Startseite Low-temperature hydrogenation of CO2 to methanol: progress in constructing catalytic active sites
Artikel Open Access

Low-temperature hydrogenation of CO2 to methanol: progress in constructing catalytic active sites

  • Jiadong Wang , Mingliang Ma , Qingxiang Ma und Tian-Sheng Zhao EMAIL logo
Veröffentlicht/Copyright: 5. Juni 2025
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

In the move away from fossil fuels, methanol is viewed as a viable alternative engine fuel and hydrogen carrier, while serving as a key component in the chemical industry. In responding to the environmental impact caused by excessive CO2 emissions and the drive to promote green H2 technology, CO2 hydrogenation to produce methanol has received increasing attention, which has addressed H2 storage and transport. The reaction is thermodynamically feasible at low temperatures with a high conversion in a single pass, but there is significant scope to enhance reaction kinetics. This article reviews the progress that has been made in engineering catalytic active sites for the heterogeneous hydrogenation of CO2 to methanol at temperatures below 200 °C, including considerations of alloying effects, doping, defect formation, active site size and dispersion, coordination effects, and surface modification. The active site valence state and size in tandem with electron transfer, hydrogen spillover, and surface alkalinity/hydrophobicity can affect catalytic performance and the prevailing reaction pathway to varying degrees. The low-temperature reaction process is briefly discussed and future research directions required to further enhance catalytic efficiency are proposed.

Abbreviations

AC

activated carbon

CNFs

carbon nanofibers

CNTs

carbon nanotubes

DFT

density functional theory

DRIFTS

diffuse reflectance infrared Fourier transform spectroscopy

EMSI

electronic metal-support interaction

ESR

electron spin resonance

EXAFS

extended X-ray absorption fine structure

FL-MoS2

few-layer molybdenum disulfide

NPs

nanoparticles

NSs

nanosheets

NWs

nanowires

Ov

oxygen vacancies

RWGS

reverse water-gas shift

SACs

single atom catalysts

SMSI

strong metal-support interaction

STY

space-time yield

Sv

sulfur vacancies

TOF

turnover frequency

TPR

temperature-programmed reduction

XANES

X-ray absorption near edge structure

XPS

X-ray photoelectron spectroscopy

1 Introduction

Methanol, which is easy to store, transport, and distribute, is a fundamental chemical feedstock that can also be used as an engine fuel and H2 carrier (Schieweck et al. 2020). The global production of methanol reached 107 million tons in 2021 (Sen et al. 2022). Given the urgency of ameliorating the negative environmental impact of CO2 emissions and the growing importance of green H2 technologies, CO2 hydrogenation to methanol (reaction 1) has been the subject of appreciable research and development in tandem with the accompanying reverse water–gas shift (RWGS, reaction 2). Implementing the reaction at low temperatures disfavors the formation of CO while favoring methanol production from a thermodynamic perspective.

(1) CO 2 + 3 H 2 CH 3 OH + H 2 O H 298 K = 49.47  kJ / mol
(2) CO 2 + H 2  CO + H 2 O H 298 K = + 41.17  kJ / mol

A significant number of studies have examined this reaction at temperatures in the 220−350 °C range (Din et al. 2019a; Jiang et al. 2020; Wu et al. 2022a). The reported conversion of the feed-gas (CO2 + H2) was low and, in most cases, significant CO was generated. Operation at higher reaction temperatures results in greater energy consumption, representing unsustainability in the drive to ensure decarbonization where energy-intensive processes are undesirable.

The performance of a range of metal/non-metal catalysts in the low-temperature hydrogenation of CO2 to methanol is summarized in Table 1. The strategies employed in constructing the active sites involve the following.

Table 1:

Catalyst performance metrics in the low-temperature hydrogenation of CO2 to methanol.

Catalysts Phases Temperature (°C) Pressure (bar) CO2 conv. (%) Methanol (sel./%) STY (mmol/g/h) TOF (h) Strategies References
Pt3Co octapods/C Gas–liquid–solid 150 32 100 173 758(Pt) Alloying effects Khan et al. (2016)
Rh75W25 nanosheets/C Gas–liquid–solid 150 32 97 107 592(Rh) Zhang et al. (2017a)
RhCo porous nanospheres/C Gas–liquid–solid 150 32 96.3 55.75 612(Rh) Zheng et al. (2018)
Pt4Co NWs/C Gas–liquid–solid 150 32 100 81.4 1773(Pt) Bai et al. (2017)
Cu93Au7/ZnO Gas–liquid–solid 200 50 3.4 100 Mosrati et al. (2023)
110 °C–Ni5Ga3 Gas–solid 200 30 3.2 100 972 Men et al. (2020)
Co4N nanosheets Gas–liquid–solid 150 32 89.4 70 25.6 Doping Wang et al. (2017)
Co–N/C Gas–liquid–solid 180 20 99 105.3 47.8 Li et al. (2024)
Er0.2CuZnO Gas–solid 170 50 8.5 89.8 1.42 Huang et al. (2024)
1Pd–10Cu/CeO2 Gas–solid 190 30 8.8 44.3 0.80 Choi et al. (2017)
h-PdMo/Mo2N Gas–solid 25 9 0.004 100 0.007 0.15 Sugiyama et al. (2023)
Cu/Mo2C Gas–liquid–solid 135 40 ∼1 93 0.76 Chen et al. (2016)
Pd/Mo2C Gas–liquid–solid 135 40 ∼1 95 0.83 Chen et al. (2016)
P–Cu/ZnO/Al2O3/MgO-R Gas–solid 190 30 9.4 78 8.18 Wang et al. (2023a)
Cu/ZrO2/CNFs/ZnO Gas–liquid–solid 180 30 9 92 1.40 6.73 Din et al. (2019b)
Cu/ZrO2/CNFs-Nb0.8 Gas–liquid–solid 180 30 9 87 1.25 3.20 Din et al. (2016)
15 Cu–ZrO2/CNFs Gas–liquid–solid 180 30 10 67 0.78 3.35 Din et al. (2017)
Pt(3)/MoOx(30)/TiO2 Gas–liquid–solid 150 60 ∼81 ∼90 Defect formation Toyao et al. (2019)
Cu/ZrO2(20CZ–OG) Gas–solid 200 40 1.5 ∼90 5.62 Chang et al. (2023)
FL-MoS2 Gas–solid 180 50 12.5 94.3 4.21 17.9(Mo) Hu et al. (2021)
MoS2-NS-60 Gas–solid 200 30 4.8 96 17.82 170(Sv) Zhang et al. (2024a)
Re(1)/TiO2 Gas–liquid–solid 150 60 82 0.1 1.83 Active site size/dispersion Ting et al. (2019)
Re/TiO2 Gas–solid 150 56 4 65 1.02 Phongprueksathat et al. (2023)
Au(1.6)/ZrO2 Gas–liquid–solid 180 40 75 1.02 20 Wu et al. (2017)
Cu/ZnO/Al2O3/ZrO2-RHT-9 Gas–solid 190 50 10.7 81.8 2.72 26.4 Xiao et al. (2017)
Cu–TiO2HT Gas–solid 200 30 9.4 96 13 0.84 Sharma et al. (2021)
CuZnO–MOF-74-350 Gas–solid 190 40 7.5 78.2 2.50 Han et al. (2022)
Cu/ZrO2(CAZ-1) Gas–solid 180 30 1.61 100 0.22 1.37 Zhao et al. (2022)
Cu–N4 Gas–liquid–solid 150 32 95.5 4.2 262.5 Coordination effects Yang et al. (2021)
Pt1@MIL Gas–liquid–solid 150 32 90.3 1.2 117 Chen et al. (2019)
7.5 % Pt/MoS2 Gas–liquid–solid 150 32 81.3 50 162.5 Li et al. (2018)
Cu/Zn-2 mM Gas–solid 200 70 ∼5 78.8 2.72 Surface modification Wang et al. (2024)
SiC QDs Gas–liquid–solid 150 32 100 169.5 0.27 mmol m−2 h−1 Peng et al. (2018)
Cu1/ZnO Gas–solid 170 30 1.8 99.1 Wu et al. (2022b)
Cu/ZnO Gas–liquid–solid 180 60 0.26 Reaction process Nieminen et al. (2018)
Cu–Cr–O Gas–liquid–solid 200 30 15 80 Fan et al. (1999)
Cu/ZnO + 3A Gas–liquid-–solid 150 50 47 91.16 Boonamnuay et al. (2021)
Cu/Mo2C + Cu–Cr Gas–liquid–solid 135 40 77 0.32 1.7 Chen et al. (2015)

Alloying effects. An alloying of metals can generate a high density of negatively charged sites that promote CO2 activation. Control over particle shape and grain boundaries serves to increase the accessibility of the active sites to the reactants, promoting hydrogenation.

Doping. The incorporation of elemental dopants provides more H adsorption sites, lowering the reaction activation energy, and promoting electron transfer and hydrogen spillover that facilitates CO2 conversion. Moreover, doping can increase active site dispersion and surface basicity.

Defect formation. Oxygen vacancies (Ov) promote electron transfer to CO2 or induce strong metal−support interactions, facilitating CO2 and H2 activation. Sulfur vacancies (Sv) can regulate C–O bond dissociation and H2 activation, enhancing catalytic performance.

Active site size/dispersion. Varying the active site size can increase activity where enhanced dispersion prevents the aggregation of active metals, improving catalyst stability.

Coordination effects. The chemical characteristics and quantity of coordinating atoms can be used to adjust active site valence state distribution, which influences the reaction pathway, lowering the activation energy and enhancing catalytic performance.

Surface hydrophobic/hydrophilic modification. Hydrophobicity suppresses the aggregation and oxidation of active species, improving stability. Hydrophilic modification involves the introduction of surface –OH groups that promote hydrogenation via hydrogen spillover.

The use of supported alloy catalysts and SiC quantum dots (QDs) catalysts has delivered high selectivity for methanol and methanol space-time yield (STY).

Different reaction processes have been adopted to promote the low-temperature hydrogenation of CO2. In the alcohol solvent-assisted process, the esterification of formic acid has been employed to lower the overall reaction energy barrier in the application of cascade catalysis to convert CO2 and H2 to methanol.

2 Alloying effects

2.1 Formation of active sites with a high negative charge density

Electron transfer is critical in CO2 activation, forming a surface carboxylate species (CO2δ−) (Graciani et al. 2014; Zheng et al. 2018). Studies have shown that alloying facilitates electron transfer from metals with a lower electronegativity to metals with a higher electronegativity, generating active metals with a high negative charge density that promote CO2 activation. Khan et al. (2016) have reported Pt–Co alloy nanocrystal preparation using a chemical solvothermal method for application in the liquid-phase hydrogenation of CO2 to methanol. The resultant Pt3Co octopods delivered a turnover frequency (TOF) of 758 h−1 at 150 °C and 32 bar CO2/H2, which was 6.1 times higher than that recorded for Pt octopods. As shown in Figure 1, the photoelectron Pt 4f7/2 binding energy of the Pt3Co octopods was shifted by 0.9 eV (71.3 → 70.4 eV) when compared with Pt octopods as a result of electron transfer from Co to Pt, increasing the negative charge associated with the Pt sites. Infrared reflection absorption spectroscopy (IRRAS) and density functional theory (DFT) calculations have indicated that Pt atoms at the apex of the Pt3Co octopods bear a high negative charge density, which promotes the activation of CO2 to generate CO2δ−. The formation of some alloy phases serves to enhance methanol synthesis, notably the Ni5Ga3 alloy which was shown to increase methanol selectivity by inhibiting the RWGS reaction and circumventing CO as by-product (Studt et al. 2014; Men et al. 2020). The application of a colloidal Pd2Ga alloy in CO2 hydrogenation delivered a methanol STY that was 4 times higher than Cu/ZnO/Al2O3 (García-Trenco et al. 2017).

Figure 1: 
XPS spectra for Pt nanocrystals (Khan et al. 2016); reproduced with permission from John Wiley and Sons (copyright 2016).
Figure 1:

XPS spectra for Pt nanocrystals (Khan et al. 2016); reproduced with permission from John Wiley and Sons (copyright 2016).

A higher negative charge density facilitates transfer of electrons to the antibonding orbitals of adsorbed CO2 and facilitates the adsorption of the (HCOO*) intermediate for further hydrogenation to methanol. Zhang et al. (2017a) have reported the synthesis of a Rh75W25 alloy NSs/C by chemical solvothermal, which was used to promote the liquid-phase hydrogenation of CO2 at 150 °C and 32 bar CO2/H2. The resultant TOFRh (592 h−1) was 4 times higher than that obtained with Rh NSs/C, with an increased methanol selectivity (from 95 % to 97 %). The electron transfer from W to Rh resulted in a lower Rh 3d5/2 XPS binding energy for Rh75W25 NSs (Figure 2).

Figure 2: 
XPS spectra for Rh nanocrystals (Zhang et al. 2017a); reproduced with permission from the American Chemical Society (copyright 2017).
Figure 2:

XPS spectra for Rh nanocrystals (Zhang et al. 2017a); reproduced with permission from the American Chemical Society (copyright 2017).

The FTIR spectrum recorded following CO2 adsorption on Rh75W25 NSs at 150 °C is shown in Figure 3 where the absorption bands at 1,500−1,600 cm−1 are attributed to CO2δ− asymmetric stretching vibrations associated with CO2 activation.

Figure 3: 
In situ DRIFT spectra of CO2 adsorption on Rh nanocrystals at 150 °C (Zhang et al. 2017a); reproduced with permission from the American Chemical Society (copyright 2017).
Figure 3:

In situ DRIFT spectra of CO2 adsorption on Rh nanocrystals at 150 °C (Zhang et al. 2017a); reproduced with permission from the American Chemical Society (copyright 2017).

Contacting Rh75W25 NSs with CO2/H2 at 150 °C resulted in CO2δ− hydrogenation to the intermediate HCOO*, generating a higher associated FTIR peak intensity (at 2,961, 1,684, 1,453, and 1,047 cm−1) relative to the CO2δ− (1,500−1,600 cm−1) signals observed for Rh NSs (Figure 4). In addition, the C = O vibration peak (1,684 cm−1) exhibited a red-shift (to 1,693 cm−1) in the transition of CO2δ− to HCOO*, indicating a stronger adsorption of HCOO*. The negatively charged active site facilitates CO2 activation to generate CO2δ−, and enhances adsorption of the HCOO* intermediate, inhibiting the formation of formaldehyde as by-product, which enables selective hydrogenation to methanol.

Figure 4: 
In situ DRIFT spectra for Rh nanocrystals exposed to CO2/H2 at 150 °C (Zhang et al. 2017a); reproduced with permission from the American Chemical Society (copyright 2017).
Figure 4:

In situ DRIFT spectra for Rh nanocrystals exposed to CO2/H2 at 150 °C (Zhang et al. 2017a); reproduced with permission from the American Chemical Society (copyright 2017).

2.2 Increased exposure of active site

The metal-carrier interface is viewed as the catalytically active site (Han et al. 2022) where the characteristics of the interface determine the level of selective CO2 hydrogenation at low temperatures (Han et al. 2021). Alloys with grain boundaries present more accessible active sites to facilitate CO2 activation. Grain boundary-rich RhCo NSs/C synthesized using a chemical solvothermal procedure, and employed in the liquid-phase hydrogenation of CO2 at 150 °C and 32 bar CO2/H2 (Zheng et al. 2018), generated a TOFRh of 612 h−1, 6.1 times and 2.5 times higher than Rh/C and RhCo NPs/C, respectively, with an associated methanol selectivity of 96.3 %, attributed to the high-density accessible grain boundaries. In situ XPS characterization has confirmed a high negative charge associated with the surface Rh atoms due to Co→Rh electron transfer that promoted the activation of CO2 to generate CO2δ− and the intermediate H3CO*.

Catalysts with faceted surfaces, rich in step and edge sites, facilitate effective CO2 hydrogenation. A “zigzag” Pt4Co NWs/C reported by Bai et al. (2017), exhibiting a Pt-rich surface with abundant steps/edges was prepared using an ultrasound-hydrothermal method. In the liquid-phase hydrogenation of CO2 at 150 °C and 32 bar CO2/H2, the catalyst delivered a TOFPt of 1773 h−1 with 100 % methanol selectivity and a STY of 81.4 mmol/g/h. This level of activity was superior to that of Pt NWs/C (TOFPt = 515 h−1 and methanol STY = 28.7 mmol/g/h), and a commercial Pt/C (TOFPt = 17 h−1 and methanol STY = 17.3 mmol/g/h). The Pt 4f7/2 XPS binding energy for Pt4Co NWs/C (70.7 eV) was significantly lower (by 0.7 eV) than that (71.4 eV) recorded for Pt NWs/C (Figure 5), indicating an electron transfer from Co to Pt that increased the negative charge on the supported Pt which enhanced CO2 activation and subsequent hydrogenation (Khan et al. 2016).

Moreover, DRIFTS analysis has shown that the absorption peaks associated with the intermediate bidentate carbonate on Pt4Co NWs/C exhibited a red-shift relative to Pt NWs/C, indicating a stronger CO2 interaction with Pt atoms in Pt4Co NWs. The results confirm that the surface structure rich in steps/edges exposes an abundance of active sites which, combined with the alloying effect, promote CO2 activation.

Figure 5: 
XPS Pt spectra for Pt–Co NWs (Bai et al. 2017); reproduced with permission from John Wiley and Sons (copyright 2017).
Figure 5:

XPS Pt spectra for Pt–Co NWs (Bai et al. 2017); reproduced with permission from John Wiley and Sons (copyright 2017).

The application of Cu-based catalysts in CO2 hydrogenation has been the subject of a number of studies (Chen et al. 2016; Chang et al. 2023; Huang et al. 2024). The combination of Cu and Au in the formation of alloy catalysts has improved performance in the low-temperature hydrogenation of CO2 to methanol. Zinc oxide supported CuAu was employed to promote CO2 liquid-phase hydrogenation at 200 °C and 50 bar CO2/H2 (Mosrati et al. 2023). When the Cu93Au7 alloy cluster loading was increased from 1 wt% to 10 wt%, the CO2 conversion increased from 0.1 % to 3.4 %, but methanol selectivity was 100 % regardless of loading. The dependence of CO2 conversion on loading demonstrates the low-temperature catalytic hydrogenation capability of the surface alloy sites. In contrast, Au/ZnO did not exhibit any activity. Commercial Cu/ZnO/Al2O3 with a Cu loading of 50.6 % delivered a CO2 conversion of 3.5 % and lower methanol selectivity (88 %). When compared with Cu/ZnO/Al2O3, Cu93Au7/ZnO (with 10 wt% Cu loading) exhibited a five-fold higher intrinsic activity for methanol synthesis with inhibited CO formation. The addition of Au inhibited Cu particle agglomeration (from 30−40 nm to 2.7−4.4 nm), facilitating the exposure of more active sites. Characterization by XPS has demonstrated that, during CO2 hydrogenation, metallic Cu was segregated in the outer shell of the nanoalloy whereas negatively charged Au represented the core of the nanoalloy. The copper segregation on the nanoalloy surface facilities better interaction with reactants and intermediates in promoting the reaction.

So far mere Ni5Ga3 non-precious metal alloy catalyst has been reported for low-temperature CO2 hydrogenation to methanol. The preparation of Ni5Ga3 requires high-temperature reduction in H2, which results in metal particle aggregation or surface reconstruction, thereby diminishing the active sites. By comparison, the preparation of noble metal alloys needs only mild reduction temperatures. To lower the reduction temperature, strategic modifications could involve tailoring the local atomic environment of Ni5Ga3 (e.g., by engineering coordination geometries, optimizing hydrogen spillover dynamics at interfaces, or incorporating noble metal species). Another approach is re-engineering the synthesis protocol to avoid the high-temperature reduction procedure, using sodium borohydride wet-chemical reduction. Furthermore, it is worth noting that the low-temperature CO2/H2 activation capability of non-precious metal alloys is relatively low, limiting their applications. Nevertheless, non-precious metal alloy catalysts are cheaper for industrial use, worthing further development.

Conventional co-precipitation, impregnation, and sol-gel methods cannot make metal salts completely into alloy phases. Alloy catalysts are currently synthesized using chemical solvothermal methods, which necessitate solvents, surfactants, and linking agents. Merely precious metal-containing alloys (Pt–Co, Rh–Co, and Rh–W) have shown promising low-temperature catalytic activity correlating to electron transfer between the metals, as demonstrated by XPS and Bader charge analysis. Current mechanistic characterization relies mainly on in situ DRIFTS and DFT calculations. Discrepancy sometimes arises because in situ spectroscopic conditions do not reach the actual reaction circumstances due to hardware limitations (e.g., inadequate pressure), rendering the analyses approximate rather than definitive. It is expected that the application of in situ EPR and operando infrared spectroscopy monitoring (steady-state versus transient analysis) (Gau et al. 2021) could provide valuable characterization information, but such exploration is lacking.

3 Use of dopants

3.1 Non-metal doping

Doping with nitrogen can introduce H adsorption sites and lower the reaction activation energy, promoting CO2 hydrogenation. N-doped Co4N NSs were prepared by ammonia pyrolysis (Wang et al. 2017) and used in the liquid-phase hydrogenation of CO2 to methanol at 150 °C and 32 bar CO2/H2. The resultant TOF (25.6 h−1) was 64 times higher than that achieved with Co NS (0.4 h−1) and 2.2 times higher than Cu/ZnO/Al2O3 (11.5 h−1). This response was attributed to the lower hydrogenation activation energy due to N doping (43.3 kJ/mol for Co4N NSs compared with 91.4 kJ/mol for Co NSs). An in situ DRIFTS analysis of Co4N NSs in a CO2 atmosphere (Figure 6a) exhibits peaks due to CO2δ− stretching vibrations (at 1,489 and 1,346 cm−1) when the temperature was raised to 120 °C, representing the lowest temperature for CO2 activation on Co4N NSs. The in situ DRIFTS recorded at 40 °C (Figure 6b) reveals stretching vibration peaks for CO2δ− and N–H (3,140 cm−1). An increase in temperature to 60 °C was accompanied by the appearance of C–H (2,955 cm−1) and O–C–O (1,626 cm−1) vibration peaks due to HCOO*. At 120 °C, the intensity of these peaks decreases, and peaks associated with the CH2O* species appear. The results suggest that Co4NHx is beneficial for the activation and transformation of CO2 at lower temperatures than observed for Co4N NSs. The mechanistic analysis suggests that H2 is dissociated on the Co, generating H atoms that are adsorbed on N sites associated with Co4N NSs, which restructs to form the active Co4NHx phase, where the H of the amino group interacts directly with CO2 to form the intermediate HCOO*. Moreover, the adsorbed H2O* activates the H of the amino group through hydrogen-bonding interactions, promoting hydrogenation.

Figure 6: 
In situ DRIFTS of Co4N NSs in (a) CO2 (b) CO2/H2 atmospheres (Wang et al. 2017); reproduced with permission from Springer Nature (copyright 2017).
Figure 6:

In situ DRIFTS of Co4N NSs in (a) CO2 (b) CO2/H2 atmospheres (Wang et al. 2017); reproduced with permission from Springer Nature (copyright 2017).

Li et al. (2024) have reported the preparation of Co–N/C catalysts rich in pyridine N–Co bonds by ammoniation of a ZIF-67 precursor. A TOF of 47.8 h−1 (methanol selectivity of 99 %) was recorded for the liquid-phase hydrogenation of CO2 at 180 °C and 20 bar CO2/H2, which was 7 times higher than that achieved with non-ammoniated Co/C (mainly containing pyrrole-type Co-Nx sites). After five reaction cycles, the activity of Co–N/C exceeded 96 % of the original value. This stable performance is attributed to a significant reduction of electron localization at the Co center (transferring electrons from Co to adjacent N atoms) as a result of the incorporation of pyridine N. The methane side-product was inhibited. Nitrogen doping served to enhance CO2 and H2 adsorption on the Co–N/C surface, lowering the energy barrier for methanol formation (93.7 kJ/mol on Co–N/C versus 106.6 kJ/mol on Co/C). Hydrogen is dissociatively adsorbed on the Co–N/C surface and interacts with N atoms to form amino groups. Theoretical and experimental studies have shown that pyridine-N has a higher affinity for bonding with Co when compared with pyrrole-N and graphitized-N. Pyridine-N–Co serves as the active site during CO2 hydrogenation. This study has established the critical role of the N configuration in determining CO2/H2 adsorption and activation.

Moreover, N doping enhances catalyst basicity for CO2 adsorption (Donphai et al. 2023). The introduction of amino groups on the surface of activated carbon (AC) increased the CO2 adsorption capacity with an increase in the surface N content to 1.38 % (Zhang et al. 2013). At 25 °C and 36 bar, CO2 adsorption on AC-NH2 and AC was 19.07 mmol/g and 16.67 mmol/g, respectively, indicating an effective adsorption on the basic N sites. An increase in temperature resulted in a lower CO2 adsorption capacity that was more pronounced for AC relative to AC-NH2. The latter exhibited CO2 uptake at room temperature (25 °C), establishing a strong affinity of AC-NH2 for CO2 adsorption that enables low-temperature reaction. The doped N species can improve the dispersion of active metals, where Cu–ZrO2/CNTs–N prepared by deposition-precipitation (Sun et al. 2018) exhibited strong interaction between pyridine-N and Cu. The addition of N promoted the dispersion and reduction of CuO, decreasing the grain size of Cu, and enhancing dissociative adsorption of H2. The results have demonstrated that pyridine-N contributes to effective CO2 adsorption and activation, promoting methanol formation.

3.2 Metal doping

3.2.1 Electron transfer promotion

Active sites with a high negative charge density are effective in the low-temperature adsorption activation of CO2 (Khan et al. 2016; Zhang et al. 2017a). Adding promoters (Mg, Ga, La, etc.) facilitates electron transfer to the active sites (Xie et al. 2021; Cored et al. 2022; Guo et al. 2022). Furthermore, the strong electronic metal-support interaction (EMSI) between the metal and support enables electrons to transfer from the support to the active metal, related closely to the Fermi level (Frost 1988). Strong EMSI can alter the d-band center of transition metals, increasing the local electron density at the active interface (Jackson et al. 2017) which affects catalytic activity. Huang et al. (2024) prepared Er0.2CuZnO by co-precipitation and achieved 8.5 % CO2 conversion with 89.8 % selectivity to methanol for reaction at 170 °C and 50 bar, superior to the conversion (5.6 %) and selectivity (84.7 %) recorded for CuZnO. The improved performance was attributed to a decrease in ZnO particle size due to Er doping that serves to increase the Cu–ZnO interface with EMSI between Cu and ZnO. Doping Er promotes the transfer of electrons from ZnO to Cu to generate Cuδ−. As shown in Figure 7, the Cu 2p3/2 binding energy in Er0.2CuZnO is lower (by 0.19 eV) relative to CuZnO, indicating a greater electron charge density associated with Cu NPs in Er0.2CuZnO. In situ DRIFTS has demonstrated that Er doping promotes CO2 adsorption activation and enhances the activation of surface carbonate species and hydrogenation of the intermediate CO* to HCO*.

Figure 7: 
Copper XPS spectra for ErCuZnO and CuZnO (Huang et al. 2024); reproduced with permission from the American Chemical Society (copyright 2024). Selectivity (44.3 %), and methanol STY (0.80 mmol/g/h) on 1Pd–10Cu/CeO2, were higher than the values (1.8 %, 84 %, and 0.31 mmol/g/h, respectively) recorded for Cu/CeO2. This response was ascribed to the interaction between highly dispersed Pd and Cu that inhibited the migration and encapsulation of the active site by reduced CeO2, resulting in a higher Cu concentration on the 1Pd–Cu/CeO2 surface. The Pd component promotes the reduction of Cu species and surface CeO2 by transferring electrons to Cu and CeO2, generating more active Cu species. In addition, the reduction of surface CeO2 (Ce4+→Ce3+) promotes the generation of Ov, enhancing the catalytic activation of CO2.
Figure 7:

Copper XPS spectra for ErCuZnO and CuZnO (Huang et al. 2024); reproduced with permission from the American Chemical Society (copyright 2024). Selectivity (44.3 %), and methanol STY (0.80 mmol/g/h) on 1Pd–10Cu/CeO2, were higher than the values (1.8 %, 84 %, and 0.31 mmol/g/h, respectively) recorded for Cu/CeO2. This response was ascribed to the interaction between highly dispersed Pd and Cu that inhibited the migration and encapsulation of the active site by reduced CeO2, resulting in a higher Cu concentration on the 1Pd–Cu/CeO2 surface. The Pd component promotes the reduction of Cu species and surface CeO2 by transferring electrons to Cu and CeO2, generating more active Cu species. In addition, the reduction of surface CeO2 (Ce4+→Ce3+) promotes the generation of Ov, enhancing the catalytic activation of CO2.

The use of metal dopants to promote the formation of active sites is an effective method to enhance catalytic performance. The CeOx/Cu (111) surface can stabilize the intermediate COOH*, promoting dissociation to generate CO*, and intermediate HCO* hydrogenation to methanol (Graciani et al. 2014). Palladium-doped Cu/CeO2 increased CO2 conversion and methanol yield (Choi et al. 2017). At 190 °C and 30 bar CO2/H2, the CO2 conversion (8.8 %), methanol selectivity (44.3 %), and methanol STY (0.80 mmol/g/h) on 1Pd-10Cu/CeO2, were higher than the values (1.8 %, 84 %, and 0.31 mmol/g/h, respectively) recorded for Cu/CeO2. This response was ascribed to the interaction between highly dispersed Pd and Cu that inhibited the migration and encapsulation of the active site by reduced CeO2, resulting in a higher Cu concentration on the 1Pd-Cu/CeO2 surface. The Pd component promotes the reduction of Cu species and surface CeO2 by transferring electrons to Cu and CeO2, generating more active Cu species. In addition, the reduction of surface CeO2 (Ce4+ → Ce3+) promotes the generation of Ov, enhancing the catalytic activation of CO2.

3.2.2 Lowering reaction activation energy and stabilizing active sites

Bulk, as opposed to surface, doping with metals can form intermetallic compounds, significantly lowering the apparent reaction activation energy and stabilizing the active phases, promoting the low-temperature hydrogenation of CO2 to methanol. Sugiyama et al. (2023) synthesized h-PdMo/Mo2N by ammonia pyrolysis, generating ordered alternating layers of Pd and Mo. Single-phase h-PdMo intermetallic NPs (h-PdMo/Mo2N) were loaded on Mo2N and compared with Pd NPs (Pd/Mo2N) loaded on Mo2N. The h-PdMo/Mo2N system exhibited good stability in air at the reaction atmosphere and superior catalytic performance to Pd/Mo2N. At reaction temperatures below 100 °C, methanol production over h-PdMo/Mo2N increased with increasing temperature, whereas trace methanol was generated over Pd/Mo2N at temperatures above 120 °C. This was attributed to a lower apparent activation energy for h-PdMo/Mo2N (27 kJ/mol) relative to Pd/Mo2N (78 kJ/mol). At 25 °C and 9 bar CO2/H2, the TOF recorded for h-PdMo/Mo2N was 0.15 h−1 with 100 % selectivity to methanol and 0.04 % CO2 conversion. This TOF exceeds that reported for a state-of-the-art Ir complex catalyst (0.02 h−1) (Kanega et al. 2021) and FL-MoS2 catalyst (0.09 h−1) (Hu et al. 2021) for reaction under comparable conditions. Moreover, at 100 °C, h-PdMo/Mo2N maintained a constant level of methanol production for 100 h, without any detectable alteration to the catalyst structure after reaction, indicating high thermal stability. Mechanistic studies have shown that CO2 is first decomposed to CO* on h-PdMo/Mo2N, followed by hydrogenation to form CH3O* and subsequently to methanol.

3.2.3 Promoting precursor reduction and hydrogen spillover

Different metal dopants significantly affect the TOF value in methanol synthesis and product distribution from CO2 hydrogenation. In the liquid-phase hydrogenation of CO2 over modified Mo2C at 135 °C and 40 bar CO2/H2 (Chen et al. 2016), the TOF decreased in order: Pd/Mo2C ≈ Cu/Mo2C > Fe/Mo2C > Co/Mo2C > Mo2C, with a methanol selectivity of 95 %, 93 %, 87 %, 84 %, and 79 %, respectively. The lower methanol selectivity in the case of Co/Mo2C and Fe/Mo2C was due to CH4 and C2–C4 hydrocarbon production. The addition of Cu and Pd promotes methanol formation whereas the incorporation of Co and Fe also facilitates C–C coupling, resulting in the formation of C2–C4 hydrocarbons. Analysis by H2-TPR has shown that the introduction of Cu shifted the Mo2C reduction peak to lower temperatures (210 °C → 170 °C). The introduction of Cu or Pd promoted the generation of H spillover. Mechanistic studies have suggested that methanol generation from CO2 proceeds via the formate pathway where the incorporation of metals modifies the properties of the active sites associated with the Mo2C surface, with the possible introduction of new active sites.

3.3 Metal oxide doping

3.3.1 Promoting H2 adsorption activation

Increasing catalytic H2 adsorption activation generates more active H atoms for CO2 hydrogenation, promoting the low-temperature reaction. Wang et al. (2023a) prepared a P-CZAM-R catalyst by co-precipitation, incorporating MgO into Cu/ZnO/Al2O3, which effectively inhibited Cu particle growth during the reduction treatment. The smaller Cu particle size promoted the formation of the active Cu–ZnO interface (Han et al. 2022), where the highly dispersed active metal exposed more H2 adsorption sites, increasing the supply of surface reactive H. H2-TPD measurements have shown that MgO doping enhanced the H2 adsorption capacity. At 190 °C and 30 bar CO2/H2, P-CZAM-R delivered a 9.4 % CO2 conversion for 120 h, with a methanol selectivity of 78 %.

Zinc oxide can promote the heterolysis of H2 to form Zn2+-H (Anderson and Nichols 1986), facilitating H2 activation (Zhang et al. 2021), enhancing H spillover and hydrogenation of adsorption surface intermediates, and improving methanol selectivity (Xu et al. 2021). The combination of ZnO and Cu-based catalysts has been shown to result in a synergism between Cu and ZnO (Chen et al. 1999), with SMSI that plays a crucial role in methanol synthesis (Lunkenbein et al. 2015). The surface synergism serves to generate Cu–Zn2+ active sites (Tu et al. 2023). Moreover, ZnO inclusion increases the adsorption strength of reaction intermediates, such as HCO, H2CO, and H3CO, and decreases the reaction energy barrier (Behrens et al. 2012). Din et al. (2019b) have doped CuZrO2/CNFs with ZnO (Cu–ZrO2), resulting in a decrease in the specific surface area (SCu) and dispersion (DCu) of copper. At 180 °C and 30 bar CO2/H2, doping CuZrO2/CNFs with 3 wt% ZnO resulted in an increase in methanol selectivity from 78 % to 92 %, methanol STY from 1.00 mmol/g/h to 1.40 mmol/g/h, and TOF from 1.41 × 10−3 s−1 to 1.87 × 10−3 s−1. However, the conversion of CO2 decreased slightly (from 14 % to 9 %), attributed to the Cu and ZnO synergistic effect. Zinc oxide promotes CuO reduction, increasing the number of active Cu–ZnO interfaces (Kattel et al. 2017b). In addition, ZnO serves as H binding site, promoting H2 adsorption (Chen et al. 1999), and contributing to the hydrogenation of CO2 and the H3CO* intermediate.

3.3.2 Increasing the dispersion of active metals and surface basicity

The addition of metal oxides can increase the dispersion of active metals. Din et al. (2016) have reported the synthesis of CZC–Nb by doping Cu/ZrO2/CNFs (CZC) with Nb2O5 using a deposition–precipitation procedure. A 9 % conversion of CO2 was achieved with CZC–Nb0.8 at 180 °C and 30 bar CO2/H2, with a methanol selectivity of 87 % and methanol STY of 1.25 mmol/g/h, exceeding the 75 % and 1.00 mmol/g/h values recorded when using CZC. The increased methanol STY due to the addition of Nb2O5 was attributed to the improved dispersion and reduction of the active Cu metal.

Doping with ZrO2 can increase the dispersion of the active components and surface basicity (Tichit et al. 2002), enhancing CO2 adsorption activation and catalytic activity. A series of CuZrO2/CNFs–O have been synthesized, containing 5, 10, 15, 20, and 25 wt% ZrO2 (Din et al. 2017), denoted as 5CZC, 10CZC, and 15CZC, respectively. Reaction at 180 °C and 30 bar CO2/H2 resulted in a 10 % CO2 conversion over 15CZC, with 67 % methanol selectivity and a methanol STY of 0.78 mmol/g/h, which was higher than the respective 3 %, 55 %, 0.56 mmol/g/h achieved using 5CZC. An increase in ZrO2 content increased the number of basic sites for CO2 adsorption activation. In addition, the specific copper surface was increased, contributing to improving methanol synthesis rate.

In the case of non-metal N-doped catalysts, the effective of doping is sensitive to treatment temperature and time. The lack of control of N doping in terms of content, and location (bulk, surface, and the bulk/surface proportions) has hampered a comprehensive understanding of possible structure-activity relationships. In the application of N-doped catalysts in gas–liquid(H2O)–solid phase hydrogenation of CO2, H2O* activates H2 via the amino group, promoting hydrogenation. Mass transfer is facilitated in gas-solid operation, but the full potential of the N-doped catalyst in low-temperature reactions has yet to be explored. While there are many possible metal doping strategies, the methods are not universal and particular procedures can only target specific catalytic systems. The degree of modulation offered by metal oxide doping is limited in terms of catalytic performance. In general, methanol selectivity drops when CO2 conversion is increased.

4 Defect engineering

4.1 Oxygen vacancy-promoted electron transfer

Oxygen vacancies (Ov) facilitate electron transfer, activating the linear CO2 molecules with a consequent alteration to the structural geometry. These vacancies promote methanol desorption from active sites, increasing overall methanol production (Chang et al. 2023). Moreover, the Ov can induce SMSI, promoting the formation of active interfaces for CO2 activation, and contributing additional stability to the dispersed active metals (Zhang et al. 2022). Toyao et al. (2019) have prepared Pt(3)/MoOx(30)/TiO2 by sequential impregnation of MoOx/TiO2 with Pt NPs. Application in the liquid-phase hydrogenation of CO2 at 150 °C (10 bar CO2 and 50 bar H2) resulted in a methanol yield of 73 %, approaching the equilibrium value. This was superior to yields recorded for other metals on MoOx(30)/TiO2 and Pt loaded on other supports (including Cu/Zn/Al2O3). In situ XANES analysis (Figure 8) has revealed that CO2 adsorption on Pt(3)/MoOx(30)/TiO2 resulted in a positive shift in the Mo K-edge signals, indicating Mo oxidation by CO2 with a consequent Ov reduction. This process serves to activate CO2 and promote the subsequent hydrogenation reaction. The Ov sites formed during the H2 pretreatment play an important role in CO2 hydrogenation. This study has established that the combination of Pt and Mo is crucial for achieving high catalytic activity, where the reduced and Ov-containing MoOx species are key to achieving low-temperature hydrogenation of CO2 to methanol.

Figure 8: 
In situ Mo K-edge XANES spectra of Pt(3)/MoOx(30)/TiO2 (Toyao et al. 2019); reproduced with permission from the American Chemical Society (copyright 2019).
Figure 8:

In situ Mo K-edge XANES spectra of Pt(3)/MoOx(30)/TiO2 (Toyao et al. 2019); reproduced with permission from the American Chemical Society (copyright 2019).

Ultrafine tetragonal ZrO2 rich in Ov (Chang et al. 2023) was used as a Cu support to catalyze CO2 hydrogenation to methanol. The Cu/ZrO2 (20CZ-OG) was prepared by oxalic acid-gel (OG) coprecipitation, which produced abundant Ov (36.2 %) generated by the self-reduction and thermal decomposition of oxalic acid complexes. By comparison, the Ov content of Cu/ZrO2 (20CZ-AC) prepared by ammonium bicarbonate (AC) co-precipitation was 23.4 %. In situ ESR (Electron Spin Resonance) analysis has shown that the Ov can accept electrons from metal clusters, resulting inelectron enrichment. Electrons transfer from Ov to CO2 serves to weaken the ESR signal, suggesting activation. In situ DRIFTS has demonstrated the contribution of Ov to CO2 adsorption and the formation of COOH* intermediates. In addition, Ov facilitates methanol desorption, increasing overall yields. At 200 °C and 40 bar CO2/H2, a 90 % methanol selectivity was achieved using 20CZ-OG with a methanol STY of 5.62 mmol/g/h, significantly higher than that obtained with 20CZ-AC (2.81 mmol/g/h) and Cu/ZnO/Al2O3 (1.56 mmol/g/h and ⁓80 % methanol selectivity).

4.2 Sulfur vacancy (Sv) regulation of C–O dissociation and H2 activation

Metal oxides such as In2O3 (Martin et al. 2016) and ZnZrOx (Wang et al. 2023b) can catalyze CO2 hydrogenation to methanol through an H-assisted dissociation mechanism at temperatures above 300 °C. In this case, the involvement of the RWGS reaction (Wang et al. 2020; Ye et al. 2024) results in low methanol selectivity. The introduction of transition metals (such as Cu, Pd, and Ni) to metal oxides can promote H2 activation and enhance catalytic activity at relatively low temperatures (<260 °C) (Danaci et al. 2016; Rui et al. 2017; Zhang et al. 2023). However, an over-hydrogenation of CO2 to CH4 and the RWGS reaction is promoted, lowering methanol selectivity (Kattel et al. 2017a; Yin et al. 2018). The interplay between C–O dissociation and H2 activation restricts a synchronous improvement in CO2 conversion and selectivity. Research has found that Sv-rich FL-MoS2 nanosheet catalysts (Hu et al. 2021) can address this limitation by promoting the dissociation of CO2 into surface-bound CO* and O*, and activating H2 on Sv. This serves to suppress deep hydrogenolysis of methanol to produce CH4, achieving enhanced methanol selectivity. Reaction at 180 °C resulted in a 12.5 % CO2 conversion over FL-MoS2, with a methanol selectivity of 94.3 %, superior to the 4.2 % CO2 conversion and methanol selectivity of 80.9 %, obtained with the commercial Cu/ZnO/Al2O3 catalyst. In addition, the FL-MoS2 catalyst exhibited stable activity over 3,000 h, demonstrating the effectiveness of Sv in promoting the low-temperature hydrogenation of CO2 to methanol.

Plane molecular structure causes MoS2 challenging to generate in-plane Sv. (Zhang et al. 2024a) synthesized MoS2-NS with in-plane Sv using NaCl-assisted method and applied it to CO2 hydrogenation to methanol. Na facilitates not only the formation of in-plane Sv, but also the adsorption of CO2 on Sv, following methanol production through the COOH* pathway, versus the dissociation pathway of CO2–CO* on Sv in absence of Na. At 200 °C and 30 bar reaction, Na-modified MoS2-NS-60 achieved methanol STY of 17.82 mmol/g/h with CO2 conversion of 4.8 % and methanol selectivity of 96 %. TOF based on total Sv reached 170 h−1. The main reason for the enhancement is the increased in-plane Sv and the promoted H2 adsorption/dissociation.

Ov and Sv can both serve as active sites to promote the adsorption and activation of reactants, and both affect the stability of reaction intermediates by altering the surface electronic state of the catalyst. Ov enhances CO2 adsorption and activation by exposing unsaturated metal sites or altering electronic structure, which requires synergistic activation of H2 through metal sites such as Cu or Pt. Ov promotes the hydrogenation of CO2 to form HCOO*/COOH* intermediates, reduces the subsequent hydrogenation energy barrier, and inhibits RWGS reaction. Sv directly adsorbs and activates CO2 and H2 by forming coordination unsaturated Mo atoms, forming CO and H atoms, avoiding over hydrogenation to produce CH4, and improving methanol selectivity. It achieves efficient catalysis through the confinement effect of two-dimensional structure and optimization of Sv density.

CO2 activation by Ov also has a limited impact in terms of low-temperature hydrogenation. A combination of precious metals (notably Pt) and partially reducible carriers (MoOx(30)/TiO2) is required to enhance the Ov contributions. The in-plane Sv associated with FL-MoS2 represents a stable catalytic but the level of CO2 conversion to methanol is still far removed from the equilibrium conversion.

5 Size effects

5.1 Role of size (cluster) in promoting methanol formation

The dispersion of metals, within a specific range, is directly related to particle size. A significant number of studies have shown that the dispersion/size of active metals affects catalytic performance (Men et al. 2019; Zhu et al. 2020a; Zhou et al. 2023). If the interface between the metal and carrier represents an active site, the high dispersion/small metal size increases the number of active interfaces and enhances catalytic activity (Fujiwara et al. 2019; Zhang et al. 2024b). A low metal dispersion (or large particle size) limits catalyst activation and reaction rate. Furthermore, different degrees of dispersion produce active sites with different properties, affecting product selectivity. Ting et al. (2019) prepared a Re(1)/TiO2 catalyst using ultrasound and impregnation. The catalyst was used to promote the liquid-phase hydrogenation of CO2 at 150 °C and 60 bar CO2/H2, delivering a TON of 44 with a methanol selectivity of 82 %. An increase in Re loading (0.2−20 wt%) was accompanied by an initial increase and subsequent decrease in activity and methanol selectivity. At 1 wt% Re, the catalytic performance was optimal. At low Re loading, a high density of isolated Re atoms increased the generation of CO. At higher loading, the formation of larger Re NPs resulted in the formation of CH4. At a Re loading of 1 wt%, highly dispersed and sub-nanometer-sized Re clusters were generated that promoted the formation of methanol. XANES characterization (Figure 9) has indicated that the Re valence state decreased with increasing reduction temperature. A low and high reduction temperatures did not result in methanol production. At the lower reduction temperature, Re exhibited a +6 valence. At higher reduction temperatures, Re exhibited a 0 valence, which resulted in Re sintering. Sample reduction temperature at 500 °C generated sub-nanometer-sized Re with a valence state between 0 and +4 that served as the active site for methanol synthesis.

Figure 9: 
XANES spectra results for Re(1)/TiO2 at different reduction temperatures: (a) L3-edge intensity and (b) binding energy chemical shifts (Ting et al. 2019); reproduced with permission from the American Chemical Society (copyright 2019).
Figure 9:

XANES spectra results for Re(1)/TiO2 at different reduction temperatures: (a) L3-edge intensity and (b) binding energy chemical shifts (Ting et al. 2019); reproduced with permission from the American Chemical Society (copyright 2019).

In a comparative study conducted at 150 °C and 56 bar CO2/H2 (Phongprueksathat et al. 2023), Re/TiO2 delivered a 4 % CO2 conversion with a methanol selectivity of 65 %, whereas Cu/ZnO/Al2O3 exhibited a CO2 conversion below 1 % and a methanol selectivity of 85 %. A higher Re loading resulted in larger Re particles that promoted by-products (CH4 and HCOOCH3) with lower CO2 conversions and methanol selectivities. At a 3 wt% Re loading, highly dispersed Reδ+ clusters with a high number of single atom Re0 species were generated. The Re0 sites act as an H2 activator to transfer H to nearby Reδ+ species responsible for activating CO2 to form the HCOO* intermediate followed by hydrogenation to produce methanol. Both Re0 and Reδ+ are required for Re/TiO2 to efficiently promote methanol formation. Insufficient H provided by Re0 gives rise to a decomposition of formic acid into CO, whereas an excess results in over-hydrogenation to CH4. In contrast, Reδ+ activates CO2, stabilizes HCOO*, and promotes hydrogenation to methanol.

The activity of catalysts is related to the size of active metal particles, resulting in structural sensitivity (Sun et al. 2017). Wu et al. (2017) have examined the effect of different Au particle sizes on CO2 hydrogenation activity at 180 °C and 40 bar CO2/H2, demonstrating that the TOF for Au(1.6)/ZrO2 (Au cluster size of 1.6 nm) was 20 h−1 with a methanol selectivity of 75 %, much higher than the values (8 h−1 with 55 % selectivity) for Au(2.4)/ZrO2. The superior selective hydrogenation performance was attributed to the formation of smaller Au particles and more Au–ZrO2 interfaces (where Au and ZrO2 interact synergistically). Xiao et al. (2017) produced a series of Cu/ZnO/Al2O3/ZrO2 catalysts using a co-precipitation hydrothermal method, applying different pH values; the samples are denoted as RHT-6, 7, 8, 9, 10, 11. An increase in pH was accompanied by an initial decrease and subsequent increase in Cu particle size, CO2 conversion, and methanol STY, indicating a negative correlation between Cu particle size and activity. The RHT-9 catalyst, prepared at a pH of 9.0, exhibited the smallest Cu particle size, stronger interaction between Cu and ZnO, and the highest active Cu dispersion. Moreover, RHT-9 delivered the highest CO2 conversion (10.7 %), methanol STY (2.72 mmol/g/h), and methanol selectivity (81.8 %), with a TOF of 7.34 × 10−3 (at the 190 °C and 50 bar CO2/H2). The results indicate a dependency of activity on Cu particle size. A smaller Cu particle size provides more active sites to convert the feed-gas, where a surface synergism between Cu and ZnO facilitates H2 activation and dissociation. Smaller active metal particles lower the energy barrier for CO2 activation, stabilizing intermediates, and improving methanol selectivity. In the case of Cu–TiO2HT and Cu–TiO2IMP catalysts synthesized using hydrothermal and initial wet impregnation methods, respectively (Sharma et al. 2021), Cu–TiO2HT exhibited smaller Cu NPs (3−5 nm) whereas a larger Cu NP size (approximately 15 nm) was associated with Cu–TiO2IMP. The Cu–TiO2HT delivered a superior catalytic performance with 9.4 % CO2 conversion and a methanol selectivity of 96 % (at 200 °C and 30 bar CO2/H2), whereas CO2 conversion over Cu–TiO2IMP was 6 % with a methanol selectivity of 62 %. The application of DFT calculations have shown that the activation barrier for CO2 on Cu–TiO2HT with a smaller Cu size (93 kJ/mol) was significantly lower than that (127 kJ/mol) determined for Cu–TiO2IMP with a larger Cu size. A higher catalytic activity and methanol selectivity for Cu–TiO2HT were attributed to an enhanced dispersion of small-sized Cu NPs, which converted CO2 to methanol via the RWGS pathway. Highly dispersed Cu particles increase the number of active sites for activation and dissociation of CO2 into CO, improving CO2 conversion. Smaller Cu NPs facilitate the formation of a greater number of Cu–TiO2 interfaces, stabilizing the reaction intermediates (COOH*, CO*, and HCO*), inhibiting desorption, and achieving high methanol selectivity.

5.2 Enhanced catalytic activity and stability due to higher metal dispersion

Improving the dispersion of active metals, preventing their aggregation, and stabilizing interfacial active sites can enhance catalyst stability. The Cu–ZnO interface plays an important role in Cu/ZnO catalyzed CO2 hydrogenation to methanol (Sun et al. 2020; Cui et al. 2021). The aggregation of Cu NPs diminishes the active Cu–ZnO interface, directly affecting catalytic stability and methanol selectivity. Bimetallic organic framework (bi–MOF) templates have been utilized to produce catalysts with well-distributed and well-arranged elements with an increased number of interfacial active sites (Qi et al. 2021). Han et al. (2022) prepared a CuZnO–MOF-74-350 using a Cu–Zn bi–MOF template that showed a preponderance of Cu–ZnO interfaces, which prevented the aggregation of Cu particles. In the CO2/H2 reaction conducted at 190 °C and 40 bar, the catalyst delivered a 7.5 % CO2 conversion with a 78.2 % methanol selectivity and stable activity for more than 100 h. In contrast, the methanol selectivity for a commercial Cu/ZnO/Al2O3 declined from 62.7 % to 44.3 % over 15 h due to a sintering of Cu and ZnO phases (An et al. 2017). The better performance shown by CuZnO-MOF-74-350 was attributed to greater availability of Cu–ZnO interfaces, which facilitated CO2 adsorption and stabilized the formate intermediate formate. The catalyst exhibited high Cu dispersion (52.0 %) when compared with the physically mixed MOF-based CuZnO–MM-350 (28.2 %), inhibiting sintering of the active metal particles.

The function of the carrier is to disperse the active metals, ideally achieving separation of single active metal atoms. When compared with clusters/NPs or large particles, single-atom metals exhibit superior catalytic performance associated with distinct sizes and valence states. Research has shown that the particle sizes (Karelovic and Ruiz 2015; Zhu et al. 2020b) and valence states (Samson et al. 2014; Yu et al. 2020; Liu et al. 2024) of active metals have a significant impact on catalytic performance. The activity of a series of Cu/ZrO2 catalysts (CAZ-X, X = 1−15, representing Cu loading) is shown in Figure 10 for reactions conducted at 180 °C and 30 bar CO2/H2 (Zhao et al. 2022). The Cu in CAZ-1/2 was present in a highly dispersed single atom form, resulting in a methanol selectivity of 100 %. The loading amount affected the activity through changing the metal dispersion or size. When the loading was below 8 wt%, the CO2 conversion exhibited a linear increase with Cu loading. The single Cu atom sites in the Cu1–O3 units are highly effective in dissociating H2 with the assistance of O atoms, activating CO2 to generate HCOO*.

Figure 10: 
Catalytic activity of CAZ-X at 180 °C and 30 bar CO2/H2 (Zhao et al. 2022). Reproduced with permission from Springer Nature (copyright 2022).
Figure 10:

Catalytic activity of CAZ-X at 180 °C and 30 bar CO2/H2 (Zhao et al. 2022). Reproduced with permission from Springer Nature (copyright 2022).

A small number of reduced Cu clusters or NPs were formed in CAZ-4, which serve as active sites for RWGS. A greater number of small Cu clusters or NPs were formed in CAZ-8, resulting in increased CO selectivity. The large Cu particles associated with CAZ-12 and CAZ-15 did not activate CO2 or contribute to CO2 conversion and product selectivity. The CAZ-15-H catalyst was prepared by pretreating CAZ-15 with HNO3, where the large Cu NPs were partially removed, generating an activity and actual Cu loading comparable to CAZ-8. This response suggests that the active hydrogenation sites are very stable, and not affected by treatment with a strong acid. A Cu loading of 8 wt% is required for effective hydrogenation activity. To date, no direct correlation has been observed between metal loading and reaction mechanisms. The loading level dictates the formation of active site configurations (e.g., single atoms, clusters, NPs), consequently altering product selectivity.

Correlating the EXAFS analysis data and DFT theoretical calculations, it was concluded that Cu in CAZ-1 before and after the reaction was present in a monodisperse state, with high thermal stability. The local structure of CAZ-1 consists of an isolated Cu and three O atoms (Cu1–O3 units) where the copper present as Cu+ is responsible for converting CO2 into methanol. The presence of copper in CAZ-15 is in the form of small Cu0 clusters or particles and Cuδ+ species (δ ≈ 1.4) serves to convert CO2 into CO and methanol, respectively.

The reaction pathway for CO2 hydrogenation to methanol at the Cuδ+ (1 < δ < 2) active sites, based on DFT calculations, is shown in Figure 11. The energy barrier for CO2 hydrogenation to form HCOO* is low (0.08 eV), whereas the energy barrier for forming COOH* is high (0.43 eV). Therefore, CO2 hydrogenation to methanol at Cuδ+ (1 < δ < 2) active sites follows the HCOO* pathway. This study has confirmed that, when Cu species are dispersed as single atoms and the ZrO2 surface exhibits uniform Cu1–O3 active sites, CO2 is selectively (100 %) converted to methanol. When the Cu species exist in clusters or small NPs, they inhibit the formation of methanol rather than serving as active sites for producing methanol. Large Cu particles are not active sites in CO2 hydrogenation. Low loadings easily form single atom Cu sites serving as efficient active sites whereas high loadings easily form Cu clusters, small Cu NPs or large Cu particles, which are not conducive to methanol generation. Single atom Cu shows being the most effective active site. However, its activation ability of CO2 or H2, hydrogenation rate, and even stability need to be further explored to raise the activity. Generally, a lower loading is advantageous from the viewpoint of practical application, for considering both the cost and the maximizing utilization of metals. However, the metal loading must be controlled to achieve a single-atom state while maximizing the loading of active metal, thereby enhancing catalytic activity. There is currently no consensus on the optimal Cu loading for commercialization.

Figure 11: 
Gibbs free energy for CO2 hydrogenation at Cuδ+ sites (Zhao et al. 2022); reproduced with permission from Springer Nature (copyright 2022).
Figure 11:

Gibbs free energy for CO2 hydrogenation at Cuδ+ sites (Zhao et al. 2022); reproduced with permission from Springer Nature (copyright 2022).

An appropriate active site size effectively promotes selective CO2 hydrogenation, but the requisite low loading to achieve a smaller size and higher dispersion limits this promotional effect. Developing active sites with the requisite size at high loading is very challenging. The use of MOF materials as supports is effective in obtaining higher loadings of well-dispersed active metals, but the feasible temperature window for catalytic operation has restricted practical applications.

6 Coordination effects

Single atom catalysts (SACs) have an adjustable coordination environment with high utilization of metal atoms, which are important features in thermal catalysis. Catalytic processes involve multiple reaction pathways, which lead to the formation of a range of products (CO, methane, formic acid, methanol, and higher alcohols) during CO2 hydrogenation, often limiting the selectivity of the target product (Zhang et al. 2017b; Li et al. 2018; Ye et al. 2022). It is necessary to control the reaction pathways in order to develop highly active and selective catalysts for CO2 hydrogenation. The single-atom active metal coordination can anchor metals and form single atoms that determine the prevalent reaction pathway, catalytic activity, and selectivity. Single atom Cu–N4 and Cu–N3 catalysts with specific coordination characteristics were prepared for use in the liquid-phase low-temperature hydrogenation of CO2 by varying melamine/active metal addition and the calcination atmosphere (Yang et al. 2021). Characterization by XANES, EXAFS, and DFT calculations have shown that Cu atoms are stable in a metallic state when anchored on C3N4. In the Cu–N4 structure, Cu atoms are located in C3N4 hole locations and coordinate with four N atoms. In the Cu–N3 structure, Cu atoms replace one C atom in C3N4 and coordinate with three N atoms. In the hydrogenation of CO2 at 150 °C and 32 bar, methanol selectivity and STY over Cu–N4 reached 95.5 % and 4.2 mmol/g/h, respectively, 4.2 times that of commercial Cu/ZnO/Al2O3 (1 mmol/g/h). The CO selectivity for Cu–N3 reached 94.3 % with a CO yield of 2.5 mmol/g/h. When Cu NPs supported on C3N4 were used, methanol selectivity and STY were 78.6 % and 1.8 mmol/g/h, respectively. The catalytic performance of single atom Cu catalysts was much higher than that of nanoparticle Cu systems. Mechanistic studies have shown that CO2 hydrogenation reaction pathways and product distribution are closely related to Cu atom coordination. The Cu–N4 SAC facilitates methanol production via the formate pathway whereas Cu–N3 SACs promotes CO formation via the RWGS pathway (Figure 12).

Figure 12: 
CO2 hydrogenation pathways and atomic configurations for (a) Cu–N4 SACs and (b) Cu–N3 SACs (Yang et al. 2021); reproduced with an open access Creative Commons Attribution (CC BY) license.
Figure 12:

CO2 hydrogenation pathways and atomic configurations for (a) Cu–N4 SACs and (b) Cu–N3 SACs (Yang et al. 2021); reproduced with an open access Creative Commons Attribution (CC BY) license.

Analysis by XPS (Figure 13) has indicated that the coordination structure directly affects the Cu valence state, where Cu in Cu–N4 SACs is present as +1 valence. The Cu–N3 SACs reveal characteristic Cu2+ and Cu+ XPS peaks, with a Cu2+/Cu+ ratio of approximately 2. In contrast, Cu NPs/C3N4 presents Cu0 and Cu2+ XPS peaks, where Cu is predominantly in the 0 valence state. Compared to Cu–N4 SACs, Cu–N3 SACs transferred more electrons from Cu to the adjacent N atoms. In the case of Cu–N4 SACs, DFT calculations have shown that the first H binds with the C atom of CO2 to form HCOO*, instead of forming the HOCO* intermediate (the free energy for HOCO* formation was higher by 0.30 eV than that for HCOO* formation), followed by hydrogenation to generate HCOOH*, H2COOH*, H2CO* and CH3OH*. The formation of H2CO* was energetically allowed through the cleavage of the C–O bond in H2COOH* with the free energy lowered by 0.36 eV. By contrast, DFT calculations indicate that the formation of HCOO* and HOCO* completely occurred on Cu–N3 SAC with the free energy of 0.59 and 0.63 eV, respectively. However, the formation of H2COOH* via HCOOH* hydrogenation was not allowed in a high free energy barrier of 0.75 eV. Thus, HOCO* was converted into CO via the RWGS pathway. In the case of Cu–N3 SACs, the formation of HCOOH* involves a high energy barrier, and H combines with the O atom of CO2 to form HOCO*, which dissociates to form CO.

Figure 13: 
Cu 2p XPS spectra for Cu–N-based catalysts (Yang et al. 2021); reproduced with an open access Creative Commons Attribution (CC BY) license.
Figure 13:

Cu 2p XPS spectra for Cu–N-based catalysts (Yang et al. 2021); reproduced with an open access Creative Commons Attribution (CC BY) license.

Chen et al. (2019) have reported that the synergism between Pt and coordinated O atoms in Pt1@MIL (single atom Pt) induced different reaction pathways, improving selective methanol production. In Pt1@MIL, Pt is atomically dispersed on MIL-101, where each Pt is bonded to four O atoms and Pt is present in the +2 valence state. In Ptn@MIL, the supported Pt forms 1.8 nm clusters. Analysis by EXAFS has demonstrated Pt–Pt metal bonding, where most of the Pt content is present in a 0 valence. Therefore, the coordination between Pt and O atoms caused Pt atoms to transfer electrons to O atoms, resulting in Pt assuming +2 valence state. In the hydrogenation of CO2 in DMF at 150 °C and 32 bar, the TOF for Pt1@MIL was 117 h−1, 5.6 times higher than Ptn@MIL (21 h−1) and 39 times that of Cu/ZnO/Al2O3 (3 h−1). Methanol selectivity was 90.3 % in the case of Pt1@MI, significantly higher than Ptn@MIL (13.3 %) where CO was the main product. The discrepant adsorption of H atoms on Pt1@MIL and Ptn@MIL induced different reaction pathways in CO2 hydrogenation. For Pt1@MIL, when the first H combined with CO2, the formation energy of HCOO* was 0.26 eV lower than that of COOH*. In this case, the first H atom was combined with the C atom of CO2 to form HCOO* as the stable intermediate. In comparison, Ptn@MIL favored the generation of COOH* whose formation energy was 0.34 eV lower than that of HCOO*. The transformation of CO2 into COOH* was the dominating reaction pathway. Mechanistic studies have revealed that Pt1@MIL promoted the HCOO* pathway whereas Ptn@MIL followed the COOH* pathway. The Pt single atom in Pt1@MIL forms an active site involving O atom coordination, where metal-ligand synergism promotes H2 dissociation with the formation of an O–H group. The hydroxyl H atom combines with CO2 to generate HCOO* as an intermediate, which is hydrogenated to HCOOH* and ultimately converted to methanol. In Ptn@MIL, Pt interacts with hydrogen to form Pt–H, where the H atom reacts with CO2 to generate COOH* as an intermediate, which dissociates to form CO.

In SACs, the interaction between neighboring active metal atoms modifies the coordination environment of the metal active site, impacting the reaction pathway and catalytic performance. Li et al. (2018) have demonstrated the synergistic effect of neighboring Pt sites on MoS2 which significantly enhanced CO2 hydrogenation activity, altering the reaction pathway and lowering the reaction activation energy (relative to isolated Pt sites). Each Pt atom and an associated activated S atom form an active site. When two active sites partially overlap or are adjacent, the two associated Pt atoms represent neighboring Pt monomers. Maintaining an atomic level dispersion of Pt and increasing the Pt loading to 7.5 % generated an abundance of neighboring Pt monomers. Pt species in atomically dispersed Pt/MoS2 were oxidized, while the oxidation of Pt species (Ptδ+) was lowered with the Pt mass loading. The XANES spectrum of Pt NPs/MoS2 was similar to that of Pt foil, indicating that the Pt species in Pt NPs/MoS2 were mainly in the 0 valence state. Therefore, Pt atoms in atomically dispersed Pt/MoS2 transferred electrons to the S atoms bonded with them. CO2 hydrogenation over 7.5 % Pt/MoS2 at 150 °C generated a TOF of 162.5 h−1, 14.8 times higher than 0.2 %Pt/MoS2. The activation energy for 7.5 % Pt/MoS2 with neighboring Pt monomers as the principal component was 72.3 kJ/mol, significantly lower than that recorded (124.7 kJ/mol) for 0.2 %Pt/MoS2 with isolated Pt monomers as the main component. This result suggests that neighboring Pt monomers lower the reaction activation energy and promote CO2 hydrogenation. Mechanistic studies have shown that the reaction pathways associated with neighboring and isolated Pt monomers are different. Isolated Pt monomers facilitate the conversion of CO2 to methanol without going through the formate pathway whereas neighboring Pt monomers first hydrogenate CO2 to generate the formate with further hydrogenation to methanol. H2 was dissociated on the activated S atom on Pt1/MoS2. As the formation energy of COOH* was 0.6 eV lower than that of HCOO*, COOH* was the stable intermediate upon the combination of the first H atom to CO2 on Pt1/MoS2 with an energy barrier of 0.40 eV (Figure 14). The energy barriers to generate HCOOH* and C(OH)2* were 1.39 and 0.74 eV, respectively. As such, the transformation of COOH* to C(OH)2* was the dominating reaction pathway. The reaction pathway on the neighboring Pt monomers differed from those on the isolated Pt monomers. After H2 dissociation on the activated S atom on Pt2iii/MoS2, COOH* was formed with an energy barrier of 0.75 eV. The energy barriers for the transformation of COOH* to HCOOH or C(OH)2* on Pt2iii/MoS2 were 1.01 or 1.34 eV, respectively; Thus, HCOOH as the intermediate was more favourable than C(OH)2* on Pt2iii/MoS2. DFT analysis (Figure 14) has shown that CO2 reacts with H on isolated Pt monomers (Pt1/MoS2) to form COOH* rather than HCOO*, with subsequent hydrogenation to generate C(OH)2*, CH(OH)2*, CHOH*, CH2OH*, and CH3OH. On neighboring Pt monomers (in Pt2iii/MoS2, the most stable model with three neighboring Pt monomers), CO2 reacts with H to form COOH*, followed by hydrogenation to give HCOOH, and subsequent formation of CH(OH)2*, CHOH*, CH2OH*, and CH3OH.

Figure 14: 
DFT studies for optimizing reaction pathways in CO2 hydrogenation over (a) Pt1/MoS2 and (b) Pt2iii/MoS2 (Li et al. 2018); reproduced with permission from Springer Nature (copyright 2018).
Figure 14:

DFT studies for optimizing reaction pathways in CO2 hydrogenation over (a) Pt1/MoS2 and (b) Pt2iii/MoS2 (Li et al. 2018); reproduced with permission from Springer Nature (copyright 2018).

The catalyst coordination structure has a substantial effect on the valence state distribution of active metals, which influences the prevailing catalytic pathway and product distribution. The coordination associated with high surface free energy SACs can result in aggregation and deactivation during catalyst preparation and reaction. Catalytic SAC systems have been applied to promote gas–liquid–solid low-temperature CO2 hydrogenation to methanol, but long-term stability has not received a detailed examination.

7 Surface modification

7.1 Hydrophobic modification

The accumulation of water on the catalyst surface and the competitive RWGS reaction affect methanol synthesis. Surface water can contribute to the sintering of the active copper species and result in CuO formation, which decreases stability and methanol selectivity (Prašnikar et al. 2019). Hydrophobic modification of the catalyst surface can impede sintering and improve the catalytic performance. Wang et al. (2024) have reported the synthesis of Cu/ZnO by oxalic acid-assisted solid-phase grinding with treatment using stearic acid as a hydrophobic functionalization agent. The results indicate that the hydrophobic treatment with appropriate concentrations of stearic acid can remove residues from the decomposition of oxalic acid, increasing specific surface area, surface basicity and reducibility of Cu/ZnO, while hindering Cu aggregation. The surface modification also increased the H2 adsorption capacity of Cu/ZnO with the formation of a surface H-rich region that enhanced the selective hydrogenation rate. Reaction at 200 °C reaction, generated a methanol selectivity and STY over Cu/Zn-2 mM of 78.8 % and 2.72 mmol/g/h, respectively, which was higher than the values (71.7 % and 2.15 mmol/g/h) obtained with Cu/ZnO. Furthermore, the Cu/Zn-2 mM was stable for over 100 h.

The influence of applicable to other supports (such as ZrO2, Al2O3) for hydrophobic modification and different types of hydrophobic modifiers on the catalyst surface and the mass transfer of CO2 have yet to be fully elucidated.

7.2 Hydrophilic modification

The surface-OH groups of catalysts determine hydrophilicity. By adjusting the surface hydrophilicity, the catalytic performance can be improved (Li and Zhao 2017). Surface hydrophilicity can influence the concentration of reactants on the catalyst surface. Peng et al. (2018) investigated the effect of surface –OH on SiC quantum dots (QDs) in CO2 hydrogenation at the molecular level. The surface of SiC QDs exhibited a large number of hydroxyl groups, contributing to hydrophilicity. At 150 °C and 32 bar CO2/H2, the methanol STY reached 169.5 mmol/g for hydrophilic SiC QDs, more than three orders of magnitude higher than the STY (0.1 mmol/g/h) obtained with hydrophobic commercial SiC (Figure 15a). The TOF of SiC QDs was 0.27 mmol/m2/h, much higher than the commercial SiC (0.01 mmol/m2/h, Figure 15b). In addition, the methanol selectivity in both cases was 100 %.

Figure 15: 
(a) Activity and (b) TOFmethanol for commercial SiC and synthetic SiC QDs (Peng et al. 2018); reproduced with permission from Elsevier (copyright 2018).
Figure 15:

(a) Activity and (b) TOFmethanol for commercial SiC and synthetic SiC QDs (Peng et al. 2018); reproduced with permission from Elsevier (copyright 2018).

The reaction activation energy associated with hydrophilic SiC QDs was 48.6 kJ/mol, significantly lower than that (94.7 kJ/mol) recorded for the hydrophobic commercial SiC. Mechanistic studies have shown that the activity enhancement of SiC QDs is closely related to hydrophilicity. The abundant active sites for H2 dissociation on the SiC (111) surface generate reactive H atoms. The H atoms of –OH on the SiC (111) surface react with CO2 to form HCOO*, releasing O atoms that combine with dissociated H atoms to regenerate –OH. This reaction pathway serves to lower the energy barrier for HCOO* formation and promotes CO2 activation, enhancing hydrogenation activity.

Furthermore, the study explored the use of Ni(OH)2, NiO, NiCo LDHs, NiCo2O4, and other hydroxyl-rich catalysts for the low-temperature CO2 hydrogenation to methanol. As expected, the metal hydroxides and LDHs exhibited significant enhancement to the activity versus their metal oxide counterparts. Specifically, the mass activity of Ni(OH)2 and NiCo LDHs, were 18.4 and 14.4 times higher than that of NiO and NiCo2O4, respectively. Besides hydroxyl groups, the enhanced activity also possibly resulted from other factors such as hydrogen dissolution, spillover, reducibility, and surface oxygen atoms as a result of different phases of metal oxides and hydroxides. The surface hydroxyl group density may be regarded as an activity descriptor for CO2 hydrogenation. Hydroxyl-rich nanocrystals were regarded as the candidate catalysts for CO2 hydrogenation. The key challenge lies in unambiguously identifying the role of hydroxyl groups and systematically excluding confounding variables. Also, the effect of modifiers on the surface hydroxyl density of nanocrystalline catalysts remains unclear.

7.3 Physical hybridization of additives

The formation of methanol can be promoted by an appropriate water coverage of the Cu single atoms on ZnO (Wu et al. 2022b). Introducing 0.11 vol% water in CO2 hydrogenation over Cu1/ZnO at 170 °C and 30 bar resulted in an increase in CO2 conversion to a maximum of 4.9 % after reaction for 3 h, with a subsequent decrease and stabilization at 1.8 %; the associated methanol selectivity was 99.1 %. In the absence of the water, CO2 conversion and methanol selectivity were 1.0 % and 93.5 %, respectively. If the water concentration is too low to disrupt the dynamic reaction equilibrium, methanol production remains unchanged. Mechanistic studies have shown that water promotes the transformation of CO2 into the COOH* and CH2O* intermediates. An increase in water content promotes the water gas shift (WGS) reaction with CO, with an increase in methanol selectivity. DFT calculations (Figure 16a) have shown that the CO2 hydrogenation reaction pathway on Cu1/ZnO under anhydrous conditions involves the reaction of CO2 with H to form HCOO* or COOH*. Formation of HCOO* is thermodynamically more favorable than COOH*, but still requires crossing an energy barrier (1.25 eV) that is higher than the barrier (0.81 eV) associated with COOH* formation. The COOH* pathway is more favorable, where COOH* is hydrogenated to form HCOOH*, which in turn generates H2COOH*, CH2O*, CH3O*, and CH3OH*. In this pathway, the apparent energy barrier for CO2 hydrogenation to methanol is 1.49 eV.

Figure 16: 
CO2 hydrogenation pathways on Cu1/ZnO (a) without and (b) with water (Wu et al. 2022b); reproduced with permission from John Wiley and Sons (copyright 2022).
Figure 16:

CO2 hydrogenation pathways on Cu1/ZnO (a) without and (b) with water (Wu et al. 2022b); reproduced with permission from John Wiley and Sons (copyright 2022).

In the case of CO2 hydrogenation on Cu1/ZnO in the presence of water, the energy barriers for the formation of COOH* and HCOO* are lower (Figure 16b), and the COOH* pathway is kinetically favorable. The apparent energy barrier mediated by water is 0.42 eV, much lower than that (1.49 eV) observed in the absence of water, confirming that water is favorable in increasing CO2 hydrogenation rate over Cu1/ZnO. Water provides H atoms to surface CO2 and reaction intermediates, lowering the activation energy barrier and increasing activity.

The hydrophilic SiC QD surface bearing –OH groups can promote methanol synthesis. However, the applicability of this strategy in the case of metal oxides and even active metals has not been fully explored. Following surface hydrophobic/hydrophilic modification, the role of the available functional groups in determining the energy barrier associated with intermediate formation and the stability of the intermediates requires further investigation. The addition of trace water can serve as a source of H to generate the COOH* intermediate, lowering the activation energy barrier, as has been demonstrated for the Cu/ZnO catalyzed hydrogenation of CO2. The addition of trace amounts of water for other catalysts has not been reported.

8 Reaction process

Industrial methanol synthesis is carried out in a fixed bed continuous flow reactor at 200−300 °C and 50−100 bar CO2/H2 (Behrens et al. 2012; Kuld et al. 2014; Yusuf and Almomani 2023). Due to thermodynamic limitations, the conversion per pass of the feed-gas is very low. Effective measures are needed to control the reaction heat. Liquid-phase processes have received significant attention, and are conducted at lower reaction temperatures (Reller et al. 2014; Scharnagl et al. 2019; Sen et al. 2022).

8.1 Catalytically active solvents

Fan et al. (1999) have reported a one-pot three-step reaction pathway for low-temperature methanol synthesis. The first step involves CO2 hydrogenation to formic acid. In Step 2, formic acid reacts with an ethanol solvent to produce ethyl formate. In Step 3, ethyl formate hydrogenation produces methanol and ethanol. Using ethanol as the solvent, reaction at 200 °C, 30 bar CO2/H2, and for 20 h, the conversion of CO2 over Cu–Cr–O was 15 % with a methanol selectivity of 80 %. The reaction when ethanol was not used as solvent proceeded at a lower rate, generating a lower methanol yield (by a factor of 10), confirming an alcohol-assisted pathway via the ethyl formate intermediate. Higher alcohols as co-catalysts promote methanol synthesis through the formic acid-ester intermediate route. Nieminen et al. (2018) reported methanol STY over Cu/ZnO of 0.26 mmol/g/h that reaction at 180 °C and 60 bar using 2-butanol as solvent. The control test using n-hexane as a solvent did not produce methanol, confirming the promotional effect of higher alcohol solvents.

Water as a product in CO2 hydrogenation to methanol blocks the surface Cu active sites, lowering reaction rate (Sahibzada et al. 1998). Utilizing molecular sieves (porous crystalline aluminosilicate compounds) as dehydrating agents has promoted methanol synthesis over Cu/ZnO. Reaction over Cu/ZnO at 150 °C and 50 bar using ethanol as the solvent, generated a methanol selectivity and yield of 87.22 % and 33.80 %, respectively (Boonamnuay et al. 2021). With the addition of a 3A molecular sieve, the in situ adsorption of water generated, during the reaction served to shift the equilibrium in favor of methanol generation, increasing the methanol selectivity and yield to 91.16 % and 42.8 %, respectively.

8.2 Cascade catalysis

The reduction of CO2 to methanol is a multi-electron (6e) conversion process, and achieving enhanced reaction performance using a single catalyst is challenging (Huff and Sanford 2011). Cascade catalysis in a single reactor utilizes a catalytic step sequence conversion to achieve high feedstock conversion. In a cascade catalytic low-temperature CO2 hydrogenation (Chen et al. 2015), a Cu–Cr catalyst was used to promote CO2 hydrogenation to the formate intermediate, whereas Cu/Mo2C catalyzed the hydrogenation of the formate to methanol. Operating the process at 135 °C and 40 bar, the selectivities of methanol and ethyl formate over Cu/Mo2C (Figure 17a) were 74 % and 20 %, respectively. The ethyl formate selectivity over Cu–Cr (Figure 17b) was 97 %, where subsequent hydrogenation to methanol was largely inhibited. The cascade reaction utilizing Cu/Mo2C and Cu–Cr (Figure 17c) resulted in an increase in methanol production by approximately 60 % compared with the total product formed in the single catalysis process. The selectivities of methanol and ethyl formate in the cascade reaction were 77 % and 20 %, respectively. The TOF associated with methanol synthesis in this cascade reaction reached 4.7 × 10−4 s−1, demonstrating that Cu/Mo2C and Cu–Cr catalysts synergistically catalyzed the hydrogenation of CO2 to methanol via the ethyl formate intermediate.

Figure 17: 
Cascade hydrogenation of CO2 over (a) Cu/Mo2C, (b) Cu-Cr, and (c) Cu-Cr +Cu/Mo2C. Reaction conditions: 135 °C; 10 bar CO2; 30 bar H2; 2 mL ethanol; 35.5 mL 1,4-dioxane. (Chen et al. 2015); reproduced with permission from the American Chemical Society (copyright 2015).
Figure 17:

Cascade hydrogenation of CO2 over (a) Cu/Mo2C, (b) Cu-Cr, and (c) Cu-Cr +Cu/Mo2C. Reaction conditions: 135 °C; 10 bar CO2; 30 bar H2; 2 mL ethanol; 35.5 mL 1,4-dioxane. (Chen et al. 2015); reproduced with permission from the American Chemical Society (copyright 2015).

Huff and Sanford (2011) have examined the application of homogeneous cascade catalysis to convert CO2 to methanol via formic acid and formate ester intermediates at 135 °C (Figure 18). The reaction involves three steps: (a) CO2 hydrogenation to formic acid; (b) esterification of formic acid to produce a formic acid-ester; (c) formic acid-ester hydrogenation releases methanol. The methanol TON generated by the cascade reaction over the three steps was 2.5, far lower than that (21) achieved by a stepwise cascade catalytic reaction, due to the incompatibility of the catalysts, where catalyst B (Sc(OTf)3) served to deactivate catalyst C ((PNN)Ru(CO)(H)). The application of a heterogeneous process can facilitate effective separation with improved performance.

Figure 18: 
Cascade hydrogenation of CO2 to methanol via formic acid and formate ester (Huff and Sanford 2011); reproduced with permission from the American Chemical Society (copyright 2011).
Figure 18:

Cascade hydrogenation of CO2 to methanol via formic acid and formate ester (Huff and Sanford 2011); reproduced with permission from the American Chemical Society (copyright 2011).

8.3 Coupled CO2 capture and hydrogenation to methanol

Kothandaraman et al. (2018) have studied the low-temperature hydrogenation of CO2 to methanol using tertiary amines and ethanol as solvents. Operating a Cu/ZnO/Al2O3 catalyst at 170 °C and 60 bar CO2/H2, with ethanol and NEt3 as promoters, the methanol yield reached 100 % (calculated based on the molar NEt3 content). As shown in Figure 19, the reaction involves three steps. In step (i), H2 is dissociated to generate surface reactive H atoms. In step (ii), CO2 reacts with NEt3 and EtOH to form an alkyl carbonate intermediate. In step (iii), the alkyl carbonate is hydrogenated to produce a formate salt that reacts with EtOH to form ethyl formate, which is then hydrogenated and dehydrated to produce methanol.

Figure 19: 
Coupled CO2 capture-hydrogenation to methanol (Kothandaraman et al. 2018); reproduced with permission from the Royal Society of Chemistry (copyright 2018).
Figure 19:

Coupled CO2 capture-hydrogenation to methanol (Kothandaraman et al. 2018); reproduced with permission from the Royal Society of Chemistry (copyright 2018).

Cascade hydrogenation of CO2 via the formate ester intermediate pathway requires at least two catalysts with the requirement for effective separation. The associated reaction kinetics for stepwise reactions and integrated cascade systems warrant further study.

9 Conclusions and future directions

Research advancements in alloying effects, doping engineering, defect engineering, size and coordination effects, and surface modification have contributed to the construction of catalytic active sites for the low-temperature hydrogenation of CO2 to methanol. The alloying effects, use of SiC QDs, and application of cascade catalysis have achieved better outcomes. Nevertheless, the level of CO2 conversion remains low.

Electron transfer associated with precious metal-containing alloys facilitates CO2 conversion due to the formation of active sites with high negative charge density, achieving a methanol selectivity up to 100 %. By introducing two active metals and reducing them to form an alloy, electrons are driven to transfer from the metal with a lower negative charge density to the one with a higher negative charge density, increasing the negative charge density of the active metals and promoting low-temperature CO2 activation. The application of non-precious metal alloys requires more research. Doping elements via pyrolysis under a controlled atmosphere or a co-precipitation method represents an efficient and straightforward approach. The use of dopants to construct catalytic active sites can enhance electron transfer, hydrogen spillover, active metal dispersion, and surface basicity, lowering the reaction activation energy. However, controlled doping, in terms of content and location, must be established to effectively adjust the electronic structure and catalytic performance of the active sites. Introducing noble metals enhances the dissociative adsorption of H2, or subjecting catalysts to H2 reduction generates more surface defects to construct active sites rich in Ov and Sv which can significantly enhance the CO2 adsorption and activation ability at low-temperature. The Ov defects formed by the combination of specific supports and active metals have shown evidence of improved CO2 activation and increased methanol production at low-temperature. The Sv sites associated with FL-MoS2 promote the dissociation of CO2 into CO* and O* at low-temperature, which are then hydrogenated to methanol, where the reaction mechanism is CO-hydrogenation pathway. The combination of Ov and Sv might be a potential strategy to enable the coexistence of two mainstream reaction pathways, promoting CO2 hydrogenation to methanol. Moreover, the impact of metal/non-metal doping on the Sv active site should enhance overall catalytic efficiency. Doping with N can introduce H adsorption sites and lower the reaction activation energy, promoting CO2 hydrogenation, where the reaction mechanism is the formate pathway. Constructing novel active sites containing N element may significantly reduce the apparent activation energy of the reaction, thereby accelerating the reaction process. The combination of SV and doping N may be a future research direction. In theory, the generation of bifunctional active sites and the simultaneous use of these two mainstream pathways will be beneficial for the progress of the reaction. The active metal size significantly impacts catalytic performance by regulating the valence metal states. Improvements in metal dispersion uniformity and sintering resistance should be the focus of future research. By reducing the loading of active metals, optimizing hydrothermal synthesis conditions during catalyst preparation, or incorporating MOF structures, the particle size and metal dispersion of the catalyst can be effectively improved. The catalyst coordination structure affects the valence states of active metals, the reaction pathway and associated activation energy barrier. The application of SACs can lower energy barriers by providing an adjustable coordination environment that ensures increased efficiency. By tuning the types and number of coordinating atoms to modify the local charge density of active metals, active sites favorable for methanol production can be engineered, thereby enhancing catalytic performance. Moreover, a hydrophobic catalyst surface can inhibit active site aggregation and increase H2 adsorption, whereas a hydrophilic surface promotes hydrogen spillover and lowers the energy barrier for the formation of hydrogenation intermediates. By introducing OH groups via template agents or urea-assisted hydrothermal methods, the hydrophilicity of the catalyst can be enhanced; hydrophobic surfaces can be created using modifiers like stearic acid to inhibit the aggregation of active metals. The trade-off between surface hydrophobicity/hydrophilicity is critical and requires fine-tuning. In terms of potentials, the defective NPs, highly dispersed alloys, or SACs structural catalysts warrant comprehensive investigation.

Three predominant mechanistic pathways have been revealed: HCOO* pathway, COOH* pathway, and CO-hydrogenation pathway, indicating the complexity and uncertainty of the reaction process and dependence of the catalysts. Strategies for constructing catalytic active sites greatly influence pathway dominance through changing the adsorption strength of the reactive intermediates. The competing mechanistic pathways substantially exist. On the whole, the precise mechanistic pathway for low-temperature CO2 hydrogenation remains controversial. Advanced characterization would deepen insights into the mechanisms, e.g., utilizing synchrotron radiation light sources coupled with in situ XAS/DRIFTS to monitor the low-temperature and high-pressure reaction, simultaneously dynamic changes in metal valence states and surface-adsorbed species are expected.


Corresponding author: Tian-Sheng Zhao, State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering, College of Chemistry & Chemical Engineering, Ningxia University, Yinchuan, 750021, China, E-mail:

Funding source: The Natural Science Foundation of Ningxia

Award Identifier / Grant number: 2022AAC02014

Funding source: The National Natural Science Foundation of China

Award Identifier / Grant number: 21965028

Acknowledgments

EditSprings for the linguistic checking.

  1. Research ethics: Not applicable.

  2. Informed consent: Not applicable.

  3. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  4. Use of Large Language Models, AI and Machine Learning Tools: None declared.

  5. Conflict of interest: The authors state no conflict of interest.

  6. Research funding: The Natural Science Fund of Ningxia (2022AAC02014) and the National Natural Science Fund of China (21965028).

  7. Data availability: Not applicable.

References

An, B., Zhang, J., Cheng, K., Ji, P., Wang, C., and Lin, W. (2017). Confinement of ultrasmall Cu/ZnOx nanoparticles in metal-organic frameworks for selective methanol synthesis from catalytic hydrogenation of CO2. J. Am. Chem. Soc. 139: 3834–3840, https://doi.org/10.1021/jacs.7b00058.Suche in Google Scholar PubMed

Anderson, A.B. and Nichols, J.A. (1986). Hydrogen on zinc oxide. Theory of its heterolytic adsorption. J. Am. Chem. Soc. 108: 4742–4746, https://doi.org/10.1021/ja00276a010.Suche in Google Scholar

Bai, S., Shao, Q., Feng, Y., Bu, L., and Huang, X. (2017). Highly efficient carbon dioxide hydrogenation to methanol catalyzed by zigzag platinum-cobalt nanowires. Small 13: 1604311, https://doi.org/10.1002/smll.201604311.Suche in Google Scholar PubMed

Behrens, M., Studt, F., Kasatkin, I., Kühl, S., Hävecker, M., Abild-Pedersen, F., Zander, S., Girgsdies, F., Kurr, P., Kniep, B.L., et al. (2012). The active site of methanol synthesis over Cu/ZnO/Al2O3 industrial catalysts. Science 336: 893, https://doi.org/10.1126/science.1219831.Suche in Google Scholar PubMed

Boonamnuay, T., Laosiripojana, N., Assabumrungrat, S., and Kim-Lohsoontorn, P. (2021). Effect 3A and 5A molecular sieve on alcohol-assisted methanol synthesis from CO2 and H2 over Cu/ZnO catalyst. Int. J. Hydrogen Energ. 46: 30948–30958, https://doi.org/10.1016/j.ijhydene.2021.04.161.Suche in Google Scholar

Chang, H., Gao, F., Luo, A., Liu, Y., Zhu, Y., He, H., and Cao, Y. (2023). Oxygen vacancy promoted carbon dioxide activation over Cu/ZrO2 for CO2-to-methanol conversion. Chem. Commun. 59: 7647–7650, https://doi.org/10.1039/d3cc01834b.Suche in Google Scholar PubMed

Chen, Y., Choi, S., and Thompson, L.T. (2015). Low-temperature CO2 hydrogenation to liquid products via a heterogeneous cascade catalytic system. ACS Catal. 5: 1717–1725, https://doi.org/10.1021/cs501656x.Suche in Google Scholar

Chen, Y., Choi, S., and Thompson, L.T. (2016). Low temperature CO2 hydrogenation to alcohols and hydrocarbons over Mo2C supported metal catalysts. J. Catal. 343: 147–156, https://doi.org/10.1016/j.jcat.2016.01.016.Suche in Google Scholar

Chen, H.Y., Lau, S.P., Chen, L., Lin, J., Huan, C.H.A., Tan, K.L., and Pan, J.S. (1999). Synergism between Cu and Zn sites in Cu/Zn catalysts for methanol synthesis. Appl. Surf. Sci. 152: 193–199, https://doi.org/10.1016/s0169-4332(99)00317-7.Suche in Google Scholar

Chen, Y., Li, H., Zhao, W., Zhang, W, Li, J., Li, W., Zheng, X., Yan, W., Zhang, Wenhua, Zhu, J., et al.. (2019). Optimizing reaction paths for methanol synthesis from CO2 hydrogenation via metal-ligand cooperativity. Nat. Commun. 10: 1885, https://doi.org/10.1038/s41467-019-09918-z.Suche in Google Scholar PubMed PubMed Central

Choi, E.J., Lee, Y.H., Lee, D.-W., Moon, D.J., and Lee, K.-Y. (2017). Hydrogenation of CO2 to methanol over Pd–Cu/CeO2 catalysts. Mol. Catal. 434: 146–153, https://doi.org/10.1016/j.mcat.2017.02.005.Suche in Google Scholar

Cored, J., Lopes, C.W., Liu, L., Soriano, J., Agostini, G., Solsona, B., Sánchez-Tovar, R., and Concepción, P. (2022). Cu-Ga3+-doped wurtzite ZnO interface as driving force for enhanced methanol production in co-precipitated Cu/ZnO/Ga2O3 catalysts. J. Catal. 407: 149–161, https://doi.org/10.1016/j.jcat.2022.01.032.Suche in Google Scholar

Cui, X., Yan, W., Yang, H., Shi, Y., Xue, Y., Zhang, H., Niu, Y., Fan, W., and Deng, T. (2021). Preserving the active Cu–ZnO interface for selective hydrogenation of CO2 to dimethyl ether and methanol. ACS Sustain. Chem. Eng. 9: 2661–2672, https://doi.org/10.1021/acssuschemeng.0c07258.Suche in Google Scholar

Danaci, S., Protasova, L., Lefevere, J., Bedel, L., Guilet, R., and Marty, P. (2016). Efficient CO2 methanation over Ni/Al2O3 coated structured catalysts. Catal. Today 273: 234–243, https://doi.org/10.1016/j.cattod.2016.04.019.Suche in Google Scholar

Din, I.U., Shaharun, M.S., Alotaibi, M.A., Alharthi, A.I., and Naeem, A. (2019a). Recent developments on heterogeneous catalytic CO2 reduction to methanol. J. CO2 Util. 34: 20–33, https://doi.org/10.1016/j.jcou.2019.05.036.Suche in Google Scholar

Din, I.U., Shaharun, M.S., Naeem, A., Tasleem, S., and Ahmad, P. (2019b). Revalorization of CO2 for methanol production via ZnO promoted carbon nanofibers based Cu-ZrO2 catalytic hydrogenation. J. Energy Chem. 39: 68–76, https://doi.org/10.1016/j.jechem.2019.01.023.Suche in Google Scholar

Din, I.U., Shaharun, M.S., Naeem, A., Tasleem, S., and Johan, M.R. (2017). Carbon nanofiber-based copper/zirconia catalyst for hydrogenation of CO2 to methanol. J. CO2 Util. 21: 145–155, https://doi.org/10.1016/j.jcou.2017.07.010.Suche in Google Scholar

Din, I.U., Shaharun, M.S., Subbarao, D., Naeem, A., and Hussain, F. (2016). Influence of niobium on carbon nanofibres based Cu/ZrO2 catalysts for liquid phase hydrogenation of CO2 to methanol. Catal. Today 259: 303–311, https://doi.org/10.1016/j.cattod.2015.06.019.Suche in Google Scholar

Donphai, W., Thepphankulngarm, N., Chaisuwan, T., Tanangteerapong, D., Rood, S.C., and Kongkachuichay, P. (2023). Catalytic performance of copper and ruthenium loaded on N-doped modified PBZ-derived carbons for CO2 hydrogenation. Chem. Eng. Sci. 274: 118693, https://doi.org/10.1016/j.ces.2023.118693.Suche in Google Scholar

Fan, L., Sakaiya, Y., and Fujimoto, K. (1999). Low-temperature methanol synthesis from carbon dioxide and hydrogen via formic ester. Appl. Catal A-Gen. 180: L11–L13, https://doi.org/10.1016/s0926-860x(98)00345-7.Suche in Google Scholar

Frost, J.C. (1988). Junction effect interactions in methanol synthesis catalysts. Nature 334: 577–580, https://doi.org/10.1038/334577a0.Suche in Google Scholar

Fujiwara, K., Tada, S., Honma, T., Sasaki, H., Nishijima, M., and Kikuchi, R. (2019). Influences of particle size and crystallinity of highly loaded CuO/ZrO2 on CO2 hydrogenation to methanol. AIChE J. 65: e16717, https://doi.org/10.1002/aic.16717.Suche in Google Scholar

García-Trenco, A., White, E.R., Regoutz, A., Payne, D.J., Shaffer, M.S.P., and Williams, C.K. (2017). Pd2Ga-based colloids as highly active catalysts for the hydrogenation of CO2 to methanol. ACS Catal. 7: 1186–1196, https://doi.org/10.1021/acscatal.6b02928.Suche in Google Scholar

Gau, A., Hack, J., Maeda, N., and Meier, D.M. (2021). Operando spectroscopic monitoring of active species in CO2 hydrogenation at elevated pressure and temperature: steady-state versus transient analysis. Energy Fuels 35: 15243–15246, https://doi.org/10.1021/acs.energyfuels.1c02592.Suche in Google Scholar

Graciani, J., Mudiyanselage, K., Xu, F., Baber, A.E., Evans, J., Senanayake, S.D., Stacchiola, D.J., Liu, P., Hrbek, J., Sanz, J.F., et al. (2014). Highly active copper-ceria and copper-ceria-titania catalysts for methanol synthesis from CO2. Science 345: 546–550, https://doi.org/10.1126/science.1253057.Suche in Google Scholar PubMed

Guo, T., Guo, Q., Li, S., Hu, Y., Yun, S., and Qian, Y. (2022). Effect of surface basicity over the supported Cu-ZnO catalysts on hydrogenation of CO2 to methanol. J. Catal. 407: 312–321, https://doi.org/10.1016/j.jcat.2022.01.035.Suche in Google Scholar

Han, H., Cui, P., Xiao, L., and Wu, W. (2021). MoCS@NSC with interfacial heterojunction nanostructure: a highly selective catalyst for synthesizing methanol from CO2 at low temperature. J. Environ. Chem. Eng. 9: 106354, https://doi.org/10.1016/j.jece.2021.106354.Suche in Google Scholar

Han, C., Zhang, H., Li, C., Huang, H., Wang, S., Wang, P., and Li, J. (2022). 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. 643: 118805, https://doi.org/10.1016/j.apcata.2022.118805.Suche in Google Scholar

Hu, J., Yu, L., Deng, J., Wang, Yong, Cheng, K., Ma, C., Zhang, Q., Wen, W., Yu, S., Pan, Y., et al.. (2021). Sulfur vacancy-rich MoS2 as a catalyst for the hydrogenation of CO2 to methanol. Nat. Catal. 4: 242–250, https://doi.org/10.1038/s41929-021-00584-3.Suche in Google Scholar

Huang, C., Zhang, S., Wang, W., Zhou, H., Shao, Z., Xia, L., Wang, H., and Sun, Y. (2024). Modulation of electronic metal-support interaction between Cu and ZnO by Er for effective low-temperature CO2 hydrogenation to methanol. ACS Catal. 14: 1324–1335, https://doi.org/10.1021/acscatal.3c04608.Suche in Google Scholar

Huff, C.A. and Sanford, M.S. (2011). Cascade catalysis for the homogeneous hydrogenation of CO2 to methanol. J. Am. Chem. Soc. 133: 18122–18125, https://doi.org/10.1021/ja208760j.Suche in Google Scholar PubMed

Jackson, C., Smith, G.T., Inwood, D.W., Leach, A.S., Whalley, P.S., Callisti, M., Polcar, T., Russell, A.E., Levecque, P., and Kramer, D. (2017). Electronic metal-support interaction enhanced oxygen reduction activity and stability of boron carbide supported platinum. Nat. Commun. 8: 15802, https://doi.org/10.1038/ncomms15802.Suche in Google Scholar PubMed PubMed Central

Jiang, X., Nie, X., Guo, X., Song, C., and Chen, J.G. (2020). Recent advances in carbon dioxide hydrogenation to methanol via heterogeneous catalysis. Chem. Rev. 120: 7984–8034, https://doi.org/10.1021/acs.chemrev.9b00723.Suche in Google Scholar PubMed

Kanega, R., Onishi, N., Tanaka, S., Kishimoto, H., and Himeda, Y. (2021). Catalytic hydrogenation of CO2 to methanol using multinuclear iridium complexes in a gas-solid phase reaction. J. Am. Chem. Soc. 143: 1570–1576, https://doi.org/10.1021/jacs.0c11927.Suche in Google Scholar PubMed

Karelovic, A. and Ruiz, P. (2015). The role of copper particle size in low pressure methanol synthesis via CO2 hydrogenation over Cu/ZnO catalysts. Catal. Sci. Technol. 5: 869–881, https://doi.org/10.1039/c4cy00848k.Suche in Google Scholar

Kattel, S., Liu, P., and Chen, J.G. (2017a). Tuning selectivity of CO2 hydrogenation reactions at the metal/oxide interface. J. Am. Chem. Soc. 139: 9739–9754, https://doi.org/10.1021/jacs.7b05362.Suche in Google Scholar PubMed

Kattel, S., Ramírez, P.J., Chen, J.G., Rodriguez, J.A., and Liu, P. (2017b). Active sites for CO2 hydrogenation to methanol on Cu/ZnO catalysts. Sci. 355: 1296–1299, https://doi.org/10.1126/science.aal3573.Suche in Google Scholar PubMed

Khan, M.U., Wang, L., Liu, Z., Gao, Z., Wang, S., Li, H., Zhang, W., Wang, M., Wang, Z., Ma, C., et al.. (2016). Pt3Co octapods as superior catalysts of CO2 hydrogenation. Angew. Chem., Int. Ed. 55: 9548–9552, https://doi.org/10.1002/anie.201602512.Suche in Google Scholar PubMed

Kothandaraman, J., Dagle, R.A., Dagle, V.L., Davidson, S.D., Walter, E.D., Burton, S.D., Hoyt, D.W., and Heldebrant, D.J. (2018). Condensed-phase low temperature heterogeneous hydrogenation of CO2 to methanol. Catal. Sci. Technol. 8: 5098–5103, https://doi.org/10.1039/c8cy00997j.Suche in Google Scholar

Kuld, S., Conradsen, C., Moses, P.G., Chorkendorff, I., and Sehested, J. (2014). Quantification of zinc atoms in a surface alloy on copper in an industrial-type methanol synthesis catalyst. Angew. Chem., Int. Ed. 53: 5941–5945, https://doi.org/10.1002/anie.201311073.Suche in Google Scholar PubMed

Li, W., Meng, R., Wang, K., Cheng, Y., Cai, D., and Zhan, G. (2024). Engineering pyridinic-N-Co sites for enhanced CO2 hydrogenation to methanol. Appl. Catal. B-Environ.: 124906, https://doi.org/10.1016/j.apcatb.2024.124906.Suche in Google Scholar

Li, H., Wang, L., Dai, Y., Pu, Z., Lao, Z., Chen, Y., Wang, M., Zheng, X., Zhu, J., Zhang, W., et al.. (2018). Synergetic interaction between neighbouring platinum monomers in CO2 hydrogenation. Nat. Nanotechnol. 13: 411–417, https://doi.org/10.1038/s41565-018-0089-z.Suche in Google Scholar PubMed

Li, Y. and Zhao, C. (2017). Enhancing water oxidation catalysis on a synergistic phosphorylated NiFe hydroxide by adjusting catalyst wettability. ACS Catal. 7: 2535–2541, https://doi.org/10.1021/acscatal.6b03497.Suche in Google Scholar

Liu, S., He, Y., Fu, W., Ren, J., Chen, J., Chen, H., Sun, R., Tang, Z., Mebrahtu, C., and Zeng, F. (2024). Synergy of Co0-Co2+ in cobalt-based catalysts for CO2 hydrogenation: quantifying via reduced and exposed atoms fraction. Appl. Catal A-Gen. 670: 119549, https://doi.org/10.1016/j.apcata.2023.119549.Suche in Google Scholar

Lunkenbein, T., Schumann, J., Behrens, M., Schlögl, R., and Willinger, M.G. (2015). Formation of a ZnO overlayer in industrial Cu/ZnO/Al2O3 catalysts induced by strong metal–support interactions. Angew. Chem., Int. Ed. 54: 4544–4548, https://doi.org/10.1002/anie.201411581.Suche in Google Scholar PubMed

Martin, O., Martín, A.J., Mondelli, C., Mitchell, S., Segawa, T.F., Hauert, R., Drouilly, C., Curulla-Ferré, D., and Pérez‐Ramírez, J. (2016). Indium oxide as a superior catalyst for methanol synthesis by CO2 hydrogenation. Angew. Chem., Int. Ed. 55: 6261–6265, https://doi.org/10.1002/anie.201600943.Suche in Google Scholar PubMed

Men, Y., Fang, X., Gu, Q., Singh, R., Wu, F., Danaci, D., Zhao, Q., Xiao, P., and Webley, P.A. (2020). Synthesis of Ni5Ga3 catalyst by hydrotalcite-like compound (HTlc) precursors for CO2 hydrogenation to methanol. Appl. Catal., B 275: 119067, https://doi.org/10.1016/j.apcatb.2020.119067.Suche in Google Scholar

Men, Y.L., Liu, Y., Wang, Q., Luo, Z.H., Shao, S., Li, Y.B., and Pan, Y.X. (2019). Highly dispersed Pt-based catalysts for selective CO2 hydrogenation to methanol at atmospheric pressure. Chem. Eng. Sci. 200: 167–175, https://doi.org/10.1016/j.ces.2019.02.004.Suche in Google Scholar

Mosrati, J., Ishida, T., Mac, H., Al-Yusufi, M., Honma, T., Parliniska-Wojtan, M., Kobayashi, Y., Klyushin, A., Murayama, T., and Abdel-Mageed, A.M. (2023). Low-temperature hydrogenation of CO2 to methanol in water on ZnO-supported CuAu nanoalloys. Angew. Chem., Int. Ed. 135: e202311340, https://doi.org/10.1002/anie.202311340.Suche in Google Scholar PubMed

Nieminen, H., Givirovskiy, G., Laari, A., and Koiranen, T. (2018). Alcohol promoted methanol synthesis enhanced by adsorption of water and dual catalysts. J. CO2 Util. 24: 180–189, https://doi.org/10.1016/j.jcou.2018.01.002.Suche in Google Scholar

Peng, Y., Wang, L., Luo, Q., Cao, Y., Dai, Y., Li, Z., Li, H., Zheng, X., Yan, W., Yang, J., et al.. (2018). Molecular-level insight into how hydroxyl groups boost catalytic activity in CO2 hydrogenation into methanol. Chem 4: 613–625, https://doi.org/10.1016/j.chempr.2018.01.019.Suche in Google Scholar

Phongprueksathat, N., Ting, K.W., Mine, S., Jing, Y., Toyoshima, R., Kondoh, H., Shimizu, K., Toyao, T., and Urakawa, A. (2023). Bifunctionality of Re supported on TiO2 in driving methanol formation in low-temperature CO2 hydrogenation. ACS Catal. 13: 10734–10750, https://doi.org/10.1021/acscatal.3c01599.Suche in Google Scholar PubMed PubMed Central

Prašnikar, A., Pavlišič, A., Ruiz-Zepeda, F., Kovač, J., and Likozar, B. (2019). Mechanisms of copper-based catalyst deactivation during CO2 reduction to methanol. Ind. Eng. Chem. Res. 58: 13021–13029, https://doi.org/10.1021/acs.iecr.9b01898.Suche in Google Scholar

Qi, T., Zhao, Y., Chen, S., Li, W., Guo, X., Zhang, Y., and Song, C. (2021). Bimetallic metal organic framework-templated synthesis of a Cu-ZnO/Al2O3 catalyst with superior methanol selectivity for CO2 hydrogenation. Mol. Catal. 514: 111870, https://doi.org/10.1016/j.mcat.2021.111870.Suche in Google Scholar

Reller, C., Pöge, M., Lißner, A., and Mertens, F.O.R.L. (2014). Methanol from CO2 by organo-cocatalysis: CO2 capture and hydrogenation in one process step. Environ. Sci. Technol. 48: 14799–14804, https://doi.org/10.1021/es503914d.Suche in Google Scholar PubMed

Rui, N., Wang, Z., Sun, K., Ye, J., Ge, Q., and Liu, C. (2017). CO2 hydrogenation to methanol over Pd/In2O3: effects of Pd and oxygen vacancy. Appl. Catal., B 218: 488–497, https://doi.org/10.1016/j.apcatb.2017.06.069.Suche in Google Scholar

Sahibzada, M., Metcalfe, I.S., and Chadwick, D. (1998). Methanol synthesis from CO/CO2/H2 over Cu/ZnO/Al2O3 at differential and finite conversions. J. Catal. 174: 111–118, https://doi.org/10.1006/jcat.1998.1964.Suche in Google Scholar

Samson, K., Śliwa, M., Socha, R.P., Góra-Marek, K., Mucha, D., Rutkowska-Zbik, D., Paul, J.-F., Ruggiero-Mikołajczyk, M., Grabowski, R., and Słoczyński, J. (2014). Influence of ZrO2 structure and copper electronic state on activity of Cu/ZrO2 catalysts in methanol synthesis from CO2. ACS Catal. 4: 3730–3741, https://doi.org/10.1021/cs500979c.Suche in Google Scholar

Scharnagl, F.K., Hertrich, M.F., Neitzel, G., Jackstell, R., and Beller, M. (2019). Homogeneous catalytic hydrogenation of CO2 to methanol: improvements with tailored ligands. Adv. Synth. Catal. 361: 374–379, https://doi.org/10.1002/adsc.201801314.Suche in Google Scholar

Schieweck, B.G., Jürling-Will, P., and Klankermayer, J. (2020). Structurally versatile ligand system for the ruthenium catalyzed one-pot hydrogenation of CO2 to methanol. ACS Catal. 10: 3890–3894, https://doi.org/10.1021/acscatal.9b04977.Suche in Google Scholar

Sen, R., Goeppert, A., and Surya Prakash, G.K. (2022). Homogeneous hydrogenation of CO2 and CO to methanol: the renaissance of low-temperature catalysis in the context of the methanol economy. Angew. Chem., Int. Ed. 61: e202207278, https://doi.org/10.1002/anie.202207278.Suche in Google Scholar PubMed PubMed Central

Sharma, S.K., Banerjee, A., Paul, B., Poddar, M.K., Sasaki, T., Samanta, C., and Bal, R. (2021). Combined experimental and computational study to unravel the factors of the Cu/TiO2 catalyst for CO2 hydrogenation to methanol. J. CO2 Util. 50: 101576, https://doi.org/10.1016/j.jcou.2021.101576.Suche in Google Scholar

Studt, F., Sharafutdinov, I., Abild-Pedersen, F., Elkjær, C.F., Hummelshøj, J.S., Dahl, S., Chorkendorff, I., and Nørskov, J.K. (2014). Discovery of a Ni-Ga catalyst for carbon dioxide reduction to methanol. Nat. Chem. 6: 320–324, https://doi.org/10.1038/nchem.1873.Suche in Google Scholar PubMed

Sugiyama, H., Miyazaki, M., Sasase, M., Kitano, M., and Hosono, H. (2023). Room-temperature CO2 hydrogenation to methanol over air-stable hcp-PdMo intermetallic catalyst. J. Am. Chem. Soc. 145: 9410–9416, https://doi.org/10.1021/jacs.2c13801.Suche in Google Scholar PubMed PubMed Central

Sun, Y., Chen, L., Bao, Y., Wang, G., Zhang, Y., Fu, M., Wu, J., and Ye, D. (2018). Roles of nitrogen species on nitrogen-doped CNTs supported Cu-ZrO2 system for carbon dioxide hydrogenation to methanol. Catal. Today 307: 212–223, https://doi.org/10.1016/j.cattod.2017.04.017.Suche in Google Scholar

Sun, Y., Huang, C., Chen, L., Zhang, Y., Fu, M., Wu, J., and Ye, D. (2020). Active site structure study of Cu/Plate ZnO model catalysts for CO2 hydrogenation to methanol under the real reaction conditions. J. CO2 Util. 37: 55–64, https://doi.org/10.1016/j.jcou.2019.11.029.Suche in Google Scholar

Sun, X., Suarez, A.I.O., Meijerink, M., Van Deelen, T., Ould-Chikh, S., Zečević, J., De Jong, K.P., Kapteijn, F., and Gascon, J. (2017). Manufacture of highly loaded silica-supported cobalt Fischer–Tropsch catalysts from a metal organic framework. Nat. Commun. 8: 1680, https://doi.org/10.1038/s41467-017-01910-9.Suche in Google Scholar PubMed PubMed Central

Tichit, D., Das, N., Coq, B., and Durand, R. (2002). Preparation of Zr-containing layered double hydroxides and characterization of the acido-basic properties of their mixed oxides. Chem. Mater. 14: 1530–1538, https://doi.org/10.1021/cm011125l.Suche in Google Scholar

Ting, K.W., Toyao, T., Siddiki, S.M.A.H., and Shimizu, K. (2019). Low-temperature hydrogenation of CO2 to methanol over heterogeneous TiO2-supported Re catalysts. ACS Catal. 9: 3685–3693, https://doi.org/10.1021/acscatal.8b04821.Suche in Google Scholar

Toyao, T., Kayamori, S., Maeno, Z., Siddiki, S.M.A.H., and Shimizu, K. (2019). Heterogeneous Pt and MoOx Co-loaded TiO2 catalysts for low-temperature CO2 hydrogenation to form CH3OH. ACS Catal. 9: 8187–8196, https://doi.org/10.1021/acscatal.9b01225.Suche in Google Scholar

Tu, W., Ren, P., Li, Y., Yang, Y., Tian, Y., Zhang, Z., Zhu, M., Chin, Y.-H.C., Gong, J., and Han, Y.-F. (2023). Gas-dependent active sites on Cu/ZnO clusters for CH3OH synthesis. J. Am. Chem. Soc. 145: 8751–8756, https://doi.org/10.1021/jacs.2c13784.Suche in Google Scholar PubMed

Wang, L., Zhang, W., Zheng, X., Chen, Y., Wu, W., Qiu, J., Zhao, Xiangchen, Zhao, Xiao, Dai, Y., and Zeng, J. (2017). Incorporating nitrogen atoms into cobalt nanosheets as a strategy to boost catalytic activity toward CO2 hydrogenation. Nat. Energy 2: 869–876, https://doi.org/10.1038/s41560-017-0015-x.Suche in Google Scholar

Wang, F., Chen, F., Guo, X., He, Y., Gao, W., Yasuda, S., Yang, G., and Tsubaki, N. (2024). Enhanced performance and stability of Cu/ZnO catalyst by hydrophobic treatment for low-temperature methanol synthesis from CO2. Catal. Today 425: 114344, https://doi.org/10.1016/j.cattod.2023.114344.Suche in Google Scholar

Wang, W., Qu, Z., Song, L., and Fu, Q. (2020). CO2 hydrogenation to methanol over Cu/CeO2 and Cu/ZrO2 catalysts: tuning methanol selectivity via metal-support interaction. J. Energy Chem. 40: 22–30, https://doi.org/10.1016/j.jechem.2019.03.001.Suche in Google Scholar

Wang, X., Yao, Z., Guo, X., Yan, Z., Ban, H., Wang, P., Yao, R., Li, L., and Li, C. (2023b). Modulating electronic interaction over Zr–ZnO catalysts to enhance CO2 hydrogenation to methanol. ACS Catal. 14: 508–521, https://doi.org/10.1021/acscatal.3c05524.Suche in Google Scholar

Wang, Peng, Zhang, H., Wang, S., and Li, J. (2023a). Controlling H2 adsorption of Cu/ZnO/Al2O3/MgO with enhancing the performance of CO2 hydrogenation to methanol at low temperature. J. Alloys Compd. 966: 171577, https://doi.org/10.1016/j.jallcom.2023.171577.Suche in Google Scholar

Wu, Qian, Liang, S., Zhang, T., Louis, B., and Wang, Q. (2022a). Current advances in bimetallic catalysts for carbon dioxide hydrogenation to methanol. Fuel 313: 122963, https://doi.org/10.1016/j.fuel.2021.122963.Suche in Google Scholar

Wu, W., Wang, Y., Luo, L., Wang, M., Li, Z., Chen, Y., Wang, Z., Chai, J., Cen, Z., Shi, Y., et al. (2022b). CO2 Hydrogenation over copper/ZnO single-atom catalysts: water-promoted transient synthesis of methanol. Angew. Chem., Int. Ed. 61: e202213024, https://doi.org/10.1002/anie.202213024.Suche in Google Scholar PubMed

Wu, C., Zhang, P., Zhang, Z., Zhang, L., Yang, G., and Han, B. (2017). Efficient hydrogenation of CO2 to methanol over supported subnanometer gold catalysts at low temperature. ChemCatChem 9: 3691–3696, https://doi.org/10.1002/cctc.201700872.Suche in Google Scholar

Xiao, S., Zhang, Y., Gao, P., Zhong, L., Li, X., Zhang, Z., Wang, H., Wei, W., and Sun, Y. (2017). Highly efficient Cu-based catalysts via hydrotalcite-like precursors for CO2 hydrogenation to methanol. Catal. Today 281: 327–336, https://doi.org/10.1016/j.cattod.2016.02.004.Suche in Google Scholar

Xie, B., Kumar, P., Tan, T.H., Asghar, A., Aguey-Zinsou, K.-F., Scott, J., and Amal, R. (2021). Doping-mediated metal-support interaction promotion towards light-assisted methanol production over Cu/ZnO/Al2O3. ACS Catal. 11: 5818–5828, https://doi.org/10.1021/acscatal.1c00332.Suche in Google Scholar

Xu, D., Hong, X., and Liu, G. (2021). Highly dispersed metal doping to ZnZr oxide catalyst for CO2 hydrogenation to methanol: insight into hydrogen spillover. J. Catal. 393: 207–214, https://doi.org/10.1016/j.jcat.2020.11.039.Suche in Google Scholar

Yang, T., Mao, X., Zhang, Y., Wu, X., Wang, L., Chu, M., Pao, C.-W., Yang, S., Xu, Y., and Huang, X. (2021). Coordination tailoring of Cu single sites on C3N4 realizes selective CO2 hydrogenation at low temperature. Nat. Commun. 12: 6022, https://doi.org/10.1038/s41467-021-26316-6.Suche in Google Scholar PubMed PubMed Central

Ye, R., Ma, L., Mao, J., Wang, X., Hong, X., Gallo, A., Ma, Y., Luo, W., Wang, B., Zhang, R., et al. (2024). A Ce-CuZn catalyst with abundant Cu/Zn-OV-Ce active sites for CO2 hydrogenation to methanol. Nat. Commun. 15: 2159, https://doi.org/10.1038/s41467-024-46513-3.Suche in Google Scholar PubMed PubMed Central

Ye, X., Ma, J., Yu, W., Pan, X., Yang, C., Wang, C., Liu, Q., and Huang, Y. (2022). Construction of bifunctional single-atom catalysts on the optimized β-Mo2C surface for highly selective hydrogenation of CO2 into ethanol. J. Energy Chem. 67: 184–192, https://doi.org/10.1016/j.jechem.2021.10.017.Suche in Google Scholar

Yin, Y., Hu, B., Li, X., Zhou, X., Hong, X., and Liu, G. (2018). Pd@zeolitic imidazolate framework-8 derived PdZn alloy catalysts for efficient hydrogenation of CO2 to methanol. Appl. Catal., B 234: 143–152, https://doi.org/10.1016/j.apcatb.2018.04.024.Suche in Google Scholar

Yu, J., Yang, M., Zhang, J., Ge, Q., Zimina, A., Pruessmann, T., Zheng, L., Grunwaldt, J.-D., and Sun, J. (2020). Stabilizing Cu+ in Cu/SiO2 catalysts with a shattuckite-like structure boosts CO2 hydrogenation into methanol. ACS Catal. 10: 14694–14706, https://doi.org/10.1021/acscatal.0c04371.Suche in Google Scholar

Yusuf, N. and Almomani, F. (2023). Highly effective hydrogenation of CO2 to methanol over Cu/ZnO/Al2O3 catalyst: a process economy & environmental aspects. Fuel 332: 126027.10.1016/j.fuel.2022.126027Suche in Google Scholar

Zhang, L., Hu, X., Wang, N., and Chen, B. (2024b). The copper size effect of CuZn/CeO2 catalyst in CO2 hydrogenation to methanol. Catal. Today 436: 114773, https://doi.org/10.1016/j.cattod.2024.114773.Suche in Google Scholar

Zhang, Xiao, Li, X., Zhang, D., Su, N.Q., Yang, W., Everitt, H.O., and Liu, J. (2017b). Product selectivity in plasmonic photocatalysis for carbon dioxide hydrogenation. Nat. Commun. 8: 14542, https://doi.org/10.1038/ncomms14542.Suche in Google Scholar PubMed PubMed Central

Zhang, C., Song, W., Sun, G., Xie, L., Wang, J., Li, K., Sun, C., Liu, H., Snape, C.E., and Drage, T. (2013). CO2 capture with activated carbon grafted by nitrogenous functional groups. Energy Fuels 27: 4818–4823, https://doi.org/10.1021/ef400499k.Suche in Google Scholar

Zhang, C., Wang, L., Chen, Y., He, X., Song, Y., Gazit, O.M., and Zhong, Z. (2023). Shifting CO2 hydrogenation from producing CO to CH3OH by engineering defect structures of Cu/ZrO2 and Cu/ZnO catalysts. Chem. Eng. J. 475: 146102, https://doi.org/10.1016/j.cej.2023.146102.Suche in Google Scholar

Zhang, C., Wang, L., Etim, U.J., Song, Y., Gazit, O.M., and Zhong, Z. (2022). Oxygen vacancies in Cu/TiO2 boost strong metal-support interaction and CO2 hydrogenation to methanol. J. Catal. 413: 284–296, https://doi.org/10.1016/j.jcat.2022.06.026.Suche in Google Scholar

Zhang, Wenbo, Wang, L., Liu, H., Hao, Y., Li, H., Khan, M.U., and Zeng, J. (2017a). Integration of quantum confinement and alloy effect to modulate electronic properties of RhW nanocrystals for improved catalytic performance toward CO2 hydrogenation. Nano Lett. 17: 788–793, https://doi.org/10.1021/acs.nanolett.6b03967.Suche in Google Scholar PubMed

Zhang, X., Zhang, G., Liu, W., Yuan, F., Wang, J., Zhu, J., Jiang, X., Zhang, A., Ding, F., Song, C., et al.. (2021). Reaction-driven surface reconstruction of ZnAl2O4 boosts the methanol selectivity in CO2 catalytic hydrogenation. Appl. Catal., B 284: 119700, https://doi.org/10.1016/j.apcatb.2020.119700.Suche in Google Scholar

Zhang, Z., Zuo, J., Luo, L., Yang, X., Ma, Z., Jin, H., Yuan, Y., Qian, Q., Chen, Q., and Luo, Y. (2024a). Sodium-assisted MoS2 for boosting CO2 hydrogenation to methanol: the crucial role of sodium in defect evolution and modification. J. Catal. 436: 115621, https://doi.org/10.1016/j.jcat.2024.115621.Suche in Google Scholar

Zhao, H., Yu, R., Ma, S., Xu, K., Chen, Y., Jiang, K., Fang, Y., Zhu, C., Liu, X., Tang, Y., et al.. (2022). The role of Cu1–O3 species in single-atom Cu/ZrO2 catalyst for CO2 hydrogenation. Nat. Catal. 5: 818–831, https://doi.org/10.1038/s41929-022-00840-0.Suche in Google Scholar

Zheng, X., Lin, Y., Pan, H., Wu, L., Zhang, W., Cao, L., Zhang, J., Zheng, L., and Yao, T. (2018). Grain boundaries modulating active sites in RhCo porous nanospheres for efficient CO2 hydrogenation. Nano Res. 11: 2357–2365, https://doi.org/10.1007/s12274-017-1841-7.Suche in Google Scholar

Zhou, X., Price, G.A., Sunley, G.J., and Copéret, C. (2023). Small cobalt nanoparticles favor reverse water-gas shift reaction over methanation under CO2 hydrogenation conditions. Angew. Chem., Int. Ed. 62: e202314274, https://doi.org/10.1002/anie.202314274.Suche in Google Scholar PubMed

Zhu, J., Zhang, G., Li, W., Zhang, X., Ding, F., Song, C., and Guo, X. (2020a). Deconvolution of the particle size effect on CO2 hydrogenation over iron-based catalysts. ACS Catal. 10: 7424–7433, https://doi.org/10.1021/acscatal.0c01526.Suche in Google Scholar

Zhu, Y., Zheng, J., Ye, J., Cui, Y., Koh, K., Kovarik, L., Camaioni, D.M., Fulton, J.L., Truhlar, D.G., Neurock, M., et al. (2020b). Copper-zirconia interfaces in UiO-66 enable selective catalytic hydrogenation of CO2 to methanol. Nat. Commun. 11: 5849, https://doi.org/10.1038/s41467-020-19438-w.Suche in Google Scholar PubMed PubMed Central

Received: 2025-01-28
Accepted: 2025-05-19
Published Online: 2025-06-05
Published in Print: 2025-08-26

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Heruntergeladen am 23.10.2025 von https://www.degruyterbrill.com/document/doi/10.1515/revce-2025-0003/html?srsltid=AfmBOoofxb4sy1r4Le404YplUiYAZf13GcfG8t7bc354H9P1HI1p_Kpb
Button zum nach oben scrollen