Next Article in Journal
PWV Inversion Model Based on Random Forest and the Trend of Its Conversion Rate with Precipitation in Hubei from 1960 to 2020
Previous Article in Journal
Centennial Variation and Mechanism of the Extreme High Temperatures in Summer over China during the Holocene Forced by Total Solar Irradiance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hydrogenation of Carbon Dioxide to Methanol over Non-Noble Catalysts: A State-of-the-Art Review

1
State Key Laboratory of Materials-Oriented Chemical Engineering, College of Chemical Engineering, Nanjing Tech University, Nanjing 211816, China
2
School of Energy Science and Engineering, Nanjing Tech University, Nanjing 211816, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Atmosphere 2023, 14(8), 1208; https://doi.org/10.3390/atmos14081208
Submission received: 12 June 2023 / Revised: 6 July 2023 / Accepted: 24 July 2023 / Published: 27 July 2023
(This article belongs to the Section Air Pollution Control)

Abstract

:
The malignant environmental changes caused by the ever-increasing amount of anthropogenic CO2 emissions have been particularly prominent in recent years. To achieve carbon mitigation and carbon neutrality, CO2 hydrogenation to methanol is regarded as a promising and sustainable route. However, the development of catalysts with exceptional performance and the establishment of a clear structure–activity relationship remain formidable challenges. Considering the lack of a state-of-the-art review on the catalytic progress of CO2 hydrogenation to methanol over non-noble catalysts, we conducted a detailed review in terms of the thermodynamic analysis, catalytic development, and reaction mechanism. In this work, we mainly reviewed the latest research progress of different catalysts including Cu-based, In2O3-based, bimetallic, solid solution, and other catalysts. Meanwhile, we summarized the effects of the support materials, promoters, and preparation methods on the catalytic performance. In addition, we also summarized the possible reaction mechanisms of direct hydrogenation of CO2 to methanol. Overall, this work would be of importance for the researchers to obtain a comprehensive understanding of the design and development of efficient catalysts for CO2 hydrogenation to methanol.

1. Introduction

Fossil fuels such as coal, oil, and natural gas have been consumed on an unprecedented scale to meet the increasing energy demands, which have caused massive CO2 emissions in the environment. To date, global anthropogenic CO2 emissions have reached 37 Gton CO2 per year, which can cause a series of environmental problems, such as serious greenhouse effects, ocean acidification, and glacier melting [1,2,3]. Therefore, the control and mitigation of CO2 emissions in the atmosphere has become an urgent task to protect the environment. Carbon capture, utilization, and storage (CCUS) strategies have been proposed to reduce CO2 emissions, which mainly can be divided into two categories: carbon capture and storage (CCS) and carbon capture and utilization (CCU) [4,5]. Compared with CCS technology, CCU technology can convert the waste CO2 into various value-added liquid fuels and platform chemicals, such as methanol, carbon monoxide, methane, dimethyl ether, and polycarbon, and is considered as a potential and sustainable pathway to achieve carbon neutrality [6,7,8]. Among various products, methanol is an important platform compound, which can be further converted into low-carbon olefins, aromatics, gasoline, and other high value-added chemicals and fuels [9,10,11]. Therefore, CO2 hydrogenation to methanol using green hydrogen has received wide attention due to its great economic value and industrial application prospects.
For CO2 hydrogenation to methanol, methanol synthesis reaction and reverse water-gas-shift (RWGS) reaction are two main competitive reactions. CO2 is firstly converted to the intermediate CO, which is hydrogenated to form *H3CO and further hydrogenated to methanol [12,13,14]. From the perspective of thermodynamics, the methanol synthesis reaction is exothermic, and the volume of the molecules is reduced, so high-pressure and low-temperature reaction conditions are conducive [2,15]. At present, thermocataltyic, photocatalytic, and electrocatalytic technologies have been successfully applied in the CO2 hydrogenation to methanol. However, the thermocatalytic technology has the most promising prospect for industrialization due to higher CO2 conversion and fewer by-products [16]. Up to now, remarkable progress has been made for the thermocatalytic conversion of CO2 to methanol in terms of catalyst development and mechanism study. Generally, the catalysts for the hydrogenation of CO2 to methanol can be roughly divided into Cu-based catalysts, noble metal-based catalysts, and mixed oxide catalysts [17,18,19,20]. Cu-based catalysts are the most widely studied catalysts, but they are prone to deactivation under reaction conditions attributed to the Ostwald ripening effect and particle migration, which seriously reduces the methanol yield [21,22,23,24]. Furthermore, noble metal-based catalysts exhibit superior stability and resilience against sintering and poisoning, making them a viable alternative to Cu-based catalysts. However, the methanol selectivity of these catalysts is relatively low, their costs are relatively high, and the product distribution is difficult to regulate due to their weak binding affinity with CO2 [25,26,27]. In recent years, indium oxide (In2O3) has been developed for the hydrogenation of CO2 to methanol, and the oxygen vacancy on its surface plays an important role in the high selectivity and activity of methanol. However, the CO2 conversion of the indium oxide-based catalyst is low. Therefore, it is of importance to add additives to enhance the ability to activate H2 and generate oxygen vacancy [28,29,30]. In addition, some bimetallic catalysts and other novel catalysts have also been used in the hydrogenation of CO2 to methanol [31,32,33].
This work is mainly divided into four sections: (1) the challenges in thermodynamics are proposed firstly. (2) The progress in conventional Cu-based catalysts with different supports, promoters, and preparation methods is deeply discussed. Additionally, an overview of the advancements of other non-noble catalysts, such as In2O3-based catalysts, bimetallic catalysts, and solid solution catalysts, is also provided. (3) The possible reaction mechanisms for the hydrogenation of CO2 to methanol are summarized in detail. (4) The summary and prospects are proposed to provide comprehensive guidelines for the design and development of efficient catalysts.

2. Thermodynamic Analysis

Large-scale commercial methanol production is mainly derived from syngas. Compared with CO as a raw material (Equation (1)), the conversion and activation of CO2 generally require numerous energy due to its highly stable molecule and low Gibbs free energy (Equation (2)) [34]. Therefore, high temperature is beneficial for promoting CO2 activation and accelerating the reaction rate. However, it should be noted that different from CO hydrogenation, the reverse water gas shift (RWGS) reaction (Equation (3)) generally competes with CO2 hydrogenation reactions, where the RWGS reaction is endothermic, whereas the CO2 hydrogenation reaction is exothermic [35,36]. Therefore, as shown in Equation (2), the operating conditions of low temperatures and high pressures are favorable. In addition, unlike CO hydrogenation, CO2 hydrogenation generally consumes more H2 and generates more water due to higher oxygen content. The presence of water accelerates the sintering of the active site in the catalyst, resulting in catalyst deactivation and the reduction of methanol production in subsequent steps [37,38,39]. Therefore, it is of importance to design an efficient catalytic system to promote CO2 activation and suppress the poisonousness of the by-products for the catalysts at suitable temperatures and pressures [40]. The CO hydrogenation to methanol reaction is as follows:
CO + 2H2 = CH3OH ∆H298k = −90.6 kJ mol−1,
The CO2 hydrogenation to methanol reaction is as follows:
CO2 + 3H2 = CH3OH + H2O ∆H298k = −49.5 kJ mol−1
The RWGS reaction is as follows:
CO2 + H2 = CO + H2O ∆H298k = 41.2 kJ mol−1

3. Catalyst Development of CO2 Hydrogenation to Methanol

3.1. Cu-Based Catalyst

In the 1940s, it was possible to synthesize methanol with CO2 as the only raw feedstock, and the relevant research report on hydrogenation of CO2 to methanol first appeared. In the 1960s, the Cu-ZnO type catalyst used for methanol industrial production was reported by Imperial Chemical Industries [41,42]. Since then, modifying Cu-ZnO catalysts by introducing different supports and promoters and optimizing the preparation methods has received more attention and is extensively studied. The catalytic performances of some of the latest Cu-based catalysts are summarized in Table 1.

3.1.1. Supports

The type of support has a great influence on the physical properties of the catalyst, and the activity of the catalyst for CO2 hydrogenation to methanol is proportional to the surface area of Cu. In Cu-based catalysts, the support with high specific surface area can disperse the active components well and prevent the catalyst from deactivation due to sintering [35]. The traditional supports primarily include Al2O3 [56,57,58], ZnO [59,60], and ZrO2 [61,62,63]. However, the water generated by the side reaction can accelerate the sintering of the active sites, which leads to the catalyst being relatively easy to deactivate, and Al2O3 has hydrophilic properties, so the development of other carriers to make up for this is necessary.
Some emerging supports have been gradually successfully developed, such as mesoporous aluminosilicate support, Mg-Al layered double hydroxide (LDH) [43,45]. For example, Kubovics et al. [46] synthesized a novel Cu-ZnO multicomponent catalyst as a 3D aerogel, using reduced graphene oxide (rGO) as a support. It was found that the addition of rGO to the catalytic system significantly increased the methanol production rate by fourfold compared to pristine Cu-ZnO NPs at 220 °C. This was mainly because the intrinsic activity of Cu was enhanced due to the strain imposed at the ZnO interphase, causing a misfit at the contact surface. Recently, MOF and zeolite have been potential supports due to their abundant pore structure and tremendous surface area. Generally, MOF and zeolite are beneficial for confining the growth of Cu particles, obtaining high Cu dispersion, maximizing interface sites, and strengthening their interaction with Cu by regulating electron transfer [64]. For example, Duma et al. [47] prepared Cu-ZnO catalysts supported on an aluminum fumarate metal-organic framework (AlFum MOF) with high surface area and good porosity, which further promoted the homogenous, uniform dispersion of Cu and Zn active sites. As a result, the catalysts exhibited good activity, with a doubling loading of Cu and Zn over the AlFum MOF. Therefore, the CO2 conversion increases from 10.8 to 45.6%, and the methanol production increases from 0.034 to 0.056 gMeOH·gcat−1·h−1. SiO2 is widely used as a support for various heterogeneous catalysts due to its low price and high specific surface area. For instance, Shawabkeh et al. [65] studied the CO2 adsorption on SiO2 at different Cu loading to investigate the CO2 interaction with the surface (mainly oxygen atom) via the sol–gel method from both experimental and theoretical points of view. The results suggested that an increase in the amount of CuO on the surface of SiO2 improved the basicity of the adsorbent, resulting in more CO2 uptake. Dalia Santa Cruz-Navarro et al. [66] synthesized Cu-based catalysts with ammonium salt and acidic ZSM-5 zeolite as support using the liquid phase (LPIE) and solid state (SSIE) ion exchange methods and compared their catalytic performances. The results showed that the catalyst prepared using the SSIE method had higher copper loading than the catalyst prepared using the LPIE method, which was attributed to the SSIE method promoting the diffusion of the volatile copper metal complexes through the internal channels of molecular sieves.

3.1.2. Promoters

The addition of promoters can change the acidity and alkalinity of the catalyst surface, enhance the interaction between the active components, form more defect sites on the surface, and change the product composition distribution. Cu-ZnO catalysts with various promoters have received wide attention. ZrO2 and Ga2O3 are considered important catalysts for the modification of Cu–ZnO [67]. For example, Sun et al. [48] designed Cu-ZnO catalysts with different molar ratios of ZrO2 using the co-current precipitation method and controlled the pH to regulate the distributions of Cu0 and Cu+ species. It was found that the CuZn10Zr (10 mol% ZrO2) catalyst had the highest space-time yield to methanol with 0.65 gMeOH·gcat−1·h−1 under the reaction conditions of 220 °C and 3 MPa. Their research showed that the addition of an appropriate amount of ZrO2 was beneficial to promoting the dispersion of Cu to provide more active sites for H2 and CO2 adsorption and activation. Cored et al. [68] studied two CuO/ZnO/Ga2O3 catalysts prepared using the co-precipitation method and compared the promoting effect of Ga3+-doped in the wurtzite ZnO lattice of a Cu/ZnO/Ga2O3 catalyst with that of a zinc gallate (ZnGa2O4) phase. The study indicated that Ga3+-doped ZnO has been considered as a more efficient promoter than ZnGa2O4 owing to the presence of surface vacancies with loosely bounded electrons, increasing the conductivity of the material and enhancing methanol selectivity versus CO formation. Among the different interfaces between the samples, the interaction of Cu and Ga is more favorable in the Ga3+-doped ZnO sample by increasing the number of surface basic centers, which is important for the stabilization of intermediate species. Sang et al. [69] systematically studied activation mechanisms of CO2 on the Ga-modified Cu surface (Figure 1) with different forms (Cu, Cu8Ga1, Cu6Ga3, and Ga2O3@Cu) using density functional theory (DFT) calculations combining the thermodynamics with chemical kinetics. DFT results and the Mulliken atomic charge of CO2 indicated that CO2 was transformed into chemisorbed CO2* with a lower reaction energy barrier, and two oxygen atoms of CO2 obtained different charges at the interface between the metal surface and Ga2O3, suggesting that the charge imbalance of the CO2 molecule was more favorable to the activation of C=O. For Cu, Cu8Ga1, and Cu6Ga3 catalysts, CO2 behaved as physical adsorption, and the oxygen atoms of CO2 gained the same charges.

3.1.3. Preparation Methods

It is well known that the preparation methods of catalysts have significant influences on the catalytic performances, which mainly include coprecipitation [22,23,49], impregnation [47,70,71], and sol–gel [65,70,72]. However, to further enhance the activity, selectivity, water tolerance, and stability of Cu-based catalysts, some emerging methods have appeared in recent reports [25,46,50,73,74]. For example, metal nanoparticles (NPs) attracted growing interest due to their outstanding applications in numerous fields, but so far, no large-scale synthesis of CuNPs has been reported using the disproportionation route and the reductive pathway. Ouyang et al. [75] report a method for obtaining Cu (0) nanoparticles (CuNPs) from readily available organocopper reagents (Figure 2). The method can be used to synthesize spherical CuNPs with excellent control of their size and shape on the decigram scale. Overall, the study offers many prospects for the rapidly developing field of copper plasmonic catalysis. Kubovics et al. [46] prepared its precursor CuOZnO NPs via conventional chemistry and designed rGO aerogels-supported CuZnO NPs with water repulsion as efficient 3D catalysts. Then, the supercritical CO2 was applied for the nano-structuration and the fabrication of 3D devices with macro- and mesoporosity, and the targeted sample was obtained through H2 reduction.
Santa et al. [66] studied the effect of the synthesis method on the physicochemical properties of Cu-based catalysts through two ion exchanges. The advantage of the ion exchange method in the liquid phase and the solid phase is that the compensating cation of the zeolite can be well exchanged with the metal ions with catalytic activity, so that the metal content in the pores reaches a high level and has good metal dispersion. The results show that both ion exchange methods are suitable for the incorporation of uniformly distributed Cu into the zeolite structure, and the solid ion exchange method has the highest exchange percentage. Ali et al. [76] prepared 30 wt% CuO 49.65 wt% ZnO2 0.35 wt% Al2O3 catalyst at glycine to nitrates (G/O) ratios between 0.1 and 1.23 through the solution combustion synthesis (SCS) method. In brief, in the experimental procedure, varying quantities of glycine were used as a fuel into a mixed nitrate precursor solution, and the generated mixture was subjected to continuous stirring. The mixture gradually transformed into a soft gel upon heating on a hotplate and underwent spontaneous combustion at 150 °C. Subsequently, the synthesized powder was subjected to calcination in a muffle furnace using static air at different temperatures. The heating and cooling rates during calcination were, respectively, set at +1 and −1 °C·min−1 for three hours. APrašnikar et al. [51] also prepared several perovskite-containing Cu/Sr/Ti materials using the one-step solution combustion method with different fuels and conducted a long-term stability test of the catalyst with the best catalytic performances (Cu5Sr3Ti2OX, citric acid, calcination at 650 °C). The results showed that the catalyst retained 91% of initial activity after 122 h and even preserved 71% of the activity. Han et al. [52] prepared hollow Cu@ZrO2 catalysts through the pyrolysis of Cu-loaded Zr-MOF and found that a lower pyrolysis temperature can enhance the porous structure and the metal–support interaction of Cu and ZrO2 (Figure 3). More specifically, low-temperature pyrolysis generated highly dispersed Cu nanoparticles with balanced Cu0/Cu+ sites, more surface basic sites, and abundant Cu-ZrO2 interface in the hollow structure, which contributed to enhancing the catalytic ability of CO2 adsorption/activation and selective hydrogenation to methanol. Qu et al. [77] anchored a small amount of Cu on the surface of a ZnAl2O4 support using the ammonia evaporation method. The interaction between Cu formed through ammonia evaporation and Zn-O structure derived from the surface of ZnAl2O4 was stronger, and it tended to form a higher proportion of Cu+, which stabilized the methoxy group and improved the methanol production efficiency.

3.2. In2O3-Based Catalyst

The traditional Cu-based catalysts have been extensively studied in CO2 hydrogenation to methanol [14,78]. However, its further application is limited by the high activity of the RWGS reaction and the poor stability and sintering of the active phase induced by H2O. Compared with the aforementioned Cu-based catalysts, In2O3 has moderate adsorption capacity of CO2 and CO and significantly better methanol selectivity than Cu, Co, and noble metal catalysts. Therefore, it has attracted extensive attention from researchers. In addition, In2O3 is easy to load and modify the surface, which can further promote the CO2 and H2 activation, and stabilize key intermediates, providing great potential for the design and preparation of efficient methanol synthesis catalysts [20,28,79].

3.2.1. Metal Promoters

Compared with traditional catalysts, In2O3 demonstrates better ability to adsorb and activate CO2 owing to its abundant oxygen vacancy on the surface. However, the weak H2 dissociation ability of In2O3 hinders the hydrogenation of carbon species, resulting in a significantly low CO2 conversion rate of approximately 6%. It is a common strategy to introduce other metal elements into the In2O3 system and form M/In2O3 structure to enhance H2 dissociation, adsorption, and overflow capacity.
As a metal with strong H2 dissociation ability, Pd is often used to further enhance the performance of In2O3-based catalysts. For example, Tian et al. [80] prepared Pd-In2O3 catalysts with high activity using the solid phase method. Pd species have high dispersibility on the prepared Pd/In2O3 catalyst, which significantly promoted H2 dissociation and provided sufficient H adatoms for CO2 hydrogenation. In addition, partial Pd2+ species can exist stably during CO2 hydrogenation, which facilitated methanol synthesis. Rui et al. [81] also prepared Pd/In2O3 catalysts by mixing the Pd/peptide composite and In2O3, followed by the removal of peptide through thermal treatment. In contrast, the catalyst was also synthesized using the conventional incipient wetness impregnation method. It is notable that the CO2 conversion of the former catalyst was higher than 20% at 300 °C, and the maximum STY was 0.89 gMeOH·gcat−1·h−1, which was higher than that of the catalyst prepared using the conventional method. This is mainly because the catalyst obtained using the former preparation method can obtain Pd-NPs with small particle size and high dispersion. The Pd NPs has a high hydrogen dissociation adsorption capacity, which can provide the hydrogen for the hydrogenation reaction and maintain the density of oxygen vacancies. Similar effects are also found for Pt, Rh, Au, and Ru supported on In2O3 [39]. For instance, Sun et al. [82] prepared Pt-In2O3 catalysts using the deposition precipitation method (DP). The synergistic effect between Pt nanoparticles and In2O3 effectively modulated the relationship between H2 activation and surface oxygen vacancy content of In2O3, promoting the efficient hydrogenation of CO2 to methanol and maintaining good stability. Rui et al. [83] prepared the Au/In2O3 catalyst using the sedimentation precipitation method. The catalyst exhibits excellent catalytic properties with 100% selectivity of the reaction below 225 °C, and the selectivity is still as high as 67.8% at 300 °C. This excellent performance is due to the Auδ+-In2O3-X interface used as the active site, which gives gold the ability to activate hydrogen by regulating the electronic structure of gold. In addition, Sun et al. [84] prepared a Ag/In2O3 catalyst in the same manner. They analyzed the interaction between the surface of Ag and In2O3 containing oxygen vacancy based on DFT. The results showed that the interface site could promote the activation and hydrogenation of CO2. Moreover, the introduction of Ag can promote the formation of surface oxygen vacancy, thus promoting the increase of oxygen vacancy sites and the adsorption and dissociation of CO2.
Recently, a study indicated that a Ni-promoted In2O3 catalyst was used to produce methanol via CO2 hydrogenation. Hensen et al. [85] prepared a Ni-promoted In2O3 catalyst using the one-step flame spray pyrolysis (FSP) method. The reduced Ni is conducive to the activation of H2 molecules, and NiO-In2O3 will form oxygen vacancies and Ni-O-In during the reaction. Specifically, the metal nitrate solution is dissolved in a mixed solution of ethanol and 2-ethylhexanoic acid at room temperature, and then, the resulting solution is injected into the nozzle of the flame synthesis device, and the catalyst is collected in a quartz filter. Frei et al. [86] synthesized the catalysts with different morphology of Ni using the coprecipitation and impregnation methods. Compared with the coprecipitation method, more stable and active catalysts were obtained using the dry impregnation method. The Ni content of different impregnating catalysts has an obvious influence on the selectivity of products. To be specific, the RWGS reaction promoted methanol synthesis to some extent, and no methane was produced when Ni content was less than 10 wt%. Considering that the special wettability of the Ni-promoted In2O3 catalyst is beneficial for forming layered structures rather than aggregate particles, their strong anchoring on oxides through alloying can obtain high catalyst stability. DFT simulation results showed that the In2O3-modulated Ni layer easily provided homogeneous cracking hydrogen to In2O3, enhancing the formation of oxygen vacancy and promoting the hydrogenation of CO2. However, it hardly can activate CO2 on its own, which overall explains the beneficial effects and the lack of methane generation. The catalyst consisting of 1 wt% Ni provided the best balance between charged and free radical hydrogen atoms, resulting in a twofold increase in the methanol space time yield (STY) compared to that achieved with pure In2O3. Wu et al. [87] reported the experimental and theoretical results of CO2 hydrogenation to methanol over the Ru-prompted In2O3 catalyst (Ru/In2O3). The results showed that the methanol selectivity of Ru/In2O3 catalyst with 1wt% Ru loading can reach 69.7%, and the STY can reach 0.57 gMeOH·gcat−1·h−1 at 300 °C and 5 MPa (Figure 4). Compared with In2O3, the Ru/In2O3 catalyst has better stability. The characterization analysis showed that Ru and In2O3 interacted with each other. Ru increased the content of oxygen vacancy on the catalyst surface, stabilized the surface structure of the catalyst, and inhibited the excessive reduction of In2O3. DFT calculation showed that the In2O3 surface makes Ru clusters more stable, and there is an obvious strong interaction between metal and support, which makes the surface structure of the catalyst more stable. Recently, Liu et al. [88] reported that Ir/In2O3-X catalysts prepared using the traditional impregnation method show high activity, selectivity, and stability for CO2 hydrogenation to methanol. It was found that there was a linear positive correlation between the amount of Ir and the catalytic activity within a certain range. The Ir/In2O3 catalyst with 10% Ir loading showed a methanol selectivity of 70% and methanol STY of 0.765 gMeOH·gcat−1·h−1 at 300 °C. The addition of Ir not only effectively avoids the excessive reduction of In2O3 but also stabilizes the oxygen vacancy on the surface of the In2O3 support, thus further improving the stability of the catalyst. In turn, In2O3 stabilizes the highly dispersed metal Ir, avoiding the Ir aggregation observed on other oxides.

3.2.2. Oxide Support

The composite oxide catalyst composed of In2O3 and other oxides can also promote the production of highly selective methanol. For example, Liu et al. [89] added ZrO2 into Pt/In2O3 catalyst to explore the effect of ZrO2 on catalytic performances of CO2 hydrogenation. The introduction of ZrO2 achieved high activity and stability of CO2 hydrogenation to methanol in the presence of high CO content. This is mainly because the addition of ZrO2 resulted in stronger electron transfer between Pt and In2O3-ZrO2 support, weakening the CO adsorption and inhibiting the overreduction of In2O3 and CO poisoning at the Pt site. In addition, it was found that the synergistic effect between the Zr-modified oxygen vacancy (In-OV-Zr) and the Pt site promoted the hydrogenation of CO2 to methanol through the formate route. Sharma et al. [19] investigated the catalytic performances of ZrO2 and CeO2-supported In2O3 catalysts for the synthesis of methanol from CO2 hydrogenation. The selectivity of In13/CeO2 for methanol is higher than that of In13/ZrO2 at 553 K, and the methanol selectivity decreases with the increase of temperature. However, it found that the ZrO2-supported In2O3 catalyst was very stable, while In2O3 supported by CeO2 was easy to be inactivated, as shown in Figure 5. This is mainly due to the hydrophilic properties [90], redox properties, and the agglomeration of CeO2 support further inhibiting the CO2 hydrogenation. In addition, it was found that the In13/CeO2 catalyst could be regenerated by washing the catalyst with Ar to a certain extent. However, repeated reaction–regeneration cycles showed that the conversion continued to decline after regeneration and was even lower in the following cycles. In summary, the adsorption of water for In13/CeO2 can recover some activity due to the partial reversibility, while it is irreversible for the structural change leading to sustaining deactivation. For comparison, In13/ZrO2 showed an excellent stability under the reaction conditions. Monoclinic ZrO2 as a support can greatly improve the activity of In2O3 for CO2 hydrogenation to methanol. Javier and Cecilia et al. [30] recently investigated electron, geometric, and interfacial phenomena, and the special role of Zr as a carrier in the catalytic hydrogenation of CO2 to methanol. According to the kinetic analysis, monoclinic ZrO2-based catalysts can activate reactants better than Al2O3 or CeO2 oxide-supported In2O3, possibly due to the superior nature of oxygen vacancy on the carrier In2O3 and the direct contribution of ZrO2 to CO2 activation through its oxygen vacancy.

3.3. Bimetallic Catalysts

Bimetals have unique electronic effects due to their unique chemical composition and geometric configuration, which can regulate the electronic, geometric, and chemical properties of their active components, thereby enhancing catalytic performance. Because of this, bimetallic catalysts have been widely used in CO2 hydrogenation to methanol in recent years, and catalysts with high catalytic activity and methanol selectivity can be obtained by adjusting the composition ratio and grain size of the two metal elements.
The metal Cu is widely used in the hydrogenation of CO2 to methanol, and Cu-containing bimetallic catalysts have been widely studied in recent years. For example, Xu et al. [91] compared the catalytic performances and surface properties of Cu-Pd bimetallic catalysts (Cu/Pd = 33.5) over different supports (SiO2, CeO2, TiO2, and ZrO2) and explored the effect of the reduction temperature on the catalytic performances. The results showed that the catalytic performances of SiO2 and ZrO2-loaded catalysts were more noticeably influenced by the reduction in temperature compared to the CeO2 and TiO2-loaded catalysts. This can be attributed to the varying degrees of interaction between Pd and Cu caused by the support and reduction in temperature. Lin et al. [92] prepared TiO2-CeO2 and TiO2-ZrO2 binary supports using the precipitation method in different ratios and then loaded Pd and Cu impregnated on the binary supports to form bimetallic catalysts. It was found that the production and selectivity of methanol were improved by the binary supports catalysts compared with the single support catalysts. This is probably because the surface area of the binary supports was significantly increased, which enhanced both the dispersion of the metal phase and gas adsorption. Meanwhile, the presence of more oxygen vacancies in the binary supports also promoted CO2 conversion. In addition, the binary carriers improved the CO2 adsorption behavior of weakly bonded species, which contributed to improving catalytic performances. Compared to the Ti-Ce binary support, the Ti-Zr binary support showed a greater improvement in the adsorption characteristics of weakly bonded CO2 and the moderately strong metal–support interaction (SMSI) effects. The methanol generation rate of Pd-Cu/Ti0.1Zr0.9O2 bimetallic catalyst reached 0.62 μmol·gcat−1·s−1. Han et al. [93] prepared Cu/ZnO catalysts with abundant Cu-ZnO interface using Cu-Zn bimetallic organic frameworks (MOFs) template strategy (X represents the Cu/Zn ratios), named (CuXZnO-MOF-74–350), which exhibited high methanol selectivity of 80% and long-term durability of 100 h under the condition of 190 °C and 4.0 MPa (Figure 6). The good catalytic performances were mainly due to the abundant Cu-ZnO interface, which facilitates the adsorption of CO2 and the formation of intermediates. Yu et al. [94] prepared Cu-Zn catalysts coated with UiO-66 using the deposition-precipitation method. The results showed that the Cu-Zn@UiO-66 catalysts with a Cu-Zn loading of 35 wt% and a Cu/Zn molar ratio of 2.5 had excellent catalytic activity. The catalyst had a maximum methanol yield of 9.1% at 240 °C and good reaction stability after a time on stream of 100 h due to the ultra-small nanoparticles (NP) confined in UiO-66 and the abundant Cu/ZnOX interface leading to the enhancement of synergistic effect of Cu and ZnO, which promotes the conversion of CO2 and the formation of methanol. In recent studies, Ni and In-based catalysts have also been used to form bimetallic oxide catalysts with Cu, both of which exhibit excellent catalytic performances. Wang et al. [95] studied the effect of Cu-Ni loading on different graphites (GO, rGO, NGO) for the synthesis of methanol from CO2 hydrogenation. The results showed that the CuNi-rGO catalyst showed 7.87% CO2 conversion and 98.7% methanol selectivity at 498 K and 4.0 MPa. This was attributed to the promotion of Cu2+ reduced by Ni and the strong chemisorption activation of CO2 caused by the CuNi-rGO catalyst. Shi et al. [96] prepared bimetallic catalysts with different Cu and In ratios via the co-precipitation method. When the molar ratio of Cu to In was 1:2, the catalysts exhibited the excellent performance due to the high dispersion and the improvement of synergistic effect of active sites. Under the condition of 260 °C, 3.0 MPa, and 7500 mL·gcat−1·h−1, the catalyst obtained CO2 conversion of 10.3%, CH3OH selectivity of 86.2%, and CH3OH STY of 0.190 gMeOH·gcat−1·h−1.
In bimetallic catalysts with Ni as the main active component, the electronic structure of the Ni atom is changed by the neighboring metal atoms, resulting in unique activity for methanol production. Studt et al. [97] synthesized and tested a series of catalysts with different Ni-Ga ratios (including Ni-rich sites and Ga-rich sites), and the results showed that Ni5Ga3 was particularly active and selective. Notably, they had superior methanol production due to their better ability to reduce the RWGS activity in favor of methanol production compared with the conventional Cu/ZnO/Al2O3 catalyst. Rasteiro et al. [98] prepared Ni5Ga3 catalysts with SiO2, ZrO2, and CeO2 as supports using the incipient wetness impregnation method. Among these three catalysts, the Ni5Ga3 catalyst supported by CeO2 was poisoned at the active site due to excessive adsorption of intermediates at the alloy–carrier interface, which resulted in the failure of intermediates to hydrogenate to methanol. The Ni5Ga3 catalyst supported by SiO2 had a poor affinity to adsorb CO2, which limited the methanol production through the alloy surface route. In contrast, the interface of the Ni5Ga3 catalyst supported by ZrO2 contributed to forming the stable intermediates and promoting hydrogen overflow, thus exhibiting the most excellent catalytic activity. This study illustrates the importance of alloy–support synergy. Otherwise, Duyar et al. [99] introduced a third metal (Au, Cu, Co) as a promoter in Ni-Ga/SiO2, which enhanced the activity of CO2 hydrogenation to methanol by weakening the interaction between the surface and the adsorbate with Au and Cu compared to Ni5Ga3/SiO2. The experimental results showed that the activity of Au-Ni-Ga increased by nearly four times, and Cu-Ni-Ga showed the highest specific activity among the three catalysts.
Pd is also often used as a bimetallic catalyst with other metals such as In, Zn, and Ga for CO2 hydrogenation to methanol. The formation of the bimetallic system facilitates the overflow of H2 to the active site, which reacts with the adsorbed CO2 and promotes methanol production. Therefore, the Pd-M (In, Zn, Ga) bimetallic system has great potential for CO2 hydrogenation to methanol. As seen in the work of Ojelade’s team [100], they conducted a comparative study of PdZn/CeO2 catalysts with different Pd/Zn composition ratios for methanol synthesis. When the Pd/Zn molar ratio was close to 1, the PdZn alloy phase on the CeO2 support was maximum, and the methanol STY maximum was 0.114 gMeOH·gcat−1·h−1. Yin et al. [101] studied PdZn alloy catalysts prepared from Pd@ZIF-8 precursors, which exhibited excellent methanol STY up to 0.65 gMeOH·gcat−1·h−1 and TOF of 972 h−1. This may be attributed to the SMSI effects between Pd and ZnO support, which are facilitated by the sub-nano-Pd being confined within the ZIF-8 pore structure. Snider et al. [102] synthesized InPd/SiO2 catalysts with different In/Pd ratios using the incipient wetness impregnation method. The catalysts showed the highest activity at 4 MPa and 300 °C when the molar ratio of In to Pd was 2:1. More specifically, the methanol selectivity and yield were 61% and 0.588 gMeOH·gcat−1·h−1, respectively. Experimental results and DFT calculations indicated that the synergistic interaction between bimetallic In-Pd particles enriched with In on the surface of In2O3 enhances the reaction activity. Tian et al. [103] used TCPP(Pd)@MIL-68(In) as a sacrificial template to synthesize the loaded Pd@In2O3 catalysts, and the obtained catalysts showed a high 81.1 gMeOH·gPd−1·h−1 within 50 h STY and 81% methanol selectivity at 295 °C and 3.0 MPa. DFT calculations showed that the InPd clusters over-reduced In2O3, which was detrimental to the improvement of the catalytic activity for methanol production by regulating the electronic interactions between Pd and In2O3. Collins et al. [104] prepared GaPd bimetallic catalysts by loading different ratios of Ga and Pd atoms (2,4,8 atom/atom) on SiO2. The results showed that the turnover frequency to methanol based on surface palladium was 200-folds higher than that of the monometallic Pd/SiO2 catalyst, because the apparent activation energy of methanol synthesis on the surface palladium is much lower than that on the monometallic Pd/SiO2 catalyst. The characterization results showed that Pd2Ga bimetallic particles play a significant role in dissociating H2 and providing hydrogen atoms to Ga2O3 in the reaction. Close contact between Pd2Ga and Ga2O3 may be the key to optimizing the performances of the bifunctional mechanism. A new Rh-In bimetallic catalyst was first reported by Li et al. [105]. They prepared a series of Rh-containing catalysts with different In/Al compositions using the wet-impregnation method. The result indicated that the optimal ratio of In to Al was 1, and the methanol yield was 0.001 gMeOH·gcat−1·h−1. The catalysts were characterized by their ability to maintain high H2 conversion in H2-deficient feed gas and inhibit the RWGS reaction. Geng et al. [106] prepared SiO2-loaded In-Ru bimetallic catalysts and investigated the promotion of In on Ru in CO2 hydrogenation to methanol. The catalyst achieved 85% methanol selectivity without methane at 240 °C and 3.4 MPa. It enhanced the charge transfer from the surface to the formic acid state, stabilized the formic acid and methoxy on the surface, and inhibited H2 activation and the conversion of CO hydrogenation to CH4 and the violent decomposition of methanol on Ru to CO.

3.4. Solid Solution Catalysts

In addition to the catalysts mentioned above, solid solution catalysts have a broad application prospect for the hydrogenation of CO2 to methanol [10]. For example, Li et al. [107] prepared a series of Ga-promoted ZnZrOx solid solution catalysts to improve the methanol activity and keep high methanol selectivity. They found that when Ga content varied from 0 to 10% (the mole fraction of Ga in the total metal molar content), the CO2 conversion fluctuated between 7.7 and 8.8%, while the methanol selectivity remained unchanged (87.5% vs. 87.4%) at 320 °C. The introduction of Ga in the ZnZrOX solid solution catalyst improved the adsorption and activation ability of the catalyst for H2 and CO2 and promoted the hydrogenation of HCOO* to CH3O* (Figure 7). Li’s team also prepared ZnO-ZrO2 solid solution catalysts with ordered mesoporous structure using the evaporation-induced self-assembly (EISA) method and compared them with the catalysts prepared using the co-precipitation method [108]. The results showed that the methanol producing rate of 20% ZnO-ZrO2 catalyst synthesized using the EISA method was 0.707 gMeOH·gcat−1·h−1, which was 1.35 times that of the co-precipitation catalyst at 320 °C and 5.5 MPa. This is mainly because the EISA catalyst has a larger specific surface area and more active sites for adsorbing CO2 and H2. The preparation method mainly dissolves the triblock copolymer P123 in ethanol and adds n-butanol zirconium and ZnCl2. After stirring overnight, the solvent is evaporated to form a gel and then calcined. The surface of the calcined catalyst needs to be washed with water and ethanol, respectively, and finally dried.
Furthermore, Tada et al. [109] explored the effect of Zn content on the hydrogenation of CO2 to methanol over ZnXZr1-XO2-X catalyst and ensured the structure of active sites through theoretical calculation and experiment. When Zn content was low, the ZnXZr1-XO2-X catalyst was mainly composed of Zn clusters (i.e., isolated [ZnOa] clusters and [ZnbOc] oligomers), which meant the formation of Zn-O-Zr sites with specific activity for CO2 hydrogenation to methanol. However, the excessive addition of Zn can result in the aggregation of ZnO clusters and the domain formation of ZrO2 and ZnO phases. Similarly, Huang et al. [110] synthesized a series of ZnO-ZrO2 composite oxides via the co-deposited method and overall studied their phase structural evolution and catalytic properties in the CO2 hydrogenation reaction. It was found that with the increase in Zr content, the phase structure of ZnO-ZrO2 composite oxides evolved from the mixture of hexagonal ZnO phase and Zn-doped ZrO2 solid solution phases to pure Zn-doped ZrO2 solid solution phase, which showed high selectivity for the formation of CH3OH. Wang et al. [111] prepared a Cu/CeZrOX solid solution catalyst using the micro-impingement flow reactor with ultrasound, which had a better micro-mixing efficiency compared with the conventional batch method (Figure 8). Under reaction conditions of 240 °C, 3 MPa, 30,000 mL·gcat−1·h−1, the Cu/CeZrOX solid solution catalyst exhibited methanol selectivity of 55.3% and STY of 0.222 gMeOH·gcat−1·h−1 at CO2 conversion of 4.67%, which was higher than other traditional Cu-based catalysts [112,113,114]. This was ascribed to the synergistic interaction between the oxygen vacancies and a higher proportion of Cu+.

3.5. Other Catalysts

Recently, some other catalysts have been widely studied as well. For example, Li et al. [115] synthesized a Mo-Co-C-N catalyst with a ZIF-67 precursor under reaction conditions of 2 MPa, 6000 mL·gcat1·h1, and 275 °C, and the CO2 conversion, methanol selectivity, and STY reached 9.2%, 58.4%, and 0.106 gMeOH·gcat1·h1, respectively. This is mainly because the addition of N provided abundant oxygen vacancies, facilitating the dissociation and adsorption of CO2 and further increasing methanol selectivity [116]. Zeng et al. [117] developed boxlike assemblages of quasi-single-layer MoS2 nanosheets, which were edge-blocked by ZnS crystallites (denoted as h-MoS2/ZnS) via a MOF-engaged method. Under reaction conditions of 260 °C, 5 MPa, and 15,000 mL·gcat1·h1, the h-MoS2/ZnS catalyst can achieve STYMeOH of 0.93 gMeOH·gcat1·h1, CO2 conversion of 9.0%, and methanol selectivity of 67.3%. Li et al. [118] reported a novel Cd/TiO2 catalyst, with methanol selectivity of 81%, CO2 conversion of 15.8%, and little CH4 selectivity. The unique electronic properties of Cd clusters on the TiO2 substrates were the main reason for the high selectivity of CO2 hydrogenation to methanol through the HCOO* pathway at the interface catalytic site.

4. Mechanism Studies

For CO2 hydrogenation to methanol, various metal-based catalysts have been studied, and Cu-based catalysts are widely employed because of their higher catalytic activity as well as better selectivity to methanol. However, the active center and the underlying catalytic mechanism are still inconclusive. It is generally recognized that metallic Cu is the active phase, which mainly includes three types of recognized active sites: the interface of Cu and ZnO, Cu0, and Cu+ species. Many researchers have proposed various mechanisms of CO2 hydrogenation to methanol over Cu-based catalysts using experiments and DFT calculation methods. Three main mechanisms have been put forward based on the nuance of interface models and chosen pathways, namely the formate (HCOO*) mechanism, the RWGS + CO-Hydro mechanism, and the trans-COOH mechanism [119]. In the following content, we mainly focus on the recent theoretical advances in reaction pathways of the three mechanisms on Cu-based catalysts and their extension to other catalyst systems for guiding the design of efficient catalysts.

4.1. The HCOO Mechanism and r-HCOO Mechanism

The majority of researchers prefer that the most possible pathway for CO2 hydrogenation to methanol is the HCOO mechanism, in which formate (HCOO*) is the key intermediate formed by CO2 reacting with preadsorbed surface atomic H via either an Eley-Rideal (ER) [120,121] or Langmuir–Hinshelwood (LH) mechanism [33,122]. HCOO is then hydrogenated to dioxymethylene (*HCOOH), followed by further hydrogenation to *H2COOH. The obtained *H2COOH is cleaved to formaldehyde (*H2CO) which is further hydrogenated to methoxy (*H3CO) and methanol (*H3COH) [123].
Yang et al. [120] explored the possible reaction pathway for CO2 hydrogenation to methanol on Cu (111) and Cu nanoparticle surfaces through a combination of experimental and theoretical methods. DFT calculations show that the hydrogenation of CO2 on the Cu surface is carried out by formate intermediates, and the overall reaction rate is limited by the hydrogenation of formate and dioxymethylene. However, CO hydrogenation was easy to be hindered, which was mainly because CO tended to be converted to formyl, and the obtained formyl is unstable and preferred to dissociate into CO and H atoms on Cu. Sun et al. [60] proposed a possible reaction mechanism of CO2 hydrogenation to methanol, which followed the formate pathway on the dual active sites of CuZn10Zr through in situ diffuse reflectance infrared Fourier transform spectra. The exposed Cu0 took to participate in the H2 activation, while on the Cu+ species, the intermediate of formate from the CO2 and H2 co-adsorption was more inclined to produce methanol rather than other by-products, improving the methanol selectivity. Qu et al. [124] explored the SMSI between Cu and ZnAl2O4 spinel and found that the spinel catalyst had a double H2 activation center. The interstitial H atoms in the spinel promoted the conversion of CO2 to formate species, while the H2 dissociated by Cu could accelerate the formation of methanol from formate. Wang et al. [125] investigated the effects of transition metal doping on CO2 hydrogenation on Cu (211) surfaces using the DFT method and found that the doping of Rh, Ni, Co, and Ru can promote CO2 hydrogenation to COOH. Studies have also found that the hydrogenation of HCOO*/H2COO* is assumed to be the rate-controlling step in the CO2 hydrogenation to methanol process for the formate-intermediated route [126]. Han et al. [74] studied the mechanistic pathway of CO2 hydrogenation to methanol over hollow Cu@ZrO2 catalysts using in situ FTIR tests. The results revealed that methanol formation followed the formate-intermediated pathway, and the synergistic effect between Cu and ZrO2 at the interface was crucial in the reaction mechanism.
In addition, the H atoms spilled over to the support surface at the metal–support interface region, participated in the formation of the carbonaceous intermediate species, and were finally converted to HCOO*, CH2O*, CH3O*, and CH3OH. Liu et al. [127] systematically explored three possible paths of CO2 hydrogenation to methanol over a single- atom Zr-doped Cu-based catalyst from the kinetic and thermodynamic aspects of each elementary reaction by the DFT method (Figure 9). The results showed that methanol was produced by the HCOO pathway, which involved the intermediates of bi-HCOO*, HCOOH*, H2COO*, H2COOH*, H2CO*, and H3CO*. In addition, the H2COO*, H2CO*, and H3CO* hydrogenation were proven to be the rate-limiting steps. Liu et al. [89] studied the pathway of methanol synthesis on Pt/In2O3 catalysts with ZrO2 addition through theoretical calculations and found that CO2 activated at the In-Ov-Zr site tends to form methanol through the formate pathway. They also calculated the Gibbs free energy of CO adsorption on the catalyst model with or without ZrO2 by DFT, and the results showed that the Gibbs free energy of CO adsorption on the model with ZrO2 was lower. This is mainly because the addition of ZrO2 significantly reduced the interaction between CO and Pt sites, making it easier for CO to desorb from Pt sites to the gas phase under reaction conditions and inhibiting the poisoning of CO on Pt sites. Wang et al. [128] found that both Cu/ZnO and trace Au-added Cu/ZnO catalysts were carried out in the formate path in the methanol synthesis reaction through in situ DRIFTS characterization and DFT calculations. The metal–oxide interface with oxygen vacancy modification is considered as the active site, and the introduction of Au produces more active sites, which promotes the activation of CO2 and the modulation of intermediate species, thereby significantly improving the catalytic performance of methanol synthesis.
In addition to the aforementioned traditional formate pathway, a revised formate (r-HCOO) pathway has been proposed. In the work of Grabow et al. [129], the partial species proposed in previous mechanisms were considered. Meanwhile, the new intermediates such as HCOOH* and CH3O2* were also involved. It was found that the CH2O, HCOOH, and HCOOCH3 were determined to be crucial byproducts through an extensive set of periodic, self-consistent DFT calculations. Moreover, the DFT results showed that the HCOO* can be converted to HCOOH*preferentially, rather than H2CO2*, as suggested in previous mechanisms. HCOOH* was then further hydrogenated to form CH3O2*, which then transformed into CH2O* by cracking its OH group. CH3O* was the final intermediate before the formation of CH3OH*. Recently, Chen et al. [130] also found that the optimal mechanism on the W-doped Rh (111) surface for CO2 hydrogenation to methanol was the r-HCOO pathway. Liu et al. [31] systematically studied the mechanisms and kinetics of CO2 hydrogenation over a serial of MOF-808-coated single-atom metal catalysts via DFT calculation methods. Two plausible CO2 hydrogenation pathways on CuII-MOF-808 catalysts were studied, namely formate path and carboxyl path. As a result, the formate pathway was more favorable, and the conversion of H2COOH*-to-H2CO* was the key step from the kinetic perspective over CuII-MOF-808 (Figure 10). The study of Li et al. [131] showed that H2 dissociation on the ZrO2 sample needed to overcome the high activation energy, which was the rate-determining step of the reaction. While the introduction of Zn in the ZnZrOx catalyst promoted the heterolytic dissociation of H2, making the dissociation of H2COOH* the rate-determining step of the reaction, the Zn-O-Zr asymmetric sites on ZnZrOx catalyst showed a synergistic effect at the atomic level, which promoted the synthesis of methanol.

4.2. The RWGS + CO-Hydro Mechanism

For the RWGS + CO-Hydro pathway, CO2 is hydrogenated to form the CO* intermediate, followed by CO hydrogenation to methanol with HCO (formyl), H2CO (formaldehyde), and H3CO (methoxy) intermediates. The RWGS + CO-Hydro mechanism has been suggested for Cu-based systems and bimetallic catalysts [7,124]. Zhang et al. [75] designed Cu+/CeZrOx interfaces by employing a series of ceria-zirconia solid solution catalysts with different Ce/Zr ratios to form a Cu+-Ov-Ce3+ structure in which Cu+ atoms were combined with the oxygen vacancies of Ce. The in situ DRIFTS results showed that the oxygen vacancies in the solid solutions can not only effectively assist the CO2 activation, modulate the electronic state of copper, and promote the formation of Cu+/CeZrOX interfaces, but also contribute to stabilizing the key *CO intermediate compound to inhibit its desorption and facilitate its further hydrogenation to methanol through the RWGS + CO-Hydro pathway. Rasteiro et al. [132] found that the reaction pathway on the Ni-Ga alloy involved both the RWGS + CO-Hydro and formate routes employing DRIFTS analyses, as confirmed by the presence of formate, methoxy, and CO intermediates. Yang et al. [133] considered both the formate pathway and the RWGS + CO-Hydro route to investigate the CO2 hydrogenation process on metal-doped Cu (111) surfaces using DFT calculations and Kinetic Monte Carlo (KMC) simulations. The calculation results indicated that the addition of Pd, Rh, Pt, and Ni can promote the formation of methanol on Cu (111). Meanwhile, the conversion of CO and HCO to CH3OH through the RWGS + CO-Hydro pathway was found to be much faster than that via the formate pathway. Liu et al. [134] employed the DFT calculation to study the reaction energy and activation barriers of all elementary steps involved in three possible hydrogenation mechanisms (Figure 11a). The calculated results indicated that the activation barrier of the rate-determining step of the RWGS + CO-Hydro mechanism was the lowest, which can be completed through the CO2* → trans-COOH* → cis-COOH* → CO* → HCO* → HCOH* → H2COH* → CH3OH* pathway (Figure 11b). Ou et al. [135] also found that methanol was generated over Pd/TiO2 catalyst following the RWGS + CO-Hydro pathway rather than the formate pathway due to its lower activation barrier. Sun et al. [136] found through in situ DRIFTS experiments that Cu+ species in Cu/SiO2 catalysts prepared using the FSP method can inhibit the desorption of CO* by stabilizing CO* intermediates and further promote the hydrogenation of CO* to CH3OH, promoting the RWGS + CO-hydro pathway.

4.3. The Trans-COOH Mechanism

The earliest study of the trans-COOH mechanism with the hydrocarboxyl (COOH*) species as the first hydrogenated species was proposed by Zhao and co-workers [137] on the basis of the experiment of Yang et al. [138]. They found that in the presence of H2O, the hydrogenation of CO2 to produce hydrocarboxyl (trans-COOH) is kinetically more favorable than formate via a unique hydrogen transfer mechanism (Figure 12). The obtained trans-COOH was then converted to hydroxymethylidyne (COH) via dihydroxycarbene (COHOH) intermediates and then continuously hydrogenated for three steps to form hydroxymethylene (HCOH), hydroxymethyl (H2COH), and methanol, respectively.
In addition, Tang et al. [33] discussed three possible reaction pathways of CO2 hydrogenation to CH3OH on Ga3Ni5 (221) surfaces based on formate (HCOO), hydrocarboxyl (COOH) formations, and RWGS reaction with CO hydrogenation through DFT calculations. It was found that the reduction of CO2 to CH3OH was more likely to occur completely through the LH mechanism, and trans-COOH was proved to be the most favorable pathway. Similarly, Liu et al. [34] also considered the three possible reaction pathways according to DFT calculations to explore the reaction mechanisms on PdCu3 (111) surface. The results showed that the CO2 hydrogenation to CH3OH preferred to occur via the COOH pathway. The adsorbed CO2 reacted with H atoms to form COOH*, which was further hydrogenated to produce COHOH* and hydroxymethylidyne. And methanol was formed by the HCOH and H2COH intermediates.

5. Summary and Outlook

The combination of CO2 and renewable H2 can be used to synthesize valuable methanol, which is not only conducive to mitigating carbon emissions but also beneficial to achieving the artificial carbon cycle. In this work, we summed up the advances in studies of reaction thermodynamics, catalyst development, and mechanism study for the hydrogenation of CO2 to methanol. In terms of reaction thermodynamics, the low-temperature and high-pressure operating conditions are proven to be favorable. In terms of catalyst development, we mainly review the research progress of Cu-based catalysts, In2O3-based catalysts, bimetallic catalysts, solid solution catalysts, and other novel catalysts. For Cu-based catalysts, the suitable supports, promoters, and preparation methods contribute to improving the number or activity of active sites to enhance the H2 adsorption and the CO2 activation and promote the formation of methanol. For In2O3-based catalysts, several strategies are employed to improve the catalytic performances. More specifically, the introduction of metal promoters can promote H2 activation and oxygen vacancy formation. The oxide support is used to improve the dispersion of the active component and prevent the active component from sintering as well. For the bimetallic catalysts, the second metal can not only inhibit the aggregation of active species but also regulate the electronic structure around the active components. For the solid solution catalysts, their application further improves the thermal stability and structural performances of traditional catalysts. In addition, the catalytic performances of novel catalysts (e.g., MoS2, MOFs, and Cd/TiO2) are briefly introduced. In terms of the mechanism study, we summarized three possible reaction pathways, namely the formate (HCOO*) mechanism, the RWGS + CO-Hydro mechanism, and the trans-COOH mechanism, and reviewed their recent theoretical advances on Cu-based catalysts and their extension to other catalytic systems.
As an important carbon-neutral technology, CO2 hydrogenation to methanol has shown great promise for industrial application. In order to further improve its process performances and technical competitiveness, it is crucial to design efficient catalysts, explore reaction mechanisms, and conduct deep kinetic study and reactor design. From the perspective of the catalyst design, increasing the dispersion and the exposed surface of Cu, regulating the interaction between Cu and ZnO, and balancing the surface H/C ratio of active sites are the main focus of future research. From the perspective of the reaction mechanisms, exploring the role of the active site and investigating the factors of the rate-determining steps through in situ catalyst characterization techniques and theoretical simulation calculations can further guide the rational design of efficient catalysts. In addition, the reaction kinetics of the CO2 and CO hydrogenation involved in the methanol synthesis process is the foundation for the subsequent reactor design and optimization of reactors, which provides constructive suggestions for large-scale industrial applications.

Author Contributions

Conceptualization, L.X. and X.C.; methodology, C.D.; software, K.H.; validation, L.X., X.C. and C.D.; formal analysis, R.G.; investigation, K.H.; resources, C.Z.; data curation, L.Z.; writing—original draft preparation, X.C.; writing—review and editing, L.X.; visualization, L.W.; supervision, L.Z.; project administration, R.G.; funding acquisition, C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the “Next Generation Carbon Upcycling Project” (Project No. 2017M1A2A2043133) through the National Research Foundation (NRF) funded by the Ministry of Science and ICT, Republic of Korea. We also appreciate the Natural Science Foundation of Jiangsu Province (BK20200694, 20KJB530002, and 21KJB480014), and the open program of the State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (2021-K32).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data are presented in the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chauvy, R.; De Weireld, G. CO2 utilization technologies in Europe: A short review. Energy Technol. 2020, 8, 2000627. [Google Scholar] [CrossRef]
  2. Ye, R.-P.; Ding, J.; Gong, W.; Argyle, M.D.; Zhong, Q.; Wang, Y.; Russell, C.K.; Xu, Z.; Russell, A.G.; Li, Q.; et al. CO2 hydrogenation to high-value products via heterogeneous catalysis. Nat. Commun. 2019, 10, 5698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Atsbha, T.A.; Yoon, T.; Seongho, P.; Lee, C.-J. A review on the catalytic conversion of CO2 using H2 for synthesis of CO, methanol, and hydrocarbons. J. CO2 Util. 2021, 44, 101413. [Google Scholar] [CrossRef]
  4. Liu, Z.; Gao, X.; Liu, B.; Ma, Q.; Zhao, T.-S.; Zang, J. Recent advances in thermal catalytic CO2 methanation on hydrotalcite-derived catalysts. Fuel 2022, 321, 124115. [Google Scholar] [CrossRef]
  5. Ra, E.C.; Kim, K.Y.; Kim, E.H.; Lee, H.; An, K.; Lee, J.S. Recycling Carbon Dioxide through Catalytic Hydrogenation: Recent Key Developments and Perspectives. ACS Catal. 2020, 10, 11318–11345. [Google Scholar]
  6. Zhou, Z.; Gao, P. Direct carbon dioxide hydrogenation to produce bulk chemicals and liquid fuels via heterogeneous catalysis. Chin. J. Catal. 2022, 43, 2045–2056. [Google Scholar] [CrossRef]
  7. Zhong, J.; Yang, X.; Wu, Z.; Liang, B.; Huang, Y.; Zhang, T. State of the art and perspectives in heterogeneous catalysis of CO2 hydrogenation to methanol. Chem. Soc. Rev. 2020, 49, 1385–1413. [Google Scholar] [CrossRef] [PubMed]
  8. Hepburn, C.; Adlen, E.; Beddington, J.; Carter, E.A.; Fuss, S.; Dowell, N.M.; Minx, J.C.; Smith, P.; Williams, C.K. The technological and economic prospects for CO2 utilization and removal. Nature 2019, 575, 87–97. [Google Scholar] [CrossRef] [Green Version]
  9. Fernández-Dacosta, C.; Van Der Spek, M.; Hung, C.R.; Oregionni, G.D.; Skagestad, R.; Parihar, P.; Gokak, D.; Strømman, A.H.; Ramirez, A. Prospective techno-economic and environmental assessment of carbon capture at a refinery and CO2 utilisation in polyol synthesis. J. CO2 Util. 2017, 21, 405–422. [Google Scholar] [CrossRef]
  10. Sha, F.; Han, Z.; Tang, S.; Wang, J.; Li, C. Hydrogenation of Carbon Dioxide to Methanol over Non-Cu-based Heterogeneous Catalysts. ChemSusChem 2020, 13, 6160–6181. [Google Scholar]
  11. Du, X.L.; Jiang, Z.; Su, D.S.; Wang, J.Q. Research progress on the indirect hydrogenation of carbon dioxide to methanol. ChemSusChem 2016, 9, 322–332. [Google Scholar] [CrossRef] [PubMed]
  12. Studt, F.; Behrens, M.; Kunkes, E.L.; Thomas, N.; Zander, S.; Tarasov, A.; Schumann, J.; Frei, E.; Varley, J.B.; Abild-Pedersen, F. The mechanism of CO and CO2 hydrogenation to methanol over Cu-based catalysts. ChemCatChem 2015, 7, 1105–1111. [Google Scholar] [CrossRef] [Green Version]
  13. Xie, S.; Li, Z.; Li, H.; Fang, Y. Integration of carbon capture with heterogeneous catalysis toward methanol production: Chemistry, challenges, and opportunities. Catal. Rev. 2023, 1–40. [Google Scholar] [CrossRef]
  14. Niu, J.; Liu, H.; Jin, Y.; Fan, B.; Qi, W.; Ran, J. Comprehensive review of Cu-based CO2 hydrogenation to CH3OH: Insights from experimental work and theoretical analysis. Int. J. Hydrogen Energy 2022, 15, 9183–9200. [Google Scholar]
  15. Guil-López, R.; Mota, N.; Llorente, J.; Millán, E.; Pawelec, B.; Fierro, J.L.G.; Navarro, R. Methanol synthesis from CO2: A review of the latest developments in heterogeneous catalysis. Materials 2019, 12, 3902. [Google Scholar] [PubMed] [Green Version]
  16. Kanuri, S.; Roy, S.; Chakraborty, C.; Datta, S.P.; Singh, S.; Dinda, S. An insight of CO2 hydrogenation to methanol synthesis: Thermodynamics, catalysts, operating parameters, and reaction mechanism. Int. J. Energy Res. 2022, 46, 5503–5522. [Google Scholar]
  17. 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]
  18. Wang, J.; Li, G.; Li, Z.; Tang, C.; Feng, Z.; An, H.; Liu, H.; Liu, T.; Li, C. A highly selective and stable ZnO-ZrO2 solid solution catalyst for CO2 hydrogenation to methanol. Sci. Adv. 2017, 3, e1701290. [Google Scholar]
  19. Sharma, P.; Ho, P.H.; Shao, J.; Creaser, D.; Olsson, L. Role of ZrO2 and CeO2 support on the In2O3 catalyst activity for CO2 hydrogenation. Fuel 2023, 331, 125878. [Google Scholar] [CrossRef]
  20. Wang, J.; Zhang, G.; Zhu, J.; Zhang, X.; Ding, F.; Zhang, A.; Guo, X.; Song, C. CO2 hydrogenation to methanol over In2O3-based catalysts: From mechanism to catalyst development. ACS Catal. 2021, 11, 1406–1423. [Google Scholar]
  21. Ay, S.; Ozdemir, M.; Melikoglu, M. Effects of metal promotion on the performance, catalytic activity, selectivity and deactivation rates of Cu/ZnO/Al2O3 catalysts for methanol synthesis. Chem. Eng. Res. Des. 2021, 175, 146–160. [Google Scholar]
  22. Natesakhawat, S.; Lekse, J.W.; Baltrus, J.P.; Ohodnicki, P.R., Jr.; Howard, B.H.; Deng, X.; Matranga, C. Active sites and structure–activity relationships of copper-based catalysts for carbon dioxide hydrogenation to methanol. ACS Catal. 2012, 2, 1667–1676. [Google Scholar]
  23. Ye, H.; Na, W.; Gao, W.; Wang, H. Carbon- Modified CuO/ZnO Catalyst with High Oxygen Vacancy for CO2 Hydrogenation to Methanol. Energy Technol. 2020, 8, 2000194. [Google Scholar]
  24. Guo, Y.; Guo, X.; Song, C.; Han, X.; Liu, H.; Zhao, Z. Capsule- Structured Copper– Zinc Catalyst for Highly Efficient Hydrogenation of Carbon Dioxide to Methanol. ChemSusChem 2019, 12, 4916–4926. [Google Scholar] [CrossRef] [PubMed]
  25. Kattel, S.; Yan, B.; Chen, J.G.; Liu, P. CO2 hydrogenation on Pt, Pt/SiO2 and Pt/TiO2: Importance of synergy between Pt and oxide support. J. Catal. 2016, 343, 115–126. [Google Scholar]
  26. Fan, L.; Fujimoto, K. Development of active and stable supported noble metal catalysts for hydrogenation of carbon dioxide to methanol. Energy Convers. Manag. 1995, 36, 633–636. [Google Scholar]
  27. Tawalbeh, M.; Javed, R.M.N.; Al-Othman, A.; Almomani, F. The novel contribution of non-noble metal catalysts for intensified carbon dioxide hydrogenation: Recent challenges and opportunities. Energy Convers. Manag. 2023, 279, 116755. [Google Scholar]
  28. Frei, M.S.; Mondelli, C.; Pérez-Ramírez, J. Development of In2O3-based catalysts for CO2-based methanol production. Chimia 2020, 74, 257–262. [Google Scholar] [CrossRef]
  29. Wang, J.; Sun, K.; Jia, X.; Liu, C.-J. CO2 hydrogenation to methanol over Rh/In2O3 catalyst. Catal. Today 2021, 365, 341–347. [Google Scholar] [CrossRef]
  30. Frei, M.S.; Mondelli, C.; Cesarini, A.; Krumeich, F.; Hauert, R.; Stewart, J.A.; Ferré, D.C.; Pérez-Ramírez, J. Role of zirconia in indium oxide-catalyzed CO2 hydrogenation to methanol. ACS Catal. 2019, 10, 1133–1145. [Google Scholar] [CrossRef]
  31. Liu, J.; Xue, W.; Zhang, W.; Mei, D. Theoretical Study on the Catalytic CO2 Hydrogenation over the MOF-808-Encapsulated Single-Atom Metal Catalysts. J. Phys. Chem. C 2023, 127, 4051–4062. [Google Scholar] [CrossRef]
  32. 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]
  33. Liu, L.; Yao, H.; Jiang, Z.; Fang, T. Theoretical study of methanol synthesis from CO2 hydrogenation on PdCu3(111) surface. Appl. Surf. Sci. 2018, 451, 333–345. [Google Scholar] [CrossRef]
  34. Ren, M.; Zhang, Y.; Wang, X.; Qiu, H. Catalytic hydrogenation of CO2 to methanol: A review. Catalysts 2022, 12, 403. [Google Scholar] [CrossRef]
  35. Jadhav, S.G.; Vaidya, P.D.; Bhanage, B.M.; Joshi, J.B. Catalytic carbon dioxide hydrogenation to methanol: A review of recent studies. Chem. Eng. Res. Des. 2014, 92, 2557–2567. [Google Scholar] [CrossRef]
  36. Zhang, C.; Yang, H.; Gao, P.; Zhu, H.; Zhong, L.; Wang, H.; Wei, W.; Sun, Y. Preparation and CO2 hydrogenation catalytic properties of alumina microsphere supported Cu-based catalyst by deposition-precipitation method. J. CO2 Util. 2017, 17, 263–272. [Google Scholar] [CrossRef]
  37. Wu, J.; Saito, M.; Takeuchi, M.; Watanabe, T. The stability of Cu/ZnO-based catalysts in methanol synthesis from a CO2 rich feed and from a CO-rich feed. Appl. Catal. A Gen. 2001, 218, 235–240. [Google Scholar] [CrossRef]
  38. Arakawa, H. Research and development on new synthetic routes for basic chemicals by catalytic hydrogenation of CO2. Stud. Surf. Sci. Catal. 1998, 114, 19–30. [Google Scholar]
  39. Zachopoulos, A.; Heracleous, E. Overcoming the equilibrium barriers of CO2 hydrogenation to methanol via water sorption: A thermodynamic analysis. J. CO2 Util. 2017, 21, 360–367. [Google Scholar] [CrossRef]
  40. Jia, X.; Sun, K.; Wang, J.; Shen, C.; Liu, C.-J. Selective hydrogenation of CO2 to methanol over Ni/In2O3 catalyst. J. Energy Chem. 2020, 50, 409–415. [Google Scholar] [CrossRef]
  41. Yuan, Y.; Qi, L.; Guo, T.; Hu, X.; He, Y.; Guo, Q. A review on the development of catalysts and technologies of CO2 hydrogenation to produce methanol. Chem. Eng. Commun. 2022, 210, 1–31. [Google Scholar]
  42. Etim, U.J.; Song, Y.; Zhong, Z. Improving the Cu/ZnO-Based Catalysts for Carbon Dioxide Hydrogenation to Methanol, and the Use of Methanol As a Renewable Energy Storage Media. Front. Energy Res. 2020, 8, 545431. [Google Scholar] [CrossRef]
  43. Jiang, Q.; Liu, Y.; Dintzer, T.; Luo, J.; Parkhomenko, K.; Roger, A.-C. Tuning the highly dispersed metallic Cu species via manipulating Brønsted acid sites of mesoporous aluminosilicate support for CO2 hydrogenation reactions. Appl. Catal. B Environ. 2020, 269, 118804. [Google Scholar]
  44. Ma, Q.; Geng, M.; Zhang, J.; Zhang, X.; Zhao, T.-S. Enhanced Catalytic Performance for CO2 Hydrogenation to Methanol over N-doped Graphene Incorporated Cu-ZnO-Al2O3 Catalysts. ChemistrySelect 2019, 4, 78–83. [Google Scholar] [CrossRef]
  45. Fang, X.; Men, Y.; Wu, F.; Zhao, Q.; Singh, R.; Xiao, P.; Du, T.; Webley, P.A. Improved methanol yield and selectivity from CO2 hydrogenation using a novel Cu-ZnO-ZrO2 catalyst supported on Mg-Al layered double hydroxide (LDH). J. CO2 Util. 2019, 29, 57–64. [Google Scholar] [CrossRef]
  46. Kubovics, M.; Trigo, A.; Sánchez, A.; Marbán, G.; Borrás, A.; Vico, J.M.; Periago, A.M.L.; Domingo, C. Role of Graphene Oxide Aerogel Support on the CuZnO Catalytic Activity: Enhancing Methanol Selectivity in the Hydrogenation Reaction of CO2. ChemCatChem 2022, 14, e202200607. [Google Scholar]
  47. Duma, Z.G.; Moma, J.; Langmi, H.W.; Louis, B.; Parkhomenko, K.; Musyoka, N.M. Towards High CO2 Conversions Using Cu/Zn Catalysts Supported on Aluminum Fumarate Metal-Organic Framework for Methanol Synthesis. Catalysts 2022, 12, 1104. [Google Scholar] [CrossRef]
  48. Sun, X.; Jin, Y.; Cheng, Z.; Lan, G.; Wang, X.; Qiu, Y.; Wang, Y.; Liu, H.; Li, Y. Dual active sites over Cu-ZnO-ZrO2 catalysts for carbon dioxide hydrogenation to methanol. J. Environ. Sci. 2023, 131, 162–172. [Google Scholar] [CrossRef]
  49. Cai, W.; Chen, Q.; Wang, F.; Li, Z.; Yu, H.; Zhang, S.; Cui, L.; Li, C. Comparison of the Promoted CuZnMxOy (M: Ga, Fe) Catalysts for CO2 Hydrogenation to Methanol. Catal. Lett. 2019, 149, 2508–2518. [Google Scholar] [CrossRef]
  50. Prašnikar, A.; Dasireddy, V.D.B.C.; Likozar, B. Scalable combustion synthesis of copper-based perovskite catalysts for CO2 reduction to methanol: Reaction structure-activity relationships, kinetics, and stability. Chem. Eng. Sci. 2022, 250, 117423. [Google Scholar]
  51. Han, X.; Li, M.; Chang, X.; Hao, Z.; Chen, J.; Pan, Y.; Kawi, S.; Ma, X. Hollow structuredCu@ZrO2 derived from Zr-MOF for selective hydrogenation of CO2 to methanol. J. Energy Chem. 2022, 71, 277–287. [Google Scholar] [CrossRef]
  52. Zhang, J.; Sun, X.; Wu, C.; Hang, W.; Hu, X.; Qiao, D.; Yan, B. Engineering Cu+/CeZrO interfaces to promote CO2 hydrogenation to methanol. J. Energy Chem. 2023, 77, 45–53. [Google Scholar] [CrossRef]
  53. Li, J.; Du, T.; Li, Y.; Jia, H.; Wang, Y.; Song, Y.; Fang, X. Novel layered triple hydroxide sphere CO2 adsorbent supported copper nanocluster catalyst for efficient methanol synthesis via CO2 hydrogenation. J. Catal. 2022, 409, 24–32. [Google Scholar] [CrossRef]
  54. Fang, X.; Xi, Y.; Jia, H.; Chen, C.; Wang, Y.; Song, Y.; Du, T. Tetragonal zirconia based ternary ZnO-ZrO2-MOX solid solution catalysts for highly selective conversion of CO2 to methanol at High reaction temperature. J. Ind. Eng. Chem. 2021, 35, 8558–8584. [Google Scholar]
  55. Ipatieff, V.N.; Monroe, G.S. Synthesis of Methanol from Carbon Dioxide and Hydrogen over Copper-Alumina Catalysts. Mechanism of Reaction. J. Am. Chem. Soc. 1945, 67, 2168–2171. [Google Scholar] [CrossRef]
  56. Li, M.M.J.; Chen, C.; Ayvalı, T.; Suo, H.; Zheng, J.; Teixeira, I.F.; Ye, L.; Zou, H.; O’Hare, D.; Tsang, S.C.E. CO2 Hydrogenation to Methanol over Catalysts Derived from Single Cationic Layer CuZnGa LDH Precursors. ACS Catal. 2018, 8, 4390–4401. [Google Scholar] [CrossRef]
  57. Lo, I.C.; Wu, H.-S. Methanol formation from carbon dioxide hydrogenation using Cu/ZnO/Al2O3 catalyst. J. Taiwan Inst. Chem. Eng. 2019, 98, 124–131. [Google Scholar] [CrossRef]
  58. Lu, P.; Chizema, L.G.; Hondo, E.; Tong, M.; Xing, C.; Lu, C.; Mei, Y.; Yang, R. CO2 Hydrogenation to Methanol via In- situ Reduced Cu/ZnO Catalyst Prepared by Formic acid Assisted Grinding. ChemistrySelect 2019, 4, 5667–5677. [Google Scholar] [CrossRef]
  59. Du, S.; Tang, W.; Lu, X.; Wang, S.; Guo, Y.; Gao, P.X. Cu- Decorated ZnO Nanorod Array Integrated Structured Catalysts for Low- Pressure CO2 Hydrogenation to Methanol. Adv. Mater. Interfaces 2017, 5, 1700730. [Google Scholar] [CrossRef]
  60. Prachumsai, W.; Pangtaisong, S.; Assabumrungrat, S.; Bunruam, P.; Nakvachiratrakul, C.; Saebea, D.; Praserthdam, P.; Soisuwan, S. Carbon dioxide reduction to synthetic fuel on zirconia supported copper-based catalysts and gibbs free energy minimization: Methanol and dimethyl ether synthesis. J. Environ. Chem. Eng. 2021, 9, 104979. [Google Scholar]
  61. Hamryszak, L.; Madej-Lachowska, M.; Grzesik, M.; Kocot, K.; Ruggiero-Mikolajczyk, M. Effect of addition of Ce, Cr or/and Ga to a Cu/Zn/Zr catalyst on the methanol synthesis from carbon dioxide and hydrogen. Przem. Chem. 2019, 98, 133–137. [Google Scholar]
  62. Wang, G.; Mao, D.; Guo, X.; Yu, J. Methanol synthesis from CO2 hydrogenation over CuO-ZnO-ZrO2-MxOy catalysts (M = Cr, Mo and W). Int. J. Hydrogen Energy 2019, 44, 4197–4207. [Google Scholar] [CrossRef]
  63. Zhang, S.; Wu, Z.; Liu, X.; Hua, K.; Shao, Z.; Wei, B.; Huang, C.; Wang, H.; Sun, Y. A Short Review of Recent Advances in Direct CO2 Hydrogenation to Alcohols. Top. Catal. 2021, 64, 371–394. [Google Scholar]
  64. Shawabkeh, R.A.; Faqir, N.M.; Rawajfeh, K.M.; Hussein, I.A.; Hamza, A. Adsorption of CO2 on Cu/SiO2 nano-catalyst: Experimental and theoretical study. Appl. Surf. Sci. 2022, 586, 152726. [Google Scholar]
  65. Cruz-Navarro, D.S.; Torres-Rodríguez, M.; Gutiérrez-Arzaluz, M.; Mugica-Álvarez, V.; Pergher, S.B. Comparative Study of Cu/ZSM-5 Catalysts Synthesized by Two Ion-Exchange Methods. Crystals 2022, 12, 1104. [Google Scholar]
  66. Biswal, T.; Shadangi, K.P.; Sarangi, P.K.; Srivastava, R.K. Conversion of carbon dioxide to methanol: A comprehensive review. Chemosphere 2022, 298, 134299. [Google Scholar] [CrossRef]
  67. Cored, J.; Lopes, C.W.; Liu, L.; Soriano, J.; Agostini, G.; Solsona, B.; Sánchez-Tovar, R.; Concepción, P. Cu-Ga3+-doped wurtzite ZnO interface as driving force for enhanced methanol production in co-precipitated Cu/ZnO/Ga2O3 catalysts. J. Catal. 2022, 407, 149–161. [Google Scholar] [CrossRef]
  68. Sang, G.; Ran, J.; Huang, X.; Ou, Z.; Tang, L. Understanding the role of Ga on the activation mechanism of CO2 over modified Cu surface by DFT calculation. Mol. Catal. 2022, 528, 112477. [Google Scholar] [CrossRef]
  69. L’hospital, V.; Angelo, L.; Zimmermann, Y.; Parkhomenko, K.; Roger, A.-C. Influence of the Zn/Zr ratio in the support of a copper-based catalyst for the synthesis of methanol from CO2. Catal. Today 2021, 369, 95–104. [Google Scholar] [CrossRef]
  70. Guo, H.; Li, Q.; Zhang, H.; Peng, F.; Xiong, L.; Yao, S.; Huang, C.; Chen, X. CO2 hydrogenation over acid-activated Attapulgite/Ce0.75Zr0.25O2 nanocomposite supported Cu-ZnO based catalysts. Mol. Catal. 2019, 476, 110499. [Google Scholar] [CrossRef]
  71. Koh, M.K.; Wong, Y.J.; Mohamed, A.R. Effect of synthesis method on the ultimate properties of copper-based catalysts supported on KIT-6 for direct CO2 hydrogenation to methanol. Mater. Today Chem. 2021, 21, 100489. [Google Scholar] [CrossRef]
  72. Paris, C.; Karelovic, A.; Manrique, R.; Le Bras, S.; Devred, F.; Vykoukal, V.; Styskalik, A.; Eloy, P.; Debecker, D.P. CO2 Hydrogenation to Methanol with Ga- and Zn-Doped Mesoporous Cu/SiO2 Catalysts Prepared by the Aerosol-Assisted Sol-Gel Process. ChemSusChem 2020, 13, 6409–6417. [Google Scholar]
  73. Xiao, J.; Mao, D.; Wang, G.; Guo, X.; Yu, J. CO2 hydrogenation to methanol over CuOZnOTiO2ZrO2 catalyst prepared by a facile solid-state route: The significant influence of assistant complexing agents. Int. J. Hydrogen Energy 2019, 44, 14831–14841. [Google Scholar] [CrossRef]
  74. Choi, H.; Oh, S.; Tran, S.B.T.; Park, J.Y. Size-controlled model Ni catalysts on Ga2O3 for CO2 hydrogenation to methanol. J. Catal. 2019, 376, 68–76. [Google Scholar] [CrossRef]
  75. Ouyang, L.; Noël, V.; Courty, A.; Campagne, J.-M.; Ouali, A.; Vrancken, E. Copper Nanoparticles with a Tunable Size: Implications for Plasmonic Catalysis. ACS Appl. Nano Mater. 2022, 5, 2839–2847. [Google Scholar] [CrossRef]
  76. Ali, S.; Kumar, D.; Mondal, K.C.; El-Naas, M.H. Development of highly active Cu-based CO2 hydrogenation catalysts by solution combustion synthesis (SCS): Effects of synthesis variables. Catal. Commun. 2022, 172, 106543. [Google Scholar]
  77. Wang, S.; Song, L.; Qu, Z. Cu/ZnAl2O4 catalysts prepared by ammonia evaporation method: Improving methanol selectivity in CO2 hydrogenation via regulation of metal-support interaction. Chem. Eng. J. 2023, 469, 144008. [Google Scholar]
  78. Murthy, P.S.; Liang, W.; Jiang, Y.; Huang, J. Cu-based nanocatalysts for CO2 hydrogenation to methanol. Energy 2021, 35, 8558–8584. [Google Scholar] [CrossRef]
  79. Zafar, F.; Zhao, R.; Ali, M.; Park, Y.M.; Roh, Y.-S.; Gao, X.; Tian, J.; Bae, J.W. Unprecedented contributions of In2O3 promoter on ordered mesoporous Cu/Al2O3 for CO2 hydrogenation to oxygenates. Chem. Eng. J. 2022, 439, 135649. [Google Scholar] [CrossRef]
  80. Tian, G.; Wu, Y.; Wu, S.; Huang, S.; Gao, J. Solid-State Synthesis of Pd/In2O3 Catalysts for CO2 Hydrogenation to Methanol. Catal. Lett. 2023, 153, 903–910. [Google Scholar] [CrossRef]
  81. Rui, N.; Wang, Z.; Sun, K.; Ye, J.; Ge, Q.; Liu, C. CO2 hydrogenation to methanol over Pd/In2O3: Effects of Pd and oxygen vacancy. Appl. Catal. B Environ. 2017, 218, 488–497. [Google Scholar] [CrossRef]
  82. Sun, K.; Rui, N.; Zhang, Z.; Sun, Z.; Ge, Q.; Liu, C.-J. A highly active Pt/In2O3 catalyst for CO2 hydrogenation to methanol with enhanced stability. Green Chem. 2020, 22, 5059–5066. [Google Scholar] [CrossRef]
  83. 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]
  84. Sun, K.; Zhang, Z.; Shen, C.; Rui, N.; Liu, C.-J. The feasibility study of the indium oxide supported silver catalyst for selective hydrogenation of CO2 to methanol. Green Energy Environ. 2022, 7, 807–817. [Google Scholar] [CrossRef]
  85. Zhu, J.; Cannizzaro, F.; Liu, L.; Zhang, H.; Kosinov, N.; Filot, I.A.W.; Rabeah, J.; Brückner, A.; Hensen, E.J.M. Ni−In Synergy in CO2 Hydrogenation to Methanol. ACS Catal. 2021, 11, 11371–11384. [Google Scholar] [CrossRef] [PubMed]
  86. 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]
  87. Wu, Q.; Shen, C.; Rui, N.; Sun, K.; Liu, C.-J. Experimental and theoretical studies of CO2 hydrogenation to methanol on Ru/In2O3. J. CO2 Util. 2021, 53, 101720. [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 studies. ACS Catal. 2021, 11, 4036–4046. [Google Scholar] [CrossRef]
  89. Sun, K.; Shen, C.; Zou, R.; Liu, C. Highly active Pt/In2O3-ZrO2 catalyst for CO2 hydrogenation to methanol with enhanced CO tolerance: The effects of ZrO2. Appl. Catal. B Environ. 2023, 320, 122018. [Google Scholar] [CrossRef]
  90. Li, W.; Huang, H.; Li, H.; Zhang, W.; Liu, H. Facile synthesis of pure monoclinic and tetragonal zirconia nanoparticles and their phase effects on the behavior of supported molybdena catalysts for methanol-selective oxidation. Langmuir 2008, 24, 8358–8366. [Google Scholar] [CrossRef]
  91. Xu, Y.M.; Ding, Z.L.; Qiu, R.; Hou, R.J. Effect of support and reduction temperature in the hydrogenation of CO2 over the Cu-Pd bimetallic catalyst with high Cu/Pd ratio. Int. J. Hydrogen Energy 2022, 47, 27973–27985. [Google Scholar] [CrossRef]
  92. Lin, F.W.; Jiang, X.; Boreriboon, N.; Song, C.S.; Wang, Z.H.; Cen, K.F. CO2 hydrogenation to methanol over bimetallic Pd-Cu catalysts supported on TiO2-CeO2 and TiO2-ZrO2. Catal. Today 2021, 371, 150–161. [Google Scholar] [CrossRef]
  93. Han, C.Y.; Zhang, H.T.; Li, C.M.; Huang, H.; Wang, S.; Wang, P.; Li, J.P. The regulation of Cu-ZnO interface by Cu-Zn bimetallic metal organic framework-templated strategy for enhanced CO2 hydrogenation to methanol. Appl. Catal. A Gen. 2022, 643, 118805. [Google Scholar] [CrossRef]
  94. Yu, J.; Chen, G.Q.; Guo, Q.S.; Guo, X.M.; Da Costa, P.; Mao, D.S. Ultrasmall bimetallic Cu/ZnOX nanoparticles encapsulated in UiO-66 by deposition-precipitation method for CO2 hydrogenation to methanol. Fuel 2022, 324, 124694. [Google Scholar] [CrossRef]
  95. Wang, C.R.; Fang, Y.H.; Liang, G.F.; Lv, X.Y.; Duan, H.M.; Li, Y.D.; Chen, D.F.; Long, M.J. Mechanistic study of Cu-Ni bimetallic catalysts supported by graphene derivatives for hydrogenation of CO2 to methanol. J. CO2 Util. 2021, 49, 101542. [Google Scholar] [CrossRef]
  96. Shi, Z.S.; Pan, M.; Wei, X.L.; Wu, D.F. Cu-In intermetallic compounds as highly active catalysts for CH3OH formation from CO2 hydrogenation. Int. J. Energy Res. 2022, 46, 1285–1298. [Google Scholar] [CrossRef]
  97. Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjaer, C.F.; Hummelshoj, J.S.; Dahl, S.; Chorkendorff, I.; Norskov, J.K. Discovery of a Ni-Ga catalyst for carbon dioxide reduction to methanol. Nat. Chem. 2014, 6, 320–324. [Google Scholar] [CrossRef] [PubMed]
  98. Rasteiro, L.F.; De Sousa, R.A.; Vieira, L.H.; Ocampo-Restrepo, V.K.; Verga, L.G.; Assaf, J.M.; Da Silva, J.L.F.; Assaf, E.M. Insights into the alloy-support synergistic effects for the CO2 hydrogenation towards methanol on oxide-supported Ni5Ga3 catalysts: An experimental and DFT study. Appl. Catal. B Environ. 2022, 302, 120842. [Google Scholar]
  99. Duyar, M.S.; Gallo, A.; Snider, J.L.; Jaramillo, T.F. Low-pressure methanol synthesis from CO2 over metal-promoted Ni-Ga intermetallic catalysts. J. CO2 Util. 2020, 39, 101151. [Google Scholar] [CrossRef]
  100. Ojelade, O.A.; Zaman, S.F.; Daous, M.A.; Al-Zahrani, A.A.; Malik, A.S.; Driss, H.; Shterk, G.; Gascon, J. Optimizing Pd:Zn molar ratio in PdZn/CeO2 for CO2 hydrogenation to methanol. Appl. Catal. A Gen. 2019, 584, 117185. [Google Scholar] [CrossRef]
  101. Yin, Y.; Hu, B.; Li, X.; Zhou, X.; Hong, X.; Liu, G. Pd@zeolitic imidazolate framework-8 derived PdZn alloy catalysts for efficient hydrogenation of CO2 to methanol. Appl. Catal. B Environ. 2018, 234, 143–152. [Google Scholar] [CrossRef]
  102. Snider, L.; Streibel, V.; Hubert, M.A.; Choksi, T.S.; Valle, E.; Upham, D.C.; Schumann, J.; Duyar, M.S.; Gallo, A.; Abild-Pedersen, F.; et al. Revealing the Synergy between Oxidend Alloy Phases on the Performance of Bimetallic In–Pd Catalysts for CO2 Hydrogenation to Methanol. ACS Catal. 2019, 9, 3399–3412. [Google Scholar] [CrossRef]
  103. Cai, Z.J.; Huang, M.; Dai, J.J.; Zhan, G.W.; Sun, F.L.; Zhuang, G.L.; Wang, Y.Y.; Tian, P.; Chen, B.; Ullah, S.; et al. Fabrication of Pd/In2O3 Nanocatalysts Derived from MIL-8(In) Loaded with Molecular Metalloporphyrin (TCPP(Pd)) Toward CO2 Hydrogenation to Methanol. ACS Catal. 2022, 12, 709–723. [Google Scholar] [CrossRef]
  104. Collins, S.E.; Baltanas, M.A.; Delgado, J.J.; Borgna, A.; Bonivardi, A.L. CO2 hydrogenationo methanol on Ga2O3-Pd/SiO2 catalysts: Dual oxide-metal sites or (bi)metallic surface sites. Catal. Today 2021, 381, 154–162. [Google Scholar] [CrossRef]
  105. Li, M.M.; Zou, H.; Zheng, J.; Wu, T.S.; Chan, T.S.; Soo, Y.L.; Wu, X.P.; Gong, X.Q.; Chen, T.; Roy, K.; et al. Methanol Synthesis at a Wide Range of H2/CO2 Ratios over a Rh-In Bimetallic Catalyst. Angew. Chem Int Edit. 2020, 59, 16039–16046. [Google Scholar] [CrossRef]
  106. Geng, F.Y.; Zhan, X.; Hicks, J.C. Promoting Methanol Synthesis and Inhibiting CO2 Methanation with Bimetallic In-Ru Catalysts. ACS Sustain. Chem. Eng. 2021, 9, 11891–11902. [Google Scholar] [CrossRef]
  107. Sha, F.; Tang, C.Z.; Tang, S.; Wang, Q.N.; Han, Z.; Wang, J.J.; Li, C. The promoting role of Gan ZnZrOx solid solution catalyst for CO2 hydrogenation to methanol. J. Catal. 2021, 04, 383–392. [Google Scholar] [CrossRef]
  108. Han, Z.; Tang, C.; Sha, F.; Tang, S.; Wang, J. CO2 hydrogenation to methanol on ZnO-ZrO2 solid solution catalysts with ordered mesoporous structure. J. Catal. 2021, 396, 242–250. [Google Scholar] [CrossRef]
  109. Tada, S.; Ochiai, N.; Kinoshita, H.; Yoshida, M.; Shimada, N.; Joutsuka, T.; Nishijima, M.; Honma, T.; Kobayashi, Y.; Iyoki, K. Active Sites on ZnXZr1-XO2-X Solid Solution Catalysts or CO2-to-Methanol Hydrogenation. ACS Catal. 2022, 12, 7748–7759. [Google Scholar] [CrossRef]
  110. Ding, Q.; Li, Z.R.; Xiong, W.; Zhang, Y.S.; Ye, A.A.I.; Huang, W.X. Structural evolution and catalytic performance in CO2 hydrogenation reaction of ZnO-ZrO2 composite oxides. Appl. Surf. Sci. 2022, 587, 152884. [Google Scholar] [CrossRef]
  111. Wang, Y.D.; Yu, H.R.; Hu, Q.; Huang, Y.P.; Wang, X.M.; Wang, Y.H.; Wang, F.H. Application of microimpinging stream reactor coupled with ultrasound in Cu/CeZrOx solid solution catalyst reparation for CO2 hydrogenation to methanol. Renew Energy 2023, 202, 834–843. [Google Scholar] [CrossRef]
  112. Shi, Z.S.; Tan, Q.Q.; Wu, D.F. Ternary copper-cerium-zirconium mixed metal oxide catalyst or direct CO2 hydrogenation to methanol. Mater. Chem. Phys. 2018, 219, 263–272. [Google Scholar] [CrossRef]
  113. Ban, H.; Li, C.; Asami, K.; Fujimoto, K. Influence of rare-earth elements (La, Ce, Nd and Pr) n the performance of Cu/Zn/Zr catalyst for CH3OH synthesis from CO2. Catal. Commun. 2014, 4, 50–54. [Google Scholar] [CrossRef]
  114. Wang, G.N.; Chen, L.M.; Sun, Y.H.; Wu, J.L.; Fu, M.L.; Ye, D.Q. Carbon dioxide ydrogenation to methanol over Cu/ZrO2/CNTs: Effect of carbon surface chemistry. RSC Adv. 2015, 5, 45320–45330. [Google Scholar] [CrossRef]
  115. Zhang, Y.H.; He, Y.M.; Cao, M.Y.; Liu, B.; Li, J.L. High selective methanol synthesis from CO2 hydrogenation over Mo-Co-C-N catalyst. Fuel 2022, 325, 124854. [Google Scholar] [CrossRef]
  116. Song, I.; Park, C.; Choi, H.C. Synthesis and properties of molybdenum disulphide: From bulk o atomic layers. RSC Adv. 2015, 5, 7495–7514. [Google Scholar] [CrossRef] [Green Version]
  117. Zhou, S.H.; Zeng, H.C. Boxlike Assemblages of Few-Layer MoS2 Nanosheets with Edge Blockage for High-Efficiency Hydrogenation of CO2 to Methanol. ACS Catal. 2022, 12, 9873–9886. [Google Scholar] [CrossRef]
  118. Wang, J.J.; Meeprasert, J.; Han, Z.; Wang, H.; Feng, Z.D.; Tang, C.Z.; Sha, F.; Tang, S.; Li, G.N.; 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]
  119. Li, Y.; Chan, S.H.; Sun, Q. Heterogeneous catalytic conversion of CO2: A comprehensive theoretical review. Nanoscale 2015, 7, 8663–8683. [Google Scholar] [CrossRef]
  120. Yang, Y.; Evans, J.; Rodriguez, J.A.; White, M.G.; Liu, P. Fundamental studies of methanol synthesis from CO2 hydrogenation on Cu(111), Cu clusters, and Cu/ZnO(0001). Phys. Chem. Chem. Phys. 2010, 12, 9909–9917. [Google Scholar] [CrossRef]
  121. Du, J.; Zong, L.; Zhang, S.; Gao, Y.; Dou, L.; Pan, J.; Shao, T. Numerical investigation on the heterogeneous pulsed dielectric barrier discharge plasma catalysis for CO2 hydrogenation at atmospheric pressure: Effects of Ni and Cu catalysts on the selectivity conversions to CH4 and CH3OH. Plasma Process. Polym. 2021, 19, 2100111. [Google Scholar] [CrossRef]
  122. Hu, Z.-M.; Takahashi, K.; Nakatsuji, H. Mechanism of the hydrogenation of CO2 to methanol on a Cu(100) surface-dipped adcluster model study. Surf. Sci. 1999, 442, 90–106. [Google Scholar] [CrossRef]
  123. Dang, S.; Yang, H.; Gao, P.; Wang, H.; Li, X.; Wei, W.; Sun, Y. A review of research progress on heterogeneous catalysts for methanol synthesis from carbon dioxide hydrogenation. Catal. Today 2019, 330, 61–75. [Google Scholar] [CrossRef]
  124. 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]
  125. Wang, Y.; Yu, M.; Zhang, X.; Gao, Y.; Liu, J.; Zhang, X.; Gong, C.; Cao, X.; Ju, Z.; Peng, Y. Density Functional Theory Study of CO2 Hydrogenation on Transition-Metal-Doped Cu(211) Surfaces. Molecules 2023, 28, 2852. [Google Scholar] [CrossRef] [PubMed]
  126. Li, M.; Hua, B.; Wang, L.-C.; Zhou, Z.; Stowers, K.J.; Ding, D. Discovery of single-atom alloy catalysts for CO2-to-methanol reaction by density functional theory calculations. Catal. Today 2022, 388–389, 403–409. [Google Scholar] [CrossRef]
  127. Liu, L.; Wang, X.; Lu, S.; Li, J.; Zhang, H.; Su, X.; Xue, F.; Cao, B.; Fang, T. Mechanism of CO2 hydrogenation over a Zr1–Cu single-atom catalyst. New J. Chem. 2022, 46, 5043–5051. [Google Scholar] [CrossRef]
  128. Xie, G.; Jin, R.; Ren, P.; Fang, Y.; Zhang, R. Boosting CO2 hydrogenation to methanol by adding trace amount of Au into Cu/ZnO catalysts. Appl. Catal. B Environ. 2023, 324, 122233. [Google Scholar] [CrossRef]
  129. 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]
  130. Chen, Q.; Chen, X.; Ke, Q. Mechanism of CO2 hydrogenation to methanol on the W-doped Rh(111) surface unveiled by first-principles calculation. Colloids Surf. A Physicochem. Eng. Asp. 2022, 638, 128332. [Google Scholar] [CrossRef]
  131. Feng, Z.; Tang, C.; Zhang, P.; Li, K.; Li, G.; Wang, J.; Feng, Z.; Li, C. Asymmetric Sites on the ZnZrOx Catalyst for Promoting Formate Formation and Transformation in CO2 Hydrogenation. J. Am. Chem. Soc. 2023, 145, 12663–12672. [Google Scholar] [CrossRef] [PubMed]
  132. Rasteiro, L.F.; Rossi, M.A.L.S.; Assaf, J.M.; Assaf, E.M. Low-pressure hydrogenation of CO2 to methanol over Ni-Ga alloys synthesized by a surfactant-assisted co-precipitation method and a proposed mechanism by DRIFTS analysis. Catal. Today 2021, 381, 261–271. [Google Scholar] [CrossRef]
  133. Yang, Y.; White, M.G.; Liu, P. Theoretical Study of Methanol Synthesis from CO2 Hydrogenation on Metal-Doped Cu(111) Surfaces. J. Phys. Chem. C 2011, 116, 248–256. [Google Scholar] [CrossRef]
  134. Liu, J.; Ke, Q.; Chen, X. First-principles investigation of methanol synthesis from CO2 hydrogenation on Cu@Pd core–shell surface. J. Mater. Sci. 2020, 56, 3790–3803. [Google Scholar] [CrossRef]
  135. Ou, Z.; Ran, J.; Niu, J.; Qin, C.; He, W.; Yang, L. A density functional theory study of CO2 hydrogenation to methanol over Pd/TiO2 catalyst: The role of interfacial site. Int. J. Hydrogen Energy 2020, 45, 6328–6340. [Google Scholar] [CrossRef]
  136. 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]
  137. 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]
  138. Yang, Y.; Mims, C.A.; Disselkamp, R.S.; Kwak, J.H.; Peden, C.H.F.; Campbell, C.T. (Non) formation of methanol by direct hydrogenation of formate on copper catalysts. J. Phys. Chem. C 2010, 114, 17205–17211. [Google Scholar] [CrossRef]
Figure 1. CO2 activation mechanism on Ga-modified Cu-based catalysts [69].
Figure 1. CO2 activation mechanism on Ga-modified Cu-based catalysts [69].
Atmosphere 14 01208 g001
Figure 2. Schematic representation of our experimental onepot procedure [75].
Figure 2. Schematic representation of our experimental onepot procedure [75].
Atmosphere 14 01208 g002
Figure 3. Illustration of synthesis of the hollow structured Cu@ZrO2 derived from Cu-MOF808 (pyrolysis temperature was 300 °C) [52].
Figure 3. Illustration of synthesis of the hollow structured Cu@ZrO2 derived from Cu-MOF808 (pyrolysis temperature was 300 °C) [52].
Atmosphere 14 01208 g003
Figure 4. CO2 hydrogenation activity of In2O3 and Ru/In2O3. (a) CO2 conversion and methanol selectivity, (b) methanol STY, (c) apparent activation energy of CO2 conversion, and (d) methanol formation rate versus time of the Ru/In2O3 catalyst for 40 h on stream [87].
Figure 4. CO2 hydrogenation activity of In2O3 and Ru/In2O3. (a) CO2 conversion and methanol selectivity, (b) methanol STY, (c) apparent activation energy of CO2 conversion, and (d) methanol formation rate versus time of the Ru/In2O3 catalyst for 40 h on stream [87].
Atmosphere 14 01208 g004
Figure 5. Evolution of the methanol STY with time on stream (TOS) over In13/ZrO2 and In13/CeO2 [19].
Figure 5. Evolution of the methanol STY with time on stream (TOS) over In13/ZrO2 and In13/CeO2 [19].
Atmosphere 14 01208 g005
Figure 6. Catalytic stability on-stream test of the continuous catalytic process over CuZnO-MOF-74–350 with VH2/CO2 = 3/1, P = 4.0 MPa, T = 190 °C, and WHSV = 4000 mL·gcat−1·h−1 [93].
Figure 6. Catalytic stability on-stream test of the continuous catalytic process over CuZnO-MOF-74–350 with VH2/CO2 = 3/1, P = 4.0 MPa, T = 190 °C, and WHSV = 4000 mL·gcat−1·h−1 [93].
Atmosphere 14 01208 g006
Figure 7. Proposed reaction mechanism for CO2 hydrogenation to methanol over 5% GaZnZrOx catalyst [107].
Figure 7. Proposed reaction mechanism for CO2 hydrogenation to methanol over 5% GaZnZrOx catalyst [107].
Atmosphere 14 01208 g007
Figure 8. (a) Cu/CeZrOX-M prepared via the new microimpinging stream reactor coupled with ultrasound, (b) 960 Cu/CeZrOX-B catalysts prepared via the conventional batch method [111].
Figure 8. (a) Cu/CeZrOX-M prepared via the new microimpinging stream reactor coupled with ultrasound, (b) 960 Cu/CeZrOX-B catalysts prepared via the conventional batch method [111].
Atmosphere 14 01208 g008
Figure 9. Potential energy profiles of the HCOO pathway in the process of methanol synthesis from CO2 hydrogenation on the Zr1–Cu surface [128].
Figure 9. Potential energy profiles of the HCOO pathway in the process of methanol synthesis from CO2 hydrogenation on the Zr1–Cu surface [128].
Atmosphere 14 01208 g009
Figure 10. Gibbs free energy profile for CO2 hydrogenation to (a) HCOOH, (b) H2CO, and (c) H3COH over CuII-MOF-808, respectively [31].
Figure 10. Gibbs free energy profile for CO2 hydrogenation to (a) HCOOH, (b) H2CO, and (c) H3COH over CuII-MOF-808, respectively [31].
Atmosphere 14 01208 g010
Figure 11. (a) Reaction networks for CO2 hydrogenation to CH3OH on the Cu@Pd core–shell surface. (b) Potential energy profiles of the HCOO pathway, the COOH pathway, and the RWGS + CO-Hydro pathway [134].
Figure 11. (a) Reaction networks for CO2 hydrogenation to CH3OH on the Cu@Pd core–shell surface. (b) Potential energy profiles of the HCOO pathway, the COOH pathway, and the RWGS + CO-Hydro pathway [134].
Atmosphere 14 01208 g011
Figure 12. Potential energy surfaces for CO2 hydrogenation to methanol on Cu (1 1 1) via the formate and hydrocarboxyl mechanisms [137].
Figure 12. Potential energy surfaces for CO2 hydrogenation to methanol on Cu (1 1 1) via the formate and hydrocarboxyl mechanisms [137].
Atmosphere 14 01208 g012
Table 1. Summary of Cu-based catalysts for the CO2 hydrogenation to methanol.
Table 1. Summary of Cu-based catalysts for the CO2 hydrogenation to methanol.
Catalyst T (°C) P (MPa) Space
Velocity a
H2/CO2
Ratio
Conv.
(%)
Sel.
(%)
STY c Ref.
CuO/ZnO-400 260 3 (G) 12,000 3 18.5 41.7 0.3 [24]
CZA-r@CZM 240 3 (G) 32,000 3 11.7 73.0 0.730 [25]
CZZ@Al-TUD-1 280 2 (G) 10,000 3 10 20.3 0.180 [43]
10NG-CZA 200 3 N.A. 3 8.2 84 N.A. [44]
aCZZ-LDH 250 3 (G) 2000 3 4.9 78.3 0.037 [45]
CuZnO@rGO 260 1 20,000 b 3 N.A. ~100 0.007 [46]
15Cu/6.4ZnO/AlFum
MOF
230 5 (G) 10,000 3 45.6 3.48 0.057 [47]
CuZn10Zr 220 3 (W) 6000 3 14.0 54.4 0.65 [48]
30Cu-ZZ66/34 280 5 (G) 25,000 N.A. 19.6 50 0.725 [49]
CuZnGa 270 3 N.A. 3 15.9 N.A. 0.136 [50]
Cu5Sr3Ti2OX 220 2 (G) 12,000 3 1.7 62.8 N.A. [51]
CM-300 220 3 (W) 15,600 3 5.0 85 0.144 [52]
Cu0.3Ce0.3Zr0.7 240 3 N.A. 3 4.1 55.3 0.192 [53]
Cu3/LTH 250 3 (G) 3000 3 23.0 73.6 0.145 [54]
3Mg-C-ZZ SSC 320 3 (G) 2000 3 12.9 81.5 N.A. [55]
a (W) = WHSV, mL·gcat−1·h−1, (G) = GHSV, h−1; b LCO2·gcat−1·h−1; c STY (gMeOH·gcat−1·h−1); N.A.: Not available.
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

Xu, L.; Chen, X.; Deng, C.; Hu, K.; Gao, R.; Zhang, L.; Wang, L.; Zhang, C. Hydrogenation of Carbon Dioxide to Methanol over Non-Noble Catalysts: A State-of-the-Art Review. Atmosphere 2023, 14, 1208. https://doi.org/10.3390/atmos14081208

AMA Style

Xu L, Chen X, Deng C, Hu K, Gao R, Zhang L, Wang L, Zhang C. Hydrogenation of Carbon Dioxide to Methanol over Non-Noble Catalysts: A State-of-the-Art Review. Atmosphere. 2023; 14(8):1208. https://doi.org/10.3390/atmos14081208

Chicago/Turabian Style

Xu, Lujing, Xixi Chen, Chao Deng, Kehao Hu, Ruxing Gao, Leiyu Zhang, Lei Wang, and Chundong Zhang. 2023. "Hydrogenation of Carbon Dioxide to Methanol over Non-Noble Catalysts: A State-of-the-Art Review" Atmosphere 14, no. 8: 1208. https://doi.org/10.3390/atmos14081208

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