction performance. For example, the high Cu dispersion can be obtained by the modification of the preparation method, leading to enhancement of the catalytic activity. Many traditional preparation methods, such as coprecipitation, impregnation, hydrothermal and sol–gel, have been widely studied in the catalysts for methanol synthesis. In recent years, the development of effective catalysts by a simple preparation method has continued to attract the attention. Lei et al. [46] prepared a series of CuO/ZnO/Al2O3 catalysts by direct combustion of citric acid, oxalic acid and urea as fuel, respectively. Among them, CO2 conversion could reach 16.2%, and the methanol yield was up to 10.3% in CuO/ZnO/Al2O3 catalysts with citric acid as fuel, owing to its large specific surface area exposed to Cu. Additionally, a suitable preparation method can regulate the electronic property of Cu, such as the ratio of both Cu0 and Cu+ species. The Cu0 is mainly used as the active site for H2 adsorption and dissociation, while the Cu+ can promote the cleavage of the C–O bond, thus affecting the methanol yield. Yu et al. [47] synthesized the Cu/SiO2 catalyst by ammonia evaporation (AE) and flame spray pyrolysis (FSP), respectively. The methanol selectivity of FSP at the same conversion reached up to 79.3%, which was much higher than that of AE (31.5%). This was because the content of Cu+ in FSP was about five times that of the AE sample. More Cu+ could inhibit the dissociation of CO and change the reaction path from formate to RWGS + CO-Hydro with a low energy barrier by the DRIFTS results (Figure 8). Arena et al. [48] prepared the Cu–ZnO/ZrO2 catalysts with a high total surface area, dispersion and the exposure of the active Cu phase based on reverse co-precipitation under ultrasound irradiation. The optimized catalyst displays a good catalytic performance for the hydrogenation of CO2 to CH3OH in comparison with a commercial Cu–ZnO/Al2O3 methanol synthesis catalyst.
The support (ZnO, Al2O3, ZrO2, etc.) can also change the Cu dispersion, Cu+, oxygen vacancies and the interaction between Cu and support, so as to influence the catalytic performance. Hu et al. [49] found that Al2O3 in supported Cu/Al2O3 catalyst could disperse Cu well and stabilize Cu+ species, leading to a strong CO2 adsorption capacity and the formation of more carbonate, bicarbonate and formate species for CO2 hydrogenation to methanol, in comparison with the Cu/ZnO catalyst. The oxygen vacancy has advantages of regulating the geometric structure and local electrons for CO2 adsorption and activation, which can be formed in the reducible supports, such as TiO2, ZrO2 and ZnO. Zhang et al. [50] thought that the oxygen vacancy with abundant electrons induced a strong metal–support interaction (SMSI) between Cu and TiO2. This could stabilize the Cu nanoparticles and promote the formation of the defective Ti3+ site. The electron was easily transferred to the electron-deficient carbon atom in CO2, forming the HCOO* intermediate, resulting in a lower activation energy and higher CO2 conversion (12.5%), relative to the Cu/TiO2 catalyst without oxygen vacancies (5.73%).
Some recent studies have shown that the interface of composite metal oxides could increase the number of oxygen vacancies. Zuo et al. [51] synthesized a series of ZrO2 and CeO2 solid solutions using a solvent evaporation self-assembly method with the supported Cu-based catalysts, which displayed a higher methanol selectivity than Cu/ZrO2. This is because the doping of Ce could enhance the interaction between Cu and solid solution for the adsorption and activation ability of CO2. Wang et al. [52] took the view that the oxygen vacancy in the ZnO-ZrO2 support could promote CO2 adsorption and activation, and the Cu enhanced the H2 dissociation, leading to the enhancement of carbonate hydrogenation to formate by the DFT calculations and in situ DRIFTS experiments. Moreover, Song et al. [53] prepared the Cu/ZnAl2O4 catalysts with the strong interaction between Cu and ZnAl2O4 support by wet impregnation method, whose CO2 conversion (11.3%) and space-time yield (STY) of methanol (242 gCH3OH kgcat−1 h−1) were higher than Cu/ZnO/Al2O3 catalysts. The authors thought the hydrogen dissociation could occur on both the ZnAl2O4 support and the high dispersion of Cu, which contributed to the formation of HCOO* intermediates with a low energy barrier.
In addition, the additives may modify the active site structure and influence the reaction performance. Tezsevin et al. [28] found Ho metal enhanced the interaction between CO2 and Cu surface, resulting in a low energy barrier of HCOO* formation (−1.02 eV) and the enhanced activation of CO2 on the holmium-doped Cu (211) surface by DFT calculations. Yan et al. [54] found that W metal doping on Cu/CeO2 catalysts could promote the irreversible reduction of Ce4+ to Ce3+ and inhibit the formation of oxygen vacancies, leading to a high STY of 12.3 molMeOH kgcat−1 h−1 by the formate pathway. In the absence of W, the oxygen vacancies on the CeO2 surface were responsible for CO2 dissociation to CO* by the RWGS+CO-Hydro path, whose STY of methanol was only 1.26 molMeOH kgcat−1 h−1. The influence of different doping ratios on the Cu-based catalyst could affect the catalytic performance. Liu et al. [55] found the variation of Ir doping content on the Cu surface changed the rate-determining step of CO2 hydrogenation to methanol. The rate-determining step was the formation of CO+OH on Ir3Cu6(111), while it was the formation of HCO* on Ir6Cu3(111).

Cu-Based Catalyst for Higher Alcohol Synthesis

The key to the higher alcohol synthesis is the coupling C–C bond, alongside the hydrogen dissociate and CO2 activation, compared with the CO2 hydrogenation into methanol. Copper-based catalysts are more conducive to promoting the non-dissociative hydrogenation of the C–O bond than its cleavage, making it easier to produce methanol rather than higher alcohols. Therefore, a series of modified Cu-based catalysts have developed to improve the formation of CHx* species as well as the growth of the carbon chain. An et al. [56] studied the cooperative Cu sites on a Zr12 cluster prepared by the metal–organic framework (MOF). The Cu2 could facilitate the hydrogen dissociation, while the Cu2+ and Cs+ provided an electron-rich environment to stabilize the formyl intermediate. They promoted C–C coupling of methanol and formyl to produce ethanol, leading to a high ethanol selectivity of >99% (Figure 9). Ding et al. [57] prepared a Cu@Na-Beta catalyst by embedding Cu nanoparticles into the Na-Beta zeolite, which displayed a good ethanol selectivity of 79% and yield of 14%. This is because the Cu nanoparticles promote CO2 hydrogenation to form the CH3* intermediate, while the zeolite skeleton limits the formation of byproducts such as methanol and formic acid. They promote the reaction of between CO2 and methyl to form acetyl acid and hydrogenation to ethanol.
In addition, other metals have been introduced on the Cu-based catalysts to promote the adsorption of CO2 by virtue of their interaction. Xu et al. [39] prepared a Cs-modified CuFeZn catalyst with a high C2+OH space-time yield of 73.4 mg gcat−1 h−1. The authors concluded the Cu-ZnO was responsible for the non-dissociative activation of CO*, iron carbide promoted the C–O bond dissociation of CO2 to alkyl intermediate and the addition of Cs improved the insertion of CO*. Zhang et al. [58] found the introduction of Ga could enhance the strong interaction between Cu and Co, and promote the formation of a rich Cu0/Cu+-CoGaOx interface. This interface served as the main active center, leading to a high ethanol selectivity of 23.8% and the ethanol STY of 1.35 mmolEtOH·gcat−1 h−1 at CO2 conversion of 17.8%.

3.1.2. Co-Based

In addition to Cu, other non-noble metals, such as cobalt (Co)-based catalysts, have also been applied for CO2 hydrogenation to alcohols. The electronic configuration of transition metal Co is 3d74s2, which has a stronger electron variation ability compared with Cu. This leads to a good C–O dissociation activation ability, and it is more likely to form methane. Recently, various methods, such as the introduction of the supports and additives, have been taken on Co-based catalysts for CO2 hydrogenation to alcohols.

Co-Based Catalysts for Methanol Synthesis

The interaction between the Co and the supports can adjust the electronic properties and increase the Co dispersion, which is used to inhibit CO2 methanation and boost the methanol selectivity. As shown in the Figure 10, Wang et al. [59] reported Co@Six catalysts with abundant Co-O-SiOn interfaces observed from in situ DRIFTS and XPS, which can inhibit the CH3O* intermediate dehydrogenation of CO and the break of the C-O bond to methane, resulting in a high methanol selectivity of 70.5% and productivity of 3.0 mmol gcat−1 h−1. The Co alloying with other metals is also a significant method to change the Co dispersion and the electronic property. The study by Din et al. [60] prepared a zeolite-based Cu-Co bimetallic catalysts, which displayed the Cu and Co bimetals in zeolite support with good morphology and dispersion. With the increase of Cu content, the number of weakly and moderately basic sites increased; meanwhile, the methanol yield enhanced from 7.5 to 25.6 g Kgcat−1 h−1. This was because the basic sites could significantly affect the adsorption and activation of CO2.

Co-Based Catalyst for Higher Alcohol Synthesis

Given the strong hydrogenation ability of Co-based catalysts, these are more favorable to C–O dissociation activation, while the carbon chain growth becomes hard. The Co-based alloy catalysts by adding other metals, forming the existence of both Co0 and Coδ+, which is considered a good method to inhibit methanation and promote chain growth for improving the catalytic higher alcohol synthesis performance. This is because the Co0 acts as the dissociative activation of CO2 and H2, and the non-dissociative activation of C–O will occur at Coδ+. For example, An et al. [61] designed a series of CoGaxAl2-xO4/SiO2 (x = 0, 0.5, 1.0, 1.5, 2.0) catalysts for CO2 hydrogenation to ethanol. The strong interaction between gallium oxide and cobalt induced the formation of the Co0-Coδ+ active site, which promoted the process of ethanol synthesis by coupling the dissociation (CHx*) and non-dissociation (CO*) of the intermediates. The ethanol selectivity reached up to 20.1% at 270 °C and 3.0 MPa on the CoGa1.0Al1.0O4/SiO2 catalyst. Zheng et al. [62] prepared a Co/La2O3-La4Ga2O9 catalyst, which formed the Co0 and Coδ+ due to the electron interaction between Ga and cobalt, based on the XPS and XRD results. A high CO2 conversion of 9.8% and a selectivity of alcohol of 74.7% on Co/La2O3-La4Ga2O9 catalyst were also obtained. The authors concluded that when CO2 adsorbed on Coδ+, the HCOO* intermediates formed, followed by decomposition into CO*. Meanwhile, CO2 was dissociated and hydrogenated into CH3* on the Co site. The coupling of CO* and CH3* could occur and form CH3CO*, and then it reacts with hydrogen into ethanol. In addition, Wang et al. [63] reported a cobalt catalyst incorporating a nickel species that could stabilize the CHx* intermediates, thus avoiding the formation of methane. Not only that, the presence of Ni also made CO2 hydrogenation form rich HCOO* species. Co promoted the insertion of CHx* species into HCOO* (Figure 11), and then hydrogenation to ethanol with a high yield of 15.8 mmol gcat−1. Therefore, the addition of other metals to Co-based catalysts may also stabilize CHx* intermediates, thereby improving the activity and inhibiting the formation of by-products, such as CH4 and CO.

3.1.3. Others

Many other non-noble metal active components, such as Cd, Fe, and Ni, have been applied in CO2 hydrogenation to alcohols because of their low cost and good catalytic hydrogenation ability. For example, Wang et al. [32] obtained TiO2-supported Cd single atoms, clusters and nanoparticles using the different Cd loadings. Among them, the Cd clusters and TiO2 support could help the stability of the HCOO* intermediate, leading to a lower energy barrier, good methanol selectivity (81%) and CO2 conversion (15.8%), confirmed by the DFT calculations [64] and DRIFTS experiments. Fe has a unique ability to promote C–C coupling, which is usually used for higher alcohol synthesis. Xi et al. [65] prepared K-promoted bimetallic Fe- and In-based catalysts on a Ce-ZrO2 support for CO2 hydrogenation to higher alcohols. When Fe was added, both the RWGS reaction and C–C coupling ability were further enhanced. In addition, the alkaline structure additive K could inhibit the excessive hydrogenation of CO2 to methane, thus improving the selectivity of higher alcohols. The optimized catalyst exhibited high CO2 conversion of 29.6% and alcohol selectivity of 28.7%.

3.2. Noble Metal-Based Catalyst

Noble metals (Pd, Pt, Au, etc.) usually serve as good electron donors or acceptors due to the large atomic radius and the outer electrons being far away from the nucleus. The noble metal-based catalysts have abundant advantages in CO2 activation via the electron interaction and hydrogen dissociation by metal–hydrogen bond, while facing the challenges of limited resources and high price. This results in a strong hydrogenation ability and C–C coupling ability for the hydrogenation of CO2 to alcohols. Therefore, how to reduce the loading of noble metals while maintaining high catalytic activity has become the main research goal.

3.2.1. Pt-Based

The expensive Pt-based catalysts exhibit a strong adsorption and dissociation hydrogen activity and a weak activation ability of CO2. Recently, the different Pt loading, support and additives have been used to further improve the catalytic performance and economic cost.

Pt-Based Catalyst for Methanol Synthesis

The interaction of Pt and supports (In2O3, TiO2, MOF, etc.) has a synergic catalytic effect, which can offer unique active centers for CO2 hydrogenation to methanol. Toyao et al. [66] prepared MoOx/TiO2-supported Pt nanoparticle catalysts for the selective hydrogenation of CO2 to CH3OH. The MoOx with surface defect sites could activate CO2, and Pt as an active metal promoted the formation of carbonate and formate, according to the in situ infrared analysis results, compared with MoOx(30)/TiO2 support, resulting in a high methanol yield of 73%. The author also analyzed the reaction of methanol synthesis with HCOOH and CO as reactants, respectively. As a result, CO hydrogenation to CH4 occurred, and the reaction rate of HCOOH was higher than that of CO2 hydrogenation to methanol. This indicated that CO2 hydrogenation follows the formate pathway. Kuwahara et al. [67] found that a Pt-loaded molybdenum suboxide nanosheet structure showed a high CO2 conversion and methanol yield, which was about 1.35 times higher than that of other molybdenum suboxide supports (including bulk, nanostrip and rod) under similar volume. This is because the high specific surface area of the Pt-loaded molybdenum suboxide nanosheet structure could form more oxygen vacancies for CO2 adsorption and activation and inhibit Pt aggregation. Gutterod and colleagues [68] reported Zr-based UiO-67 MOF encapsulated Pt nanoparticle catalysts with good catalytic performance for CO2 hydrogenation to methanol. Steady-state and transient kinetic studies involving H/D and 13C/12C exchange found that the interface derived from strong interaction between Pt and the linker-deficient Zr6O8 nodes acted as the main active centers to enhance the formation of methanol, in which the Pt was responsible for hydrogen dissociation, and the adsorbed CO2 and hydrogen overflowed would form formate species at the Zr node (Figure 12a).
In addition, the different sizes of Pt have significant effects on the reactivity performance of synthesis methanol. Han et al. [69] found the atomically dispersed Ptn+, formed by the interaction between In2O3 and Pt, could promote the CO2 hydrogenation to methanol, due to the heterolysis of H2 on the atomically dispersed Ptn+ as the Lewis acid. Nevertheless, the Pt nanoparticles were inclined to the RWGS reaction, owing to the homolysis of H2 and formation of CO. Furthermore, Li et al. [70] proposed adjacent Pt monomers on MoS2 support, which exhibited a high catalytic activity compared with the isolated Pt monomers. Through theoretical study, it was found that the main intermediate, COOH* hydrogenation to C(OH)2*, was the main reaction step on the isolated Pt monomer. However, the synergistic effect between two Pt atoms on adjacent Pt monomers could promote COOH* hydrogenation into HCOOH with a low activation energy, resulting in a good catalytic activity (Figure 12b,c).
Figure 12. (a) Schematic presentation of the postulated reaction mechanism of CO2 hydrogenation to the formate intermediate in CH3OH formation at the Pt–Zr node interface [68]. Steps for the addition of an H atom to COOH* over (b) isolated Pt monomers and (c) adjacent Pt on MoS2 [70].
Figure 12. (a) Schematic presentation of the postulated reaction mechanism of CO2 hydrogenation to the formate intermediate in CH3OH formation at the Pt–Zr node interface [68]. Steps for the addition of an H atom to COOH* over (b) isolated Pt monomers and (c) adjacent Pt on MoS2 [70].
Catalysts 14 00232 g012

Pt Base Catalyst for Higher Alcohols

The Pt-based catalyst shows excellent catalytic performance in the synthesis of methanol due to its excellent hydrogen dissociation ability. Not only that, Pt also has the ability to promote the RWGS reaction, and the resultant CO can be used as an intermediate to further generate higher alcohols, thus increasing the selectivity of higher alcohols in the CO2 hydrogenation process. Ouyang and co-workers [71] prepared nanorod and nanoplate Co3O4-supported Pt catalysts (Pt/Co3O4-p and Pt/Co3O4-r), respectively. After hydrogen pretreatment at 200 °C, the part of Co3O4 on Pt/Co3O4-p was reduced to Co and CoO, while the Co3O4 and CoO co-existed on Pt/Co3O4-r. The synergistic effect of between Pt and Co3O4-p oxygen vacancy in Pt/Co3O4-p promoted a higher yield of higher alcohol of 0.56 mmol gcat−1 h−1. Subsequently, Liu et al. [72] further studied the influence of the ordered mesoporous structure of Co3O4-m. In addition to the enhancement of RWGS reaction by means of Pt, the Co metal was conducive to the hydrogen dissociation, and the surface oxygen vacancies formed by unreduced Co3O4 could enhance CO2 adsorption. The ordered mesoporous structure of Co3O4 could help the CHx further form C2+ chains via the growth of carbon chains.

3.2.2. Pd-Based

Similarly, Pd-based catalysts show good dissociation of hydrogen but a weak binding ability between Pd and CO2, which take advantage of the interaction between Pd and supports to regulate the electronic state and geometry structure of active site on Pd-based catalysts for CO2 hydrogenation to alcohols.

Pd-Based Catalyst for Methanol Synthesis

A series of Pd/ZnO catalysts with different Pd particle sizes, varying from 1.6 to 7.9 nm, were synthesized by Zhang et al. [73] using the atomic layer deposition (ALD) technique. With the increment of Pd particle sizes, the Pd/ZnO catalysts displayed a strong metal–support interaction was found in the XPS and H2-TPR results, followed by a higher CH3OH selectivity. This was because the stronger bond between Pd and CO intermediates occurs on the larger the particle size catalysts through the RWGS+ CO-Hydro pathway. In a report by Tian et al. [74], the highly dispersed Pd species reduced the energy barrier of H2 dissociation, thus promoting hydrogen dissociation and the strong metal–support interaction between Pd and MnO/In2O3 promoted hydrogen spillover. The CO2 was adsorbed at the defective oxygen site, forming the formate intermediate, resulting in a good methanol selectivity of 70% at 280 °C. Since the oxygen vacancy is key to CO2 conversion, the promoters were introduced into the Pd-based catalyst to improve the oxygen vacancy. Jiang et al. [75] added a series of promoters, including La, Zr, Cr, Ca, Co, Cd, Zn, Bi and K, on the Pd/CeO2 catalyst by the co-impregnate method. Along with the oxygen vacancy increases, the CO2 conversion showed an approximate linear enhancement, and the maximum presented on the Zr-modified Pd/CeO2 catalyst, owing to the entrance of Zr into the CeO2 lattice (Figure 13).

Pd-Based Catalyst for Higher Alcohols

The metallic Pd size on supported Pd based catalysts can not only influence H2 dissociation, butalso have an important impact on the C–C coupling in the formation of higher alcohols. Caparros et al. [76] prepared a series of different Pd particle sizes on the Pd/Fe2O3 catalysts. For the single-atom Pd/Fe3O4 catalyst, the selectivity towards ethanol reached up to 97.5% at 300 °C and the activity was 413 mmolEtOH gPd−1 h−1. But when the Pd single atoms evolved progressively into Pd nanoparticles, the ethanol selectivity decreased gradually. This was because the interaction between Pd single atoms and Fe3O4 assisted a particular architecture for C-C coupling, and the transformation of CO2 into CO occurred on the Fe3O4 support through the RWGS reaction. In addition, Lou et al. [77] prepared Pd dimers with a uniform Pd2O4 configuration in the Pd2/CeO2 catalyst for CO2 hydrogenation to ethanol. According to the DFT calculation, CO2 inclined to dissociate into CO* on Pd dimer, instead of the COOH* and HCOO* from CO2 hydrogenation with high energy barrier. Since the energy barrier of CO desorption was high, the CO hydrogenation to CH2OH* occurred. However, the dissociation ability of CH2OH* was stronger than in comparison to the CH2OH* hydrogenation methanol, leading to the formation of a CHx* intermediate. The coupling of the strong CO* adsorption and the presence of CHx* intermediate would occur and produce the ethanol. The Pd2/CeO2 catalyst could achieve a high ethanol selectivity of 99.2% and yield of 45.6 gethanol gPd−1 h−1.

3.2.3. Rh-Based

Although rhodium-based catalysts are commonly used for hydrogenation of syngas to ethanol [78], in the hydrogenation of CO2 to alcohols, their catalytic performance might depend largely on the interaction between Rh and supports.

Rh-Based Catalyst for Methanol Synthesis

A recent study by Wang et al. [79] found that the highly dispersed Rh species on the Rh/In2O3 catalyst enhanced the hydrogen dissociation and hydrogen spillover, which promoted the formation of oxygen vacancy and increased the alkalinity of the In2O3 surface, thus promoting the adsorption and activation of CO2. This led to a high CO2 conversion of 17.1%, methanol selectivity of 56.1% and space-time yield of 0.5448 gMeOH h−1 gcat−1 under 300 °C, 5 MPa. In order to enhance the CO2 activation by means of increasing oxygen vacancy, Lu et al. [80] further prepared the In2O3-ZrO2-supported Rh-based catalysts (Rh/In2O3-ZrO2). The addition of ZrO2 could inhibit the reduction of In2O3, boost oxygen vacancy, and stabilize the catalyst structure, resulting in a good 66.5% methanol selectivity and space-time yield of 0.684 gMeOH h−1 gcat−1 at a high CO2 conversion of 18.1% at 300 °C and 5 MPa. Additionally, the oxygen vacancy can be adjusted by the different crystalline phases. Cao et al. [81] prepared a series of Rh-based catalysts with various crystalline phases of TiO2 (p25, anatase and rutile). Among them, the rutile-supported Rh-based catalyst displayed a high methanol selectively, while the major products of p25 and anatase were the methane and CO, respectively (Figure 14). The authors found that the highest content of oxygen vacancy emerged on the rutile-supported Rh-based catalyst, which could be conducive to the formation of methanol by the CO* intermediate.

Rh-Based Catalyst for Higher Alcohols

Although Ru has good hydrogen dissociation ability, the recent studies have focused on the enhancement of the dissociation of C–O bonds and C–C bond coupling for the formation of higher alcohols by means of the additives (e.g., Ti, V, Na and Li) in Rh-supported catalysts. Zheng et al. [82] constructed a Rh1/CeTiOx single-atom catalyst by embedding monoatomic Rh onto a Ti-doped CeO2 support. The catalyst characterization demonstrated that synergistic effects of Ti-doping and monoatomic Rh could generate oxygen-vacancy-Rh Lewis-acid–base pairs, which could improve CO2 activation and C–O bond cleavage of CHxOH* into CHx* and COOH* into CO* species, respectively. This, consequently, led to the remarkable high TOF of 493.1 h−1 activity and 99.1% ethanol selectivity. The Ti-doping also induced Rh–O bond structural reconstruction, which contributed to the improved catalytic stability. Wang et al. [83] prepared the Rh/MCM-41 catalyst with VOx dopants forming the presence of Rh0-Rhx+ groups. Notably, the Rh+ species deriving from the electronic interaction between VOx and Rh could promote CO adsorption and dissociated into CHx*, followed by the CHx* carbonylation to CH3CO* (Figure 15a). Given the synergistic effect on the high dispersion Rh species by the confinement MCM-41 effect and the formation of VOx-Rh interface sites, a good CO2 conversion ~12% and ethanol selectivity ~24% were obtained.
In addition, the introduction of alkali metals into the Rh-based catalysts has been considered as an effective method in regulating the electronic state and intermediate species. For example, Zhang et al. [84] reported the Na-modified Rh nanoparticles embedded in the zeolite Silicalite-1 (Na-Rh@S-1) catalyst showed a high space-time ethanol yield of 72 mmol gRh−1 h−1 for CO2 hydrogenation to ethanol. The Na+ induced the formation of Rh+ and Rh0, which accelerated the CO* dissociation to CHx*, along with the coupling of adsorbed CHx* and CO* to form ethanol (Figure 15b), confirmed by in situ DRIFTS results. On the contrary, there was only the Rh0 species as the main product of the Rh@S-1 catalyst without Na+, which underwent the hydrogen dissociation and CO2 hydrogenation to CO*, followed by the further dissociation and hydrogenation to CHx* and CH4. Kusama et al. [85] reported similar results to the above in the Rh/Al2O3 catalyst, in which the main product of the Rh/Al2O3 catalysts was the methane, while Rh/Al2O3 catalysts with Li modification changed the electronic state of the catalyst, resulting in the increase of intermediate bridge CO* to form the ethanol, with a good ethanol selectivity of 15.5% at 7% CO2 conversion at 513 K and 5 MPa.

3.2.4. Others

Similarly, the other noble metals, such as Ru and Ir, also have a good ability of hydrogen dissociation and CO2 activation. As their methanol selectivity is usually low with the common support, the In2O3 with good methanol activity was used for the support of methanol synthesis due to the strong metal–support interaction. Wu et al. [86] found that the oxygen vacancy derived from hydrogen spillover from the Ru/In2O3 catalyst could promote CO2 adsorption and activation by the strong metal–carrier interaction between Ru and In2O3 support. Combined by the DFT and energy barrier analysis, this oxygen vacancy enhances the direct decomposition of CO2 into CO*, accompanied by the continuous hydrogenation to CH3OH with a methanol space-time yield of 0.57 gmethanol gcat−1 h−1 at 300 °C. Xiong et al. [87] further improved Ru/In2O3 catalyst by adding ZrO2 in order to form the solid solution of ZrO2 and In2O3. The formed In-Ov-Zr oxygen vacancy structure serves as the main active site, leading to a higher the space-time yield, which increased by more than 20%, in comparison with the Ru/In2O3 catalyst. In addition, Shen et al. [88] designed a series of Ir/In2O3 catalysts with different Ir loading for methanol synthesis from CO2 hydrogenation, which displayed a high dispersion Ir and oxygen vacancy of In2O3 by means of the interaction between iridium and In2O3. As the Ir content increased, the higher activity was obtained, because it could promote the formation of oxygen vacancies and stabilize them. The optimized Ir/In2O3 catalyst showed a high methanol selectivity over 70% and space-time yield of 0.765 gMeOH h−1 gcat−1 at CO2 conversion of 17.7% under the reaction conditions of 300 °C and 5 MPa. The metal Ir promoted hydrogen dissociation and In2O3 reduction to produce oxygen vacancies, resulting in altering the formate route of In2O3 to the RWGS+CO-Hydro pathway with a lower energy barrier on Ir/In2O3 catalyst.

4. Metal Oxide-Based Catalyst

As a result of their oxygen vacancy and reducibility, the metal oxides are used to adsorb and activate CO2 for CO2 hydrogenation reaction. Recently, the metal oxide catalysts have been extensively studied in the hydrogenation of CO2 to produce alcohols.

4.1. In2O3-Based

In2O3 catalysts are prone to form the oxygen vacancy for CO2 adsorption and activation, as well as the InOx for hydrogen dissociation, thus showing high activity and good stability in the CO2 hydrogenation reaction.

4.1.1. In2O3-Based Catalyst for Methanol Synthesis

In the methanol synthesis process, the other support is usually introduced because their interaction can improve the dispersion and prevent sintering. Zhang et al. [89] investigated In2O3/m-ZrO2 and the CO2 hydrogenation reaction pathway under reaction conditions. They pointed out that m-ZrO2 as support could promote the reduction of In2O3 to InOx rather than In0. There are more oxygen vacancies on the In2O3/m-ZrO2, owing to the weak In-O bond, weakened by the strong interaction between ZrO2 and In2O3. Not only that, when the concentration of In on the surface was low, InOx (0 < x < 1.5) would “dissolve” into the m-ZrO2 subsurface, resulting in the formation of CO via RWGS. As the content of In increased, the content of InOx increased, which promoted CO2 hydrogenation to methanol by the HCOO* in Figure 16a,b. A study was reported by Shi et al. [90], where they synthesized rod In2O3 catalysts modified by different contents of graphene oxide (GO) by precipitation method for CO2 hydrogenation to methanol. GO would increase the dispersion of In2O3 and convert cubic In2O3 (440) into hexagonal In2O3 (110). These two different forms of In2O3 would form c-In2O3(440)/h-In2O3(110) homojunction and interaction. The results of DFT showed that the energy barrier forming an oxygen vacancy could be reduced by homojunction. Hydrogenation of CO2 on the oxygen vacancy would form formates and methanol. The selectivity of methanol (over 76%) and the conversion of CO2 (10.4%) on the optimized In2O3/GO catalyst were both improved compared with the In2O3 catalyst.
In addition, the introduction of metals is used to increase hydrogen dissociation and improve the performance of the In2O3 catalyst for CO2 hydrogenation due to its weak hydrogen dissociation ability. Fang et al. [91] prepared an In2O3/Co/C-N catalyst, in which the Co metal promoted the hydrogen activation, the oxygen vacancy acted on the active site for CO2 adsorption, the presence of N enhanced the electron interaction between Co and In2O3, preventing the In2O3 sintering. As a result, a high CO2 conversion rate of 9.4% and methanol selectivity rate of 88.4% were obtained via the HCOO* pathway. Tian et al. [92] found the addition of Pd to the In2O3 catalyst could promote H2 dissociation, thus providing sufficient H atoms for CO2 hydrogenation. Experiments confirmed that there were abundant oxygen vacancies in In2O3, which could be used as the active site for CO2 activation, as well as prevent Pd sintering. The methanol yield of the Pd-based catalyst (331.7 gMeOH kgcat−1 h−1) was much higher than that of pure In2O3 (92.8 gMeOH kgcat−1 h−1). Therefore, the catalytic activity was significantly improved through the synergistic effect between In2O3 and the active metal Pd. Rui et al. [93] reported an In2O3-supported Au catalyst, which exhibited a high space-time yield of methanol of 0.47 gMeOH/(h·gcat) at 300 °C, 5 MPa, and 21,000 cm3 h−1 gcat−1. The Auδ+−In2O3−x interface formed by electron transfer between Au and In2O3 behaved as the active centers (Figure 16c,d), in which Auδ+ was responsible for H2 dissociation and In2O3−x with abundant oxygen vacancies promoted CO2 activation.
Figure 16. (a) Schematic diagram and (b) performance diagram of CO2 hydrogenation on In2O3/m-ZrO2 catalyst [89]. (c) In Auger peaks and (d)Au 4f peaks for the Au/In2O3 catalyst [93].
Figure 16. (a) Schematic diagram and (b) performance diagram of CO2 hydrogenation on In2O3/m-ZrO2 catalyst [89]. (c) In Auger peaks and (d)Au 4f peaks for the Au/In2O3 catalyst [93].
Catalysts 14 00232 g016

4.1.2. In2O3-Based Catalyst for Higher Alcohol Synthesis

Although the In2O3-based catalyst has advantages of H2 dissociation and oxygen vacancy for CO2 adsorption toward CO2 hydrogenation to methanol, both the C–O dissociation activation and carbon chain growth still remain challenges. Recently, some metals have been introduced into the In2O3 catalyst to solve the question above for the hydrogenation of CO2 to higher alcohols. Ye et al. [94] designed a Ir1−In2O3 single-atom catalyst with a high ethanol selectivity (>99%) and initial turnover frequency (481 h−1). The authors thought that single-atom Ir, formed by the oxygen vacancy on In2O3, could promote the CO2 and activation by the Ir-C bond and oxygen vacancy adsorption on the O in CO2 molecule (Figure 17). According to DFT calculations, the energy barrier of CO2 dissociation to CO* on Ir/In2O3 was 0.48 eV, which was lower than the pure In2O3 of 0.77 eV. In addition, the Ir could promote the hydrogen dissociation and form the CH3O* intermediates, followed by the coupling of CO* and CH3O* to form ethanol. Witoon et al. [95] found that adding K-Co in the In2O3 catalyst could form the KCoO2 and CoO. The CoO could promote the non-dissociative activation of C–O to form CO*, while the formation of K–O–Co could inhibit the hydrogenation of CxHy* intermediates to form hydrocarbons by means of enhancing the strong adsorption of H2. As a result, the K-Co-modified In2O3 catalyst showed a good higher alcohol space-time yield of 169.6 g kgcat−1h−1.

4.2. Solid Solution

In recent years, bimetallic composite oxide catalysts (In2O3–ZrO2, ZnO-ZrO2, CexZr1-xO2) have attracted much attention for CO2 hydrogenation to alcohols because of their excellent stability, abundant oxygen vacancy and unique synergistic effect of metals. Tada et al. [96] studied the effect of different Zn contents on ZnxZr1−xO2−x catalysts. With the increase of Zn content, the main active site converted from the Zn–O–Zr sites to the ZnO nanoparticles, confirmed by the results of catalytic performance evaluation and structural analysis. In the Zn–O–Zr site, the Zr could promote the CO2 activation, while the Zn was responsible for the hydrogen dissociation. Their synergistic effect resulted in a high methanol yield of 1.9 mmol h−1 g−1 at 300 °C by the formed formates and methoxys. Furthermore, Sha et al. [97] prepared the Ga-modified ZnZrOx (GaZnZrOx) solid solution catalyst, which showed a 8.8% CO2 conversion and 630 mg gcat−1 h−1 methanol space-time yield at 320 °C. The addition of Ga promoted the adsorption and dissociation of H2 and increased the amount of oxygen vacancy for CO2 activation, thus accelerating the hydrogenation of HCOO* to CH3O* by the formate pathway. In addition, the weak hydrogen dissociation ability of the ZnZrOx could be enhanced by introducing metal, thus promoting the conversion of CO2. Lee et al. [98] studied a Pd-promoted ZnZrOx solid solution catalyst for CO2 hydrogenation to methanol. The atomically dispersed Pd, induced by the ZnZrOx anchoring, contributed to the production of more surface oxygen vacancies and hydrogen dissociation. Moreover, the presence of Pd accelerated the formation of HCOO* and CH3O* intermediates through HCOO pathway, resulting in a low apparent activation energy and higher STYMeOH of 0.69 g gcat−1 h−1, in comparison to the ZnZrOx of 0.55 g gcat−1 h−1.

5. Conclusions

This overview highlights the latest research developments in the catalyst synthetic schemes and structural properties of metal-based and metal oxide-based catalysts for CO2 hydrogenation reaction, as well as their effect on reaction mechanisms. First of all, we have reported and discussed the important role of reaction mechanisms and catalyst design for methanol and higher alcohol synthesis from CO2 hydrogenation based on the main metal-based and metal oxide-based catalysts. The metal dispersion, support modification, promoter addition and their structural optimization in supported heterogeneous catalyst systems has great influence on the CO2 and hydrogen molecule adsorption, activation, hydrogenation reaction and desorption of reactants and intermediates, thus determining the catalytic performance of catalysts in methanol and higher alcohol formation from CO2 hydrogenation. Moreover, based on the thermodynamics analysis and reaction mechanism for methanol synthesis (HCOO*, r-HCOO*, RWGS+CO-Hydro and trans-COOH pathways), as well as the formation higher alcohols (CO-, formate- and methanol-mediated pathways) in detail, many advances in comprehension and design of active centers for CO2 hydrogenation have been proposed, such as the metal dispersion, metal alloy, metal–support interaction, geometry and electronic properties, as well as the defects and coordination structure. However, there are still numerous challenges in accurately and comprehensively comprehending the active centers, owing to the complexity and obstacles of heterogeneous catalysts. Aiming at improving the performance of metal-based and metal oxide-based catalysts, several strategies in catalyst preparation, characterization techniques and theoretical calculations needs to be further developed in the future. Therefore, this review could put forward valuable guidelines for the design of efficient catalysts based on the great advances in understanding and improving active centers of heterogeneous catalysts, such as, enhancing CO2 activation via controlling the oxygen vacancies and physicochemical features of supports, promoting the dissociative adsorption and spillover of H2 by the metal size and stabilizing the key intermediates for the CO2 hydrogenation to alcohols reaction.

Author Contributions

Conceptualization and writing—original draft preparation, G.C. and Y.L. (Yingjie Lou); figure and table preparation, M.Z. and Y.L. (Yuming Li); writing—review and editing, G.J. and C.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (22109177), the Young Elite Scientists Sponsorship Program by CAST (2023QNRC001), and the Carbon Neutrality Research Institute Fund (20230303).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. Koytsoumpa, E.I.; Bergins, C.; Kakaras, E. The CO2 economy: Review of CO2 capture and reuse technologies. J. Supercrit. Fluids 2018, 132, 3–16. [Google Scholar] [CrossRef]
  2. Divins, N.J.; Kordus, D.; Timoshenko, J.; Sinev, I.; Zegkinoglou, I.; Bergmann, A.; Chee, S.W.; Widrinna, S.; Karslioglu, O.; Mistry, H.; et al. Operando high-pressure investigation of size-controlled CuZn catalysts for the methanol synthesis reaction. Nat. Commun. 2021, 12, 1435. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, J.; Guo, C.; Fang, S.; Zhao, X.; Li, L.; Jiang, H.; Liu, Z.; Fan, Z.; Xu, W.; Xiao, J.; et al. Accelerating electrochemical CO2 reduction to multi-carbon products via asymmetric intermediate binding at confined nanointerfaces. Nat. Commun. 2023, 14, 1298. [Google Scholar] [CrossRef] [PubMed]
  4. Yuan, L.; Hung, S.-F.; Tang, Z.-R.; Chen, H.M.; Xiong, Y.; Xu, Y.-J. Dynamic Evolution of Atomically Dispersed Cu Species for CO2 Photoreduction to Solar Fuels. ACS Catal. 2019, 9, 4824–4833. [Google Scholar] [CrossRef]
  5. Das, S.; Perez-Ramirez, J.; Gong, J.; Dewangan, N.; Hidajat, K.; Gates, B.C.; Kawi, S. Core-shell structured catalysts for thermocatalytic, photocatalytic, and electrocatalytic conversion of CO2. Chem. Soc. Rev. 2020, 49, 2937–3004. [Google Scholar] [CrossRef] [PubMed]
  6. Li, Y.; Zheng, S.; Liu, H.; Xiong, Q.; Yi, H.; Yang, H.; Mei, Z.; Zhao, Q.; Yin, Z.-W.; Huang, M.; et al. Sequential co-reduction of nitrate and carbon dioxide enables selective urea electrosynthesis. Nat. Commun. 2024, 15, 176. [Google Scholar] [CrossRef] [PubMed]
  7. Li, Z.; Wu, W.; Wang, M.; Wang, Y.; Ma, X.; Luo, L.; Chen, Y.; Fan, K.; Pan, Y.; Li, H.; et al. Ambient-pressure hydrogenation of CO2 into long-chain olefins. Nat. Commun. 2022, 13, 2396. [Google Scholar] [CrossRef] [PubMed]
  8. Hung, S.F.; Xu, A.; Wang, X.; Li, F.; Hsu, S.H.; Li, Y.; Wicks, J.; Cervantes, E.G.; Rasouli, A.S.; Li, Y.C.; et al. A metal-supported single-atom catalytic site enables carbon dioxide hydrogenation. Nat. Commun. 2022, 13, 819. [Google Scholar] [CrossRef] [PubMed]
  9. 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]
  10. Wang, K.; Zhu, Y.; Gu, M.; Hu, Z.; Chang, Y.-C.; Pao, C.-W.; Xu, Y.; Huang, X. A Derivative of ZnIn2S4 Nanosheet Supported Pd Boosts Selective CO2 Hydrogenation. Adv. Funct. Mater. 2023, 33, 2215148. [Google Scholar] [CrossRef]
  11. Melaet, G.; Ralston, W.T.; Li, C.-S.; Alayoglu, S.; An, K.; Musselwhite, N.; Kalkan, B.; Somorjai, G.A. Evidence of Highly Active Cobalt Oxide Catalyst for the Fischer–Tropsch Synthesis and CO2 Hydrogenation. J. Am. Chem. Soc. 2014, 136, 2260–2263. [Google Scholar] [CrossRef] [PubMed]
  12. Pastor-Perez, L.; Shah, M.; le Sache, E.; Reina, T.R. Improving Fe/Al2O3 Catalysts for the Reverse Water-Gas Shift Reaction: On the Effect of Cs as Activity/Selectivity Promoter. Catalysts 2018, 8, 608. [Google Scholar] [CrossRef]
  13. Yang, H.; Zhang, C.; Gao, P.; Wang, H.; Li, X.; Zhong, L.; Wei, W.; Sun, Y. A review of the catalytic hydrogenation of carbon dioxide into value-added hydrocarbons. Catal. Sci. Technol. 2017, 7, 4580–4598. [Google Scholar] [CrossRef]
  14. Liu, S.; Bian, K.; Yang, H.; Sun, M.; Nie, X.; Zhang, X.; Hou, S.; Zhang, G.; Song, C.; Guo, X. Methane transformation into light olefins using methanol as co-reactant over Mo/HZSM-5. Chem. Eng. Sci. 2024, 283, 119431. [Google Scholar] [CrossRef]
  15. Wang, X.; Gao, X.; Song, F.; Wang, X.; Cao, G.; Zhang, J.; Han, Y.; Zhang, Q. Direct Oxidation of Methanol to Polyoxymethylene Dimethyl Ethers over FeMo@HZSM-5 Core–Shell Catalyst. ACS Catal. 2024, 14, 1093–1097. [Google Scholar] [CrossRef]
  16. Ye, Y.; Abou-Hamad, E.; Gong, X.; Shoinkhorova, T.B.; Dokania, A.; Gascon, J.; Chowdhury, A.D. Mapping the Methanol-to-Gasoline Process Over Zeolite Beta. Angew. Chem. Int. Edit. 2023, 62, e202303124. [Google Scholar] [CrossRef] [PubMed]
  17. 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]
  18. 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]
  19. Sancho-Sanz, I.; Korili, S.A.; Gil, A. Catalytic valorization of CO2 by hydrogenation: Current status and future trends. Catal. Rev. Sci. Eng. 2021, 65, 698–772. [Google Scholar] [CrossRef]
  20. Hussain, I.; Jalil, A.A.; Izan, S.M.; Azami, M.S.; Kidam, K.; Ainirazali, N.; Ripin, A. Thermodynamic and experimental explorations of CO2 methanation over highly active metal-free fibrous silica-beta zeolite (FS@SiO2-BEA) of innovative morphology. Chem. Eng. Sci. 2021, 229, 116015. [Google Scholar] [CrossRef]
  21. 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] [PubMed]
  22. Li, Y.; Chan, S.H.; Sun, Q. Heterogeneous catalytic conversion of CO2: A comprehensive theoretical review. Nanoscale 2015, 7, 8663–8683. [Google Scholar] [CrossRef] [PubMed]
  23. Martínez-Suárez, L.; Siemer, N.; Frenzel, J.; Marx, D. Reaction Network of Methanol Synthesis over Cu/ZnO Nanocatalysts. ACS Catal. 2015, 5, 4201–4218. [Google Scholar] [CrossRef]
  24. Chen, S.; Zhang, J.; Song, F.; Zhang, Q.; Yang, G.; Zhang, M.; Wang, X.; Xie, H.; Tan, Y. Induced high selectivity methanol formation during CO2 hydrogenation over a CuBr2-modified CuZnZr catalyst. J. Catal. 2020, 389, 47–59. [Google Scholar] [CrossRef]
  25. Lam, E.; Jose Corral-Perez, J.; Larmier, K.; Noh, G.; Wolf, P.; Comas-Vives, A.; Urakawa, A.; Coperet, C. CO2 Hydrogenation on Cu/Al2O3: Role of Metal/Support Interface in Driving Activity and Selectivity of a Bifunctional Catalyst. Angew. Chem. Int. Edit. 2019, 58, 13989–13996. [Google Scholar] [CrossRef] [PubMed]
  26. Chang, X.; Han, X.; Pan, Y.; Hao, Z.; Chen, J.; Li, M.; Lv, J.; Ma, X. Insight into the Role of Cu-ZrO2 Interaction in Methanol Synthesis from CO2 Hydrogenation. Ind. Eng. Chem. Res. 2022, 61, 6872–6883. [Google Scholar] [CrossRef]
  27. Dang, S.; Qin, B.; Yang, Y.; Wang, H.; Cai, J.; Han, Y.; Li, S.; Gao, P.; Sun, Y. Rationally designed indium oxide catalysts for CO2 hydrogenation to methanol with high activity and selectivity. Sci. Adv. 2020, 6, eaaz2060. [Google Scholar] [CrossRef]
  28. Tezsevin, I.; Senkan, S.; Onal, I.; Duzenli, D. DFT Study on the Hydrogenation of CO2 to Methanol on Ho-Doped Cu(211) Surface. J. Phys. Chem. C 2020, 124, 22426–22434. [Google Scholar] [CrossRef]
  29. Grabow, L.C.; Mavrikakis, M. Mechanism of Methanol Synthesis on Cu through CO2 and CO Hydrogenation. ACS Catal. 2011, 1, 365–384. [Google Scholar] [CrossRef]
  30. Cui, Z.; Meng, S.; Yi, Y.; Jafarzadeh, A.; Li, S.; Neyts, E.C.; Hao, Y.; Li, L.; Zhang, X.; Wang, X.; et al. Plasma-Catalytic Methanol Synthesis from CO2 Hydrogenation over a Supported Cu Cluster Catalyst: Insights into the Reaction Mechanism. ACS Catal. 2022, 12, 1326–1337. [Google Scholar] [CrossRef]
  31. Cai, Z.; Huang, M.; Dai, J.; Zhan, G.; Sun, F.-l.; Zhuang, G.-L.; Wang, Y.; Tian, P.; Chen, B.; Ullah, S.; et al. Fabrication of Pd/In2O3 Nanocatalysts Derived from MIL-68(In) Loaded with Molecular Metalloporphyrin (TCPP(Pd)) Toward CO2 Hydrogenation to Methanol. ACS Catal. 2022, 12, 709–723. [Google Scholar] [CrossRef]
  32. Wang, J.; Meeprasert, J.; Han, Z.; Wang, H.; Feng, Z.; Tang, C.; Sha, F.; Tang, S.; Li, G.; Pidko, E.A.; et al. Highly dispersed Cd cluster supported on TiO2 as an efficient catalyst for CO2 hydrogenation to methanol. Chin. J. Catal. 2022, 43, 761–770. [Google Scholar] [CrossRef]
  33. Michiels, R.; Engelmann, Y.; Bogaerts, A. Plasma Catalysis for CO2 Hydrogenation: Unlocking New Pathways toward CH3OH. J. Phys. Chem. C 2020, 124, 25859–25872. [Google Scholar] [CrossRef]
  34. Kattel, S.; Yan, B.; Yang, Y.; Chen, J.G.; Liu, P. Optimizing Binding Energies of Key Intermediates for CO2 Hydrogenation to Methanol over Oxide-Supported Copper. J. Am. Chem. Soc. 2016, 138, 12440–12450. [Google Scholar] [CrossRef]
  35. Shen, C.; Bao, Q.; Xue, W.; Sun, K.; Zhang, Z.; Jia, X.; Mei, D.; Liu, C.-j. Synergistic effect of the metal-support interaction and interfacial oxygen vacancy for CO2 hydrogenation to methanol over Ni/In2O3 catalyst: A theoretical study. J. Energy Chem. 2022, 65, 623–629. [Google Scholar] [CrossRef]
  36. Tang, Q.; Shen, Z.; Huang, L.; He, T.; Adidharma, H.; Russell, A.G.; Fan, M. Synthesis of methanol from CO2 hydrogenation promoted by dissociative adsorption of hydrogen on a Ga3Ni5(221) surface. Phys. Chem. Chem. Phys. 2017, 19, 18539–18555. [Google Scholar] [CrossRef] [PubMed]
  37. Zhao, Y.-F.; Yang, Y.; Mims, C.; Peden, C.H.F.; Li, J.; Mei, D. Insight into methanol synthesis from CO2 hydrogenation on Cu(111): Complex reaction network and the effects of H2O. J. Catal. 2011, 281, 199–211. [Google Scholar] [CrossRef]
  38. An, K.; Zhang, S.; Wang, J.; Liu, Q.; Zhang, Z.; Liu, Y. A highly selective catalyst of Co/La4Ga2O9 for CO2 hydrogenation to ethanol. J. Energy Chem. 2021, 56, 486–495. [Google Scholar] [CrossRef]
  39. Xu, D.; Ding, M.; Hong, X.; Liu, G.; Tsang, S.C.E. Selective C2+ Alcohol Synthesis from Direct CO2 Hydrogenation over a Cs-Promoted Cu-Fe-Zn Catalyst. ACS Catal. 2020, 10, 5250–5260. [Google Scholar] [CrossRef]
  40. 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. Edit. 2018, 57, 6104–6108. [Google Scholar] [CrossRef]
  41. Wang, X.; Ramirez, P.J.; Liao, W.; Rodriguez, J.A.; Liu, P. Cesium-Induced Active Sites for C-C Coupling and Ethanol Synthesis from CO2 Hydrogenation on Cu/ZnO(000(1)over-bar) Surfaces. J. Am. Chem. Soc. 2021, 143, 13103–13112. [Google Scholar] [CrossRef] [PubMed]
  42. He, Z.; Qian, Q.; Ma, J.; Meng, Q.; Zhou, H.; Song, J.; Liu, Z.; Han, B. Water-Enhanced Synthesis of Higher Alcohols from CO2 Hydrogenation over a Pt/Co3O4 Catalyst under Milder Conditions. Angew. Chem. Int. Ed. Engl. 2016, 55, 737–741. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, C.; Mu, R.; Wang, G.; Song, J.; Tian, H.; Zhao, Z.-J.; Gong, J. Hydroxyl-mediated ethanol selectivity of CO2 hydrogenation. Chem. Sci. 2019, 10, 3161–3167. [Google Scholar] [CrossRef] [PubMed]
  44. 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-Environ. 2021, 291, 120073. [Google Scholar] [CrossRef]
  45. Chen, K.; Yu, J.; Liu, B.; Si, C.; Ban, H.; Cai, W.; Li, C.; Li, Z.; Fujimoto, K. Simple strategy synthesizing stable CuZnO/SiO2 methanol synthesis catalyst. J. Catal. 2019, 372, 163–173. [Google Scholar] [CrossRef]
  46. Lei, H.; Hou, Z.; Xie, J. Hydrogenation of CO2 to CH3OH over CuO/ZnO/Al2O3 catalysts prepared via a solvent-free routine. Fuel 2016, 164, 191–198. [Google Scholar] [CrossRef]
  47. Yu, J.; Yang, M.; Zhang, J.; Ge, Q.; Zimina, A.; Pruessmann, T.; Zheng, L.; Grunwaldt, J.-D.; Sun, J. Stabilizing Cu+ in Cu/SiO2 Catalysts with a Shattuckite-Like Structure Boosts CO2 Hydrogenation into Methanol. ACS Catal. 2020, 10, 14694–14706. [Google Scholar] [CrossRef]
  48. Arena, F.; Barbera, K.; Italiano, G.; Bonura, G.; Spadaro, L.; Frusteri, F. Synthesis, characterization and activity pattern of Cu-ZnO/ZrO2 catalysts in the hydrogenation of carbon dioxide to methanol. J. Catal. 2007, 249, 185–194. [Google Scholar] [CrossRef]
  49. Hu, J.; Li, Y.; Zhen, Y.; Chen, M.; Wan, H. In situ FTIR and ex situ XPS/HS-LEIS study of supported Cu/Al2O3 and Cu/ZnO catalysts for CO2 hydrogenation. Chin. J. Catal. 2021, 42, 367–375. [Google Scholar] [CrossRef]
  50. Zhang, C.; Wang, L.; Jerome Etim, U.; Song, Y.; Gazit, O.M.; Zhong, Z. Oxygen Vacancies in Cu/TiO2 Boost Strong Metal-Support Interaction and CO2 Hydrogenation to Methanol. J. Catal. 2022, 413, 284–296. [Google Scholar] [CrossRef]
  51. Zuo, J.; Na, W.; Zhang, P.; Yang, X.; Wen, J.; Zheng, M.; Wang, H. Enhanced activity of CexZr1−xO2 solid solutions supported Cu-based catalysts for hydrogenation of CO2 to methanol. Mol. Catal. 2022, 526, 112357. [Google Scholar] [CrossRef]
  52. Wang, Y.; Kattel, S.; Gao, W.; Li, K.; Liu, P.; Chen, J.G.; Wang, H. Exploring the ternary interactions in Cu-ZnO-ZrO2 catalysts for efficient CO2 hydrogenation to methanol. Nat. Commun. 2019, 10, 1166. [Google Scholar] [CrossRef] [PubMed]
  53. Song, L.; Wang, H.; Wang, S.; Qu, Z. Dual-site activation of H2 over Cu/ZnAl2O4 boosting CO2 hydrogenation to methanol. Appl. Catal. B-Environ. 2023, 322, 122137. [Google Scholar] [CrossRef]
  54. Yan, Y.; Wong, R.J.; Ma, Z.; Donat, F.; Xi, S.; Saqline, S.; Fan, Q.; Du, Y.; Borgna, A.; He, Q.; et al. CO2 hydrogenation to methanol on tungsten-doped Cu/CeO2 catalysts. Appl. Catal. B-Environ. 2022, 306, 121098. [Google Scholar] [CrossRef]
  55. Liu, L.; Wang, C.; Xue, F.; Li, J.; Zhang, H.; Lu, S.; Su, X.; Cao, B.; Huo, W.; Fang, T. DFT investigation of CO2 hydrogenation to methanol over Ir-doped Cu surface. Mol. Catal. 2022, 528, 112460. [Google Scholar] [CrossRef]
  56. An, B.; Li, Z.; Song, Y.; Zhang, J.; Zeng, L.; Wang, C.; Lin, W. Cooperative copper centres in a metal-organic framework for selective conversion of CO2 to ethanol. Nat. Catal. 2019, 2, 709–717. [Google Scholar] [CrossRef]
  57. Ding, L.; Shi, T.; Gu, J.; Cui, Y.; Zhang, Z.; Yang, C.; Chen, T.; Lin, M.; Wang, P.; Xue, N.; et al. CO2 Hydrogenation to Ethanol over Cu@Na-Beta. Chem 2020, 6, 2673–2689. [Google Scholar] [CrossRef]
  58. Zhang, G.; Fan, G.; Zheng, L.; Li, F. Ga-Promoted CuCo-Based Catalysts for Efficient CO2 Hydrogenation to Ethanol: The Key Synergistic Role of Cu-CoGaOxInterfacial Sites. ACS Appl. Mater. Interfaces 2022, 14, 35569–35580. [Google Scholar] [CrossRef]
  59. 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]
  60. Din, I.U.; Alotaibi, M.A.; Alharthi, A.I.; Al-Shalwi, M.N.; Alshehri, F. Green synthesis approach for preparing zeolite based Co-Cu bimetallic catalysts for low temperature CO2 hydrogenation to methanol. Fuel 2022, 330, 125643. [Google Scholar] [CrossRef]
  61. An, K.; Zhang, S.; Wang, H.; Li, N.; Zhang, Z.; Liu, Y. Co0-Coδ+ active pairs tailored by Ga-Al-O spinel for CO2-to-ethanol synthesis. Chem. Eng. J. 2022, 433, 134606. [Google Scholar] [CrossRef]
  62. Zheng, J.; An, K.; Wang, J.; Li, J.; Liu, Y. Direct synthesis of ethanol via CO2 hydrogenation over the Co/La-Ga-O composite oxide catalyst. J. Fuel Chem. Technol. 2019, 47, 697–708. [Google Scholar] [CrossRef]
  63. 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]
  64. Li, G.; Meeprasert, J.; Wang, J.; Li, C.; Pidko, E.A. CO2 Hydrogenation to Methanol over Cd4/TiO2 Catalyst: Insight into Multifunctional Interface. ChemCatChem 2022, 14, e202101646. [Google Scholar] [CrossRef] [PubMed]
  65. Xi, X.; Zeng, F.; Zhang, H.; Wu, X.; Ren, J.; Bisswanger, T.; Stampfer, C.; Hofmann, J.P.; Palkovits, R.; Heeres, H.J. CO2 Hydrogenation to Higher Alcohols over K-Promoted Bimetallic Fe-In Catalysts on a Ce-ZrO2 Support. ACS Sustain. Chem. Eng. 2021, 9, 6235–6249. [Google Scholar] [CrossRef]
  66. Toyao, T.; Kayamori, S.; Maeno, Z.; Siddiki, S.M.A.H.; Shimizu, K.-i. Heterogeneous Pt and MoOxCo-Loaded TiO2 Catalysts for Low-Temperature CO2 Hydrogenation To Form CH3OH. ACS Catal. 2019, 9, 8187–8196. [Google Scholar] [CrossRef]
  67. Kuwahara, Y.; Jinda, A.; Ge, H.; Hamahara, K.; Yamashita, H. CO2 Hydrogenation to Methanol over a Pt-Loaded Molybdenum Suboxide Nanosheet with Abundant Surface Oxygen Vacancies. J. Phys. Chem. C 2023, 127, 4942–4952. [Google Scholar] [CrossRef]
  68. Gutterod, E.S.; Lazzarini, A.; Fjermestad, T.; Kaur, G.; Manzoli, M.; Bordiga, S.; Svelle, S.; Lillerud, K.P.; Skulason, E.; Oien-Odegaard, S.; et al. Hydrogenation of CO2 to Methanol by Pt Nanoparticles Encapsulated in UiO-67: Deciphering the Role of the Metal-Organic Framework. J. Am. Chem. Soc. 2020, 142, 999–1009. [Google Scholar] [CrossRef] [PubMed]
  69. Han, Z.; Tang, C.; Wang, J.; Li, L.; Li, C. Atomically dispersed Ptn+ species as highly active sites in Pt/In2O3 catalysts for methanol synthesis from CO2 hydrogenation. J. Catal. 2021, 394, 236–244. [Google Scholar] [CrossRef]
  70. Li, H.; Wang, L.; Dai, Y.; Pu, Z.; Lao, Z.; Chen, Y.; Wang, M.; Zheng, X.; Zhu, J.; Zhang, W.; et al. Synergetic interaction between neighbouring platinum monomers in CO2 hydrogenation. Nat. Nanotechnol. 2018, 13, 411–417. [Google Scholar] [CrossRef]
  71. 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]
  72. 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]
  73. Zhang, L.; Liu, X.; Wang, H.; Cao, L.; Huang, C.; Li, S.; Zhang, X.; Guan, Q.; Shao, X.; Lu, J. Size-dependent strong metal-support interaction in Pd/ZnO catalysts for hydrogenation of CO2 to methanol. Catal. Sci. Technol. 2021, 11, 4398–4405. [Google Scholar] [CrossRef]
  74. Tian, G.; Wu, Y.; Wu, S.; Huang, S.; Gao, J. CO2 hydrogenation to methanol over Pd/MnO/In2O3 catalyst. J. Environ. Chem. Eng. 2022, 10, 106965. [Google Scholar] [CrossRef]
  75. Jiang, F.; Jiang, F.; Wang, S.; Xu, Y.; Liu, B.; Liu, X. Catalytic Activity for CO2 Hydrogenation is Linearly Dependent on Generated Oxygen Vacancies over CeO2-Supported Pd Catalysts. ChemCatChem 2022, 14, e202200422. [Google Scholar] [CrossRef]
  76. Caparros, F.J.; Soler, L.; Rossell, M.D.; Angurell, I.; Piccolo, L.; Rossell, O.; Llorca, J. Remarkable Carbon Dioxide Hydrogenation to Ethanol on a Palladium/Iron Oxide Single-Atom Catalyst. ChemCatChem 2018, 10, 2365–2369. [Google Scholar] [CrossRef]
  77. Lou, Y.; Jiang, F.; Zhu, W.; Wang, L.; Yao, T.; Wang, S.; Yang, B.; Yang, B.; Zhu, Y.; Liu, X. CeO2 supported Pd dimers boosting CO2 hydrogenation to ethanol. Appl. Catal. B-Environ. 2021, 291, 120122. [Google Scholar] [CrossRef]
  78. Zhu, C.; Xu, K.; Fang, Y.; Lv, Q.; Jiang, K.; Zhao, H.; Chen, Y.; Wang, P.; Yang, H.; Wu, L.; et al. Synergetic effect of Ce/Zr for ethanol synthesis from syngas over Rh-based catalyst. Fuel 2023, 334, 126770. [Google Scholar] [CrossRef]
  79. Wang, J.; Sun, K.; Jia, X.; Liu, C. CO2 hydrogenation to methanol over Rh/In2O3 catalyst. Catal. Today 2021, 365, 341–347. [Google Scholar] [CrossRef]
  80. Lu, Z.; Wang, J.; Sun, K.; Xiong, S.; Zhang, Z.; Liu, C. CO2 hydrogenation to methanol over Rh/In2O3–ZrO2 catalyst with improved activity. GreenChE 2022, 3, 165–170. [Google Scholar] [CrossRef]
  81. Cao, F.; Gong, N.; Ma, Z.; Wang, X.; Tan, M.; Wu, Y.; Tan, Y. Controlling CO2 hydrogenation selectivity by Rh-based catalysts with different crystalline phases of TiO2. Chem. Commun. 2022, 58, 4219–4222. [Google Scholar] [CrossRef] [PubMed]
  82. Zheng, K.; Li, Y.; Liu, B.; Jiang, F.; Xu, Y.; Liu, X. Ti-doped CeO2 Stabilized Single-Atom Rhodium Catalyst for Selective and Stable CO2 Hydrogenation to Ethanol. Angew. Chem. Int. Edit. 2022, 61, e202210991. [Google Scholar] [CrossRef] [PubMed]
  83. Wang, G.; Luo, R.; Yang, C.; Song, J.; Xiong, C.; Tian, H.; Zhao, Z.-J.; Mu, R.; Gong, J. Active sites in CO2 hydrogenation over confined VOx-Rh catalysts. Sci. China-Chem. 2019, 62, 1710–1719. [Google Scholar] [CrossRef]
  84. Zhang, F.; Zhou, W.; Xiong, X.; Wang, Y.; Cheng, K.; Kang, J.; Zhang, Q.; Wang, Y. Selective Hydrogenation of CO2 to Ethanol over Sodium-Modified Rhodium Nanoparticles Embedded in Zeolite Silicalite-1. J. Phys. Chem. C 2021, 125, 24429–24439. [Google Scholar] [CrossRef]
  85. Kusama, H.; Okabe, K.; Sayama, K.; Arakawa, H. CO2 hydrogenation to ethanol over promoted Rh/SiO2 catalysts. Catal. Today 1996, 28, 261–266. [Google Scholar] [CrossRef]
  86. Wu, Q.; Shen, C.; Rui, N.; Sun, K.; Liu, C. Experimental and theoretical studies of CO2 hydrogenation to methanol on Ru/In2O3. J. CO2 Util. 2021, 53, 101720. [Google Scholar] [CrossRef]
  87. Xiong, S.; Lu, Z.; Shen, C.; Liu, C. ZrO2 promoted Ru/In2O3 catalyst for selective hydrogenation of CO2 to methanol. Chem. Eng. Sci. 2023, 282, 119246. [Google Scholar] [CrossRef]
  88. Shen, C.; Sun, K.; Zhang, Z.; Rui, N.; Jia, X.; Mei, D.; Liu, C. Highly Active Ir/In2O3 Catalysts for Selective Hydrogenation of CO2 to Methanol: Experimental and Theoretical Studiese. ACS Catal. 2021, 11, 4036–4046. [Google Scholar] [CrossRef]
  89. Zhang, X.; Kirilin, A.V.; Rozeveld, S.; Kang, J.H.; Pollefeyt, G.; Yancey, D.F.; Chojecki, A.; Vanchura, B.; Blum, M. Support Effect and Surface Reconstruction in In2O3/m-ZrO2 Catalyzed CO2 Hydrogenation. ACS Catal. 2022, 12, 3868–3880. [Google Scholar] [CrossRef]
  90. Shi, Y.; Su, W.; Kong, L.; Wang, J.; Lv, P.; Hao, J.; Gao, X.; Yu, G. The homojunction formed by h-In2O3(110) and c-In2O3(440) promotes carbon dioxide hydrogenation to methanol on graphene oxide modified In2O3. J. Colloid Interface Sci. 2022, 623, 1048–1062. [Google Scholar] [CrossRef]
  91. Fang, T.; Liu, B.; Lian, Y.; Zhang, Z. Selective Methanol Synthesis from CO2 Hydrogenation over an In2O3/Co/C-N Catalyst. Ind. Eng. Chem. Res. 2020, 59, 19162–19167. [Google Scholar] [CrossRef]
  92. Tian, G.; Wu, Y.; Wu, S.; Huang, S.; Gao, J. Solid-State Synthesis of Pd/In2O3 Catalysts for CO2 Hydrogenation to Methanol. Catal. Lett. 2022, 153, 903–910. [Google Scholar] [CrossRef]
  93. Rui, N.; Zhang, F.; Sun, K.; Liu, Z.; Xu, W.; Stavitski, E.; Senanayake, S.D.; Rodriguez, J.A.; Liu, C.-J. Hydrogenation of CO2 to Methanol on a Auδ+−In2O3−x Catalyst. ACS Catal. 2020, 10, 11307–11317. [Google Scholar] [CrossRef]
  94. Ye, X.; Yang, C.; Pan, X.; Ma, J.; Zhang, Y.; Ren, Y.; Liu, X.; Li, L.; Huang, Y. Highly Selective Hydrogenation of CO2 to Ethanol via Designed Bifunctional Ir1-In2O3 Single-Atom Catalyst. J. Am. Chem. Soc. 2020, 142, 19001–19005. [Google Scholar] [CrossRef] [PubMed]
  95. Witoon, T.; Numpilai, T.; Nijpanich, S.; Chanlek, N.; Kidkhunthod, P.; Cheng, C.K.; Ng, K.H.; Vo, D.-V.N.; Ittisanronnachai, S.; Wattanakit, C.; et al. Enhanced CO2 hydrogenation to higher alcohols over K-Co promoted In2O3 catalysts. Chem. Eng. J. 2022, 431, 133211. [Google Scholar] [CrossRef]
  96. Tada, S.; Ochiai, N.; Kinoshita, H.; Yoshida, M.; Shimada, N.; Joutsuka, T.; Nishijima, M.; Honma, T.; Yamauchi, N.; Kobayashi, Y.; et al. Active Sites on ZnxZr1−xO2−x Solid Solution Catalysts for CO2-to-Methanol Hydrogenation. ACS Catal. 2022, 12, 7748–7759. [Google Scholar] [CrossRef]
  97. Sha, F.; Tang, C.; Tang, S.; Wang, Q.; Han, Z.; Wang, J.; Li, C. The promoting role of Ga in ZnZrOxsolid solution catalyst for CO2 hydrogenation to methanol. J. Catal. 2021, 404, 383–392. [Google Scholar] [CrossRef]
  98. Lee, K.; Anjum, U.; Araújo, T.P.; Mondelli, C.; He, Q.; Furukawa, S.; Pérez-Ramírez, J.; Kozlov, S.M.; Yan, N. Atomic Pd-promoted ZnZrOxsolid solution catalyst for CO2 hydrogenation to methanol. Appl. Catal. B-Environ. 2022, 304, 120994. [Google Scholar] [CrossRef]
Figure 1. Various reaction pathways of CO2 hydrogenation to alcohols. A, HCOO* pathway; B, r-HCOO* pathway; C, RWGS + CO-Hydro pathway; D, trans-COOH* pathway; E, CO mediated pathway; F, formate mediated pathway; G, methanol mediated pathway. Blue line, C–H bond break or formation; brown line, C–O bond break or formation; green line, C–C bond coupling; orange line, O-H bond break or formation.
Figure 1. Various reaction pathways of CO2 hydrogenation to alcohols. A, HCOO* pathway; B, r-HCOO* pathway; C, RWGS + CO-Hydro pathway; D, trans-COOH* pathway; E, CO mediated pathway; F, formate mediated pathway; G, methanol mediated pathway. Blue line, C–H bond break or formation; brown line, C–O bond break or formation; green line, C–C bond coupling; orange line, O-H bond break or formation.
Catalysts 14 00232 g001
Figure 2. (a) In situ DRIFTS spectra of the CO2 hydrogenation over the Cu-ZrO2 catalyst at different times [26]. (b) Schematic illustration of CO2 hydrogenation pathways on c-In2O3 and h-In2O3 surfaces, respectively. The black line shows methanol formation. The red line shows CO formation [27].
Figure 2. (a) In situ DRIFTS spectra of the CO2 hydrogenation over the Cu-ZrO2 catalyst at different times [26]. (b) Schematic illustration of CO2 hydrogenation pathways on c-In2O3 and h-In2O3 surfaces, respectively. The black line shows methanol formation. The red line shows CO formation [27].
Catalysts 14 00232 g002
Figure 3. Reaction pathway of CO2 hydrogenation to methanol on the Pd/In2O3 catalyst [31].
Figure 3. Reaction pathway of CO2 hydrogenation to methanol on the Pd/In2O3 catalyst [31].
Catalysts 14 00232 g003
Figure 4. Potential energy diagram for CO2 hydrogenation to CH3OH on (a) Ti3O6H6/Cu(111) and (b) Zr3O6H6/Cu(111). “TS” corresponds to the transition state [34].
Figure 4. Potential energy diagram for CO2 hydrogenation to CH3OH on (a) Ti3O6H6/Cu(111) and (b) Zr3O6H6/Cu(111). “TS” corresponds to the transition state [34].
Catalysts 14 00232 g004
Figure 5. The pathway of CO2 hydrogenation to higher alcohol on the Cu-Fe-Zn catalyst [39].
Figure 5. The pathway of CO2 hydrogenation to higher alcohol on the Cu-Fe-Zn catalyst [39].
Catalysts 14 00232 g005
Figure 6. In situ Fourier Transform infrared spectroscopy of CoAlOx-600 in CO2 hydrogenation to alcohols [40].
Figure 6. In situ Fourier Transform infrared spectroscopy of CoAlOx-600 in CO2 hydrogenation to alcohols [40].
Catalysts 14 00232 g006
Figure 7. (a) Schematic of CO2 hydrogenation over the Rh-based catalyst with or without hydroxyl groups on TiO2. (b) In situ DRIFTS of RhFeLi/TiO2 catalyst [43].
Figure 7. (a) Schematic of CO2 hydrogenation over the Rh-based catalyst with or without hydroxyl groups on TiO2. (b) In situ DRIFTS of RhFeLi/TiO2 catalyst [43].
Catalysts 14 00232 g007
Figure 8. Schematic illustration of CO2 hydrogenation to methanol over Cu/SiO2 [47].
Figure 8. Schematic illustration of CO2 hydrogenation to methanol over Cu/SiO2 [47].
Catalysts 14 00232 g008
Figure 9. Bimetallic Cu centers supported on Zr12-SBUs in a MOF for producing C2 + oxygenates [56].
Figure 9. Bimetallic Cu centers supported on Zr12-SBUs in a MOF for producing C2 + oxygenates [56].
Catalysts 14 00232 g009
Figure 10. (a) Schematic illustration over Co@Six catalyst. (b) Dependences of FTIR signal intensity of HCOO* and CH3O* species on time during the CO2 hydrogenation over Co@Si0.95 catalyst [59].
Figure 10. (a) Schematic illustration over Co@Six catalyst. (b) Dependences of FTIR signal intensity of HCOO* and CH3O* species on time during the CO2 hydrogenation over Co@Si0.95 catalyst [59].
Catalysts 14 00232 g010
Figure 11. Schematic illustration of CO2 hydrogenation to ethanol reaction mechanism on CoAlOx catalyst [63].
Figure 11. Schematic illustration of CO2 hydrogenation to ethanol reaction mechanism on CoAlOx catalyst [63].
Catalysts 14 00232 g011
Figure 13. Schematic illustration of catalytic CO2 hydrogenation mechanism over Pd/CeO2 catalyst [75].
Figure 13. Schematic illustration of catalytic CO2 hydrogenation mechanism over Pd/CeO2 catalyst [75].
Catalysts 14 00232 g013
Figure 14. (a) Schematic illustration of CO2 hydrogenation to methanol over different Rh/TiO2 catalysts (b). Reaction results over Rh-based catalysts with different TiO2 supports [81].
Figure 14. (a) Schematic illustration of CO2 hydrogenation to methanol over different Rh/TiO2 catalysts (b). Reaction results over Rh-based catalysts with different TiO2 supports [81].
Catalysts 14 00232 g014
Figure 15. Schematic diagram of CO2 hydrogenation with (a) VOx-promoted Rh/MCM-41 catalyst [83] and (b) Na-promoted Rh@S-1 catalyst. [84].
Figure 15. Schematic diagram of CO2 hydrogenation with (a) VOx-promoted Rh/MCM-41 catalyst [83] and (b) Na-promoted Rh@S-1 catalyst. [84].
Catalysts 14 00232 g015
Figure 17. Schematic illustration of the preparation processes of the Ir1-In2O3 catalyst [94].
Figure 17. Schematic illustration of the preparation processes of the Ir1-In2O3 catalyst [94].
Catalysts 14 00232 g017
Table 1. Possible reactions involved in CO2 hydrogenation to alcohols.
Table 1. Possible reactions involved in CO2 hydrogenation to alcohols.
EntryReactionReaction Equation∆G298K
(kJ/mol)
∆H298K (kJ/mol)
1CO2 hydrogenation to methanolCO2(g) + 2H2(g) ⇋ CH3OH(g) + H2O(g)3.5−49.3
2Reverse water–gas shift reactionCO2(g) + H2(g) ⇋ CO(g) + H2O(g)28.641.2
3CO hydrogenation to methanolCO(g) + 2H2(g) ⇋ CH3OH(g)/−90.6
4CO2 hydrogenation to alkanenCO2 + (3n + 1)H2 ⇋ CnH2n+2 + 2nH2O//
5CO2 hydrogenation to olefinnCO2 + 3nH2 ⇋ CnH2n + 2nH2O//
6CO2 hydrogenation to higher alcoholsnCO2 + 3nH2 ⇋ CnH2n+1OH + (2n − 1)H2O//
7CO2 hydrogenation to ethanol2CO2 + 6H2 ⇋ C2H5OH(g) + 3H2O(g)−65.62−173.74
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

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. https://doi.org/10.3390/catal14040232

AMA Style

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(4):232. https://doi.org/10.3390/catal14040232

Chicago/Turabian Style

Cui, Guoqing, Yingjie Lou, Mingxia Zhou, Yuming Li, Guiyuan Jiang, and Chunming Xu. 2024. "Review of Mechanism Investigations and Catalyst Developments for CO2 Hydrogenation to Alcohols" Catalysts 14, no. 4: 232. https://doi.org/10.3390/catal14040232

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