Next Article in Journal
Photocatalytic Degradation of Tetracycline by Supramolecular Materials Constructed with Organic Cations and Silver Iodide
Previous Article in Journal
Application of Low-Cost Plant-Derived Carbon Dots as a Sustainable Anode Catalyst in Microbial Fuel Cells for Improved Wastewater Treatment and Power Output
Previous Article in Special Issue
The Advanced Synthesis of MOFs-Based Materials in Photocatalytic HER in Recent Three Years
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Progress of Metal-Organic Framework for Boosting CO2 Conversion

1
Tianjin Key Laboratory of Structure and Performance for Functional Molecules, College of Chemistry, Tianjin Normal University, Tianjin 300387, China
2
State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China
3
School of Chemistry and Chemical Engineering, Qilu University of Technology, Jinan 250353, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1582; https://doi.org/10.3390/catal12121582
Submission received: 14 November 2022 / Revised: 29 November 2022 / Accepted: 1 December 2022 / Published: 5 December 2022
(This article belongs to the Special Issue Metal-Organic Framework Based Catalysts for Energy Applications)

Abstract

:
With the rapid development of modern society, environmental problems, including excessive amounts of CO2 released in the atmosphere, are becoming more and more serious. It is necessary to develop new materials and technologies to reduce pollution. Among them, metal–organic frameworks (MOFs) have shown potential for application in the area of catalysis due to their ultra-high specific surface area, structural versatility, and designability as well as ease of modification and post-synthesis. Herein, we summarize recent research advances by use of MOFs for boosting CO2 conversion. Furthermore, challenges and possible research directions related to further exploration are also discussed.

1. Introduction

As the global economy has grown at a rapid pace since the 21st century, energy and environmental issues have become increasingly prominent. In recent years, the consumption of fossil fuels has provided 85% of the world’s total energy consumption, and this proportion is likely to remain high in the years to come [1,2,3,4,5]. However, fossil fuels, for example, are being depleted, and their combustion is causing serious environmental pollution problems such as global warming [6,7,8]. Therefore, the search for green and efficient new energy sources has become a hot topic of research for scientists worldwide. Just as plants can convert CO2 in the atmosphere into organic compounds such as oxygen and sugar through photosynthesis [9,10,11,12,13], if humans can efficiently use CO2 in the air to produce small organic molecules of high energy density such as carbon monoxide, formic acid, methanol, and methane, they can both effectively reduce the accumulated CO2 in the atmosphere and the greenhouse effect, and also produce fuel and more fine chemical products, alleviating mankind’s heavy reliance on fossil fuels [14,15,16,17].
While CO2 is a major component of greenhouse gases, it is also an important renewable carbon resource. The conversion of CO2 into value-added chemicals and fuels is a promising pathway for reducing CO2 emissions and will contribute to the goal of sustainable development and carbon neutrality [18,19,20]. The soaring CO2 levels in the atmosphere due to the excessive use of fossil fuels have broken the original carbon cycle system and degraded our living environment, forcing us to urgently adjust and optimize the current energy supply structure [21]. At the same time, CO2 shows great potential for exploitation as a cheap, abundant, non-toxic, and renewable carbon feedstock [22]. The conversion of CO2 into value-added chemicals is therefore of great importance for the reduction of fossil fuel use and the achievement of carbon neutrality targets [23].
However, the CO2 reduction reaction (CO2–RR) now requires a high level of CO2 gas purity, and although direct CO2 reduction using air is a goal pursued by scientists, no related work has been reported [24,25,26,27]. In recent years, researchers have developed various methods for CO2 reduction, such as electroreduction, photoreduction, thermal reduction and bioreduction. To date, a variety of catalytic materials including graphene, graphitic carbon nitride (g-C3N4), layered oxides and hydroxides, metal–organic framework materials (MOFs) and covalent–organic framework materials (COFs) have been extensively investigated in the field of CO2 catalysis due to their unique physicochemical properties [28,29,30,31].
Metal–organic framework materials are a promising new generation of adsorbent materials due to their high specific surface area, high porosity, ease of functionalization, and adjustable structure [32,33,34]. They are a promising new generation of materials for catalytic CO2 conversion. In particular, MOFs have the following advantages as catalysts for photocatalytic or electrocatalytic CO2 reduction: (1) the structural diversity and designability of MOF components (nodes, linkers, and pores) enables catalyst design at the molecular level; (2) the high porosity of MOFs facilitates efficient transport of reactants and products across the active site and catalyst; (3) the precise structure of MOFs allows the molecular level to explore the mechanism of photocatalysis or electrocatalysis; (4) the excellent CO2 adsorption capacity of MOFs leads to high CO2 concentration around the active site, which is ideal for photocatalytic or electrocatalytic CO2 reduction. With all these advantages, MOF has been successfully applied in photocatalytic and electrocatalytic CO2 reduction. This paper provides a comprehensive review of recent advances in the research of MOF materials for CO2–RR catalysis and a brief overview of several commonly used research areas for catalytic CO2 conversion and introduces the research progress in CO2 reduction in the recent five years [35,36,37,38].

2. Strategies for CO2 Recycling

It has been found that carbon dioxide is a linear molecule with a C=O bond energy of about 750 KJ mol−1. Therefore, its conversion must involve the breaking of C=O, a process that requires a large amount of energy [39,40,41,42,43]. The literature reports that the direct reduction of CO2 involves the gain of an electron in the CO2 molecule, which changes the geometrical configuration of the CO2 molecule from linear to curved, forming a CO2 anion radical intermediate (CO2). This process requires a negative potential to be achieved (−1.9 V vs NHE), making the catalytic reduction of CO2 difficult. However, when a proton source is introduced into the reaction system, proton-assisted multi-electron transfer allows the CO2 catalytic reduction process to bypass the CO2 anion radical intermediate (Figure 1), allowing the catalytic reduction of CO2 to proceed at a lower potential [44,45,46,47,48]. While the explicit impact of pH on the HER is still an ongoing area of research, it has been thought that increasing local pH helps favor CO2–RR over HER, mainly due to the decreasing overpotential for the formation of C2+ products, as discussed earlier. On the other hand, a high local pH shifts equilibria of the acid–base reactions toward (bi)carbonates, which can reduce the concentration of CO2 near the surface, decreasing selectivity toward CO2–RR and favoring the HER instead [36]. Depending on the number of protons and electrons involved in the catalytic reduction process (Table 1), CO2 can be reduced to carbon monoxide [CO(2e)], formic acid [HCOOH(2e)], formaldehyde [HCHO(4e)], methanol [CH3OH(6e)], methane [CH4(8e)], ethylene [C2H4(12e)], ethanol [CH3CH2OH(12e)], ethane [C2H6(14e)], and propanol [C3H7OH(18e)], among other products [49,50,51,52,53,54,55,56].
For the production and final combustion of CO2 reduction/hydrogenation products to represent a closed cycle (Figure 1), the final source of electrons and protons for CO2 reduction must be water (H2O), since the combustion of hydrogenated carbon products releases H2O [4]. CO2 reduction/hydrogenation/fixation therefore follows the overall equation:
xCO2 + yH2O → product + zO2 CO2 recycling reaction
Although photosynthesis by plants and other phototrophic organisms fixes a much larger amount of carbon globally than is released by human activity, using biomass as a carbon source is an easy carbon cycling strategy. However, the use of bio-based carbon is limited by the amount of land required. Markets for biofuels such as ethanol and biodiesel require energy-intensive care (e.g., fertilisers, pesticides, etc.) and in some cases have led to deforestation, which significantly reduces the capacity for CO2 bioconversion. As mentioned earlier, biomass energy with carbon capture and storage (BECCS) has the potential to achieve net negative emissions; however, the land requirements are huge, meaning that some 14 million square kilometers of land is needed for plants to offset the current global carbon emissions of 11 GtC/year. It is clearly not feasible to reuse roughly the same area of land for global food crop production. Therefore, although important, harnessing the power of natural photosynthesis can only play a limited role in achieving net zero emissions, leaving room for other CO2 recovery strategies [57].

3. Electrochemical CO2 Reduction

3.1. Basic Issues of CO2ET

CO2 electro-chemical transformation (CO2ET) systems consist of electrodes, electrolytes, and electrolytic cells. Metal-based materials, metal-organic frameworks, and carbon materials are widely used as electrocatalysts for CO2ET due to their low cost, high conductivity, and tunable properties [58]. CO2ET electrodes are usually made from powder catalysts, polymer binders, and collectors. However, the binder usually leads to pore blockage of the catalyst, hindering the performance of the powder-based electrode due to loss of mass transfer channels [59]. Fortunately, emerging self-assembly and electrochemical deposition technologies offer stand-alone and binder-free strategies for electrode preparation with enhanced catalytic activity, improved stability, and reduced cost [60].

3.2. Construction of Electrolytic Cells

The electrolytic cell is the site where the electrocatalytic reduction of carbon dioxide takes place, and its design has a great influence on the mass transfer and thus on the efficiency of the carbon dioxide reduction. Electrolytic cells for general laboratory use fall into two main categories: H-type electrolytic cells and microfluidic flow electrolytic cells. A typical H-type electrolytic cell consists of a cathode and anode chamber and three electrodes (working electrode, reference electrode, and counter electrode). The cathode and anode chambers are separated by a proton exchange membrane or ion exchange membrane (nafion 117, etc.), which enables conduction while avoiding the diffusion of cathode products to the anode for oxidation, and the two chambers can be filled with different electrolytes so that the working electrode and counter electrode can operate under optimum conditions, respectively [61].
The working electrode and counter electrode of a microfluidic flow electrolytic cell are separated by a flowing electrolyte. The working electrode is usually a gas diffusion electrode, where the gas is slowly introduced into the cell by the diffusion layer, which ensures that the carbon dioxide is better covered by the catalyst and increases the chance of the carbon dioxide participating in the reaction. This ensures that the carbon dioxide is better covered by the catalyst and increases the chance of the carbon dioxide participating in the reaction, thus increasing the conversion rate of carbon dioxide and improving the electrolysis efficiency [22].

3.3. Type of Electrolyte

The electrolyte provides the medium of electron and proton transfer for the electrocatalytic reduction of carbon dioxide. Differences in type and concentration will greatly affect the activity and selectivity of the catalyst, and it should be noted that impurities in the electrolyte (metal ions or organic matter) in the electrolyte may poison the catalyst, causing a reduction in activity or even deactivation. In general, there are three types of electrolytes used in the laboratory: aqueous electrolytes, organic phase electrolytes, and ionic solutions electrolytes [62].

3.4. Metal–Organic Frameworks in Electrocatalytic CO2–RR

As a class of porous materials, MOFs contain both a large number of single-molecule metal junctions and structurally controlled organic groups. In contrast to conventional inorganic porous materials, MOFs are long-range ordered porous crystalline materials with a highly periodic arrangement of organic ligands and metal nodes [62,63]. In addition, the physical and chemical properties of MOFs can be adjusted by modulating the composition of the compound, metal nodes, and organic ligands at the molecular level; these excellent properties allow them to be used in a wide range of applications in electrocatalytic CO2–RR [64].

3.4.1. Fe–MOFs with Iron as the Active Centre

Due to the tunable structure and properties of porous materials, researchers have developed a series of MOFs and their derivatives as CO2–RR electrocatalysts, with the goal of designing and synthesizing highly active, selective, and stable electrocatalysts to achieve efficient electrocatalytic CO2–RR. According to relevant studies, the selection of suitable metal centres is extremely important for the synthesis of high-performance catalysts. So far, transition metals such as Fe, Co, Ni, and Cu have been good catalytic active sites in MOFs catalysts because they can form reduced or oxidised states and have good chemical stability. For example, Fe-MOFs with Fe as the active centre have shown good catalytic performance in electrocatalytic CO2–RR reactions by exploiting the Lewis acidity and redox activity of Fe [65]. Hods’ group [66] reported the electrophoretic deposition of Fe-MOF-525 films on fluorine-doped tin oxide (FTO) glass. The catalyst has a large active site, because the Lewis acid of the active center iron has good adsorption performance for CO2 molecules. It can reach a current density of 150 mA/cm2 during catalytic reduction and has a Faraday efficiency of approximately 100% for the reduction of CO2 to CO. The same Fe-PCN-222 MOFs have also been investigated for electrocatalytic CO2–RR, where the metal cluster zirconium oxygene as a linker of the MOFs does not have a catalytic active role and the porphyrin central iron is the catalytic centre, catalyzing CO2 to CO with a selectivity of 99%.

3.4.2. Co-MOFs with Cobalt as the Active Centre

Cobalt is also a well performing catalytic active site in electrocatalytic CO2–RR [67], with cobalt as the catalytic active centre mainly exploiting the diversity of cobalt valence states (Co2+, Co3+, and Co4+) [68]. To demonstrate that metallic Co also has CO2–RR properties, Yang’s group [69] prepared a kind of dispersed nano cobalt porphyrin MOFs in situ with a Faraday efficiency of about 76% CO under potentiostatic electrolytic conditions at −0.7 V (vs. RHE) and was able to maintain stability for >7 h. To further understand the process of the reduction reaction, in situ spectroelectrochemical measurements of the prepared samples were carried out (Figure 2), and the results showed that the CO2 adsorbed on the cobalt porphyrins may be subjected to a reduction reaction through the interaction of Co1+ with single electrons [70].

3.4.3. Cu-MOFs with Copper as the Active Centre

Copper is one of the most widely studied catalytic active centres because of its large CO2 molecular adsorption energy [71], which can trap other CO2 molecules to participate in the reaction while the catalytic reaction is incomplete, resulting in the production of polycarbonate products. The onset potential for CO2 reduction of Cu-MOF’s electrode occurred at potentials approximately 0.2 V more positive than those observed with a Cu metal electrode. The Hing group [72] prepared a Cu-MOF material by mixing red ammonia with aqueous copper sulphate solution and applied it to electrocatalytic CO2–RR. The authors found that in a 0.5 M KHCO3 solution, the MOF material with Cr–Mn as the metal node had an onset potential around 0.2 V (vs. RHE) higher than that of the Cu metal electrode when catalysed under the same conditions. To explain this phenomenon, the authors show by means of density flooding (DFT) calculations that the reason for this phenomenon is mainly due to the lower electron cloud density of copper, which leads to a stronger adsorption of CO2 onto Cu-MOFs and a higher amount of CO2 per active site, thus exhibiting a lower onset potential for Cu-MOFs. Since then, MOFs materials with copper as the catalytic active centre have been widely used in electrocatalytic CO2–RR reactions. In subsequent studies, researchers have used the specific adsorption of metallic copper on CO2 molecules to generate polycarbonate products in electrocatalytic CO2–RR reactions. A typical MOFs catalyst is Cu3(BTC)2, generated using benzoic acid as the organic ligand, and experimental results show that this copper-based MOF can effectively catalyse the reduction of CO2 to 90% pure oxalic acid with a Faraday efficiency of 51% [73]. Yang and co-workers [74] also reported an IL-templated electro-synthesis of MOF [Cu2(L)] on a copper electrode. This integrated electrode consisting of a thin dense coating of Cu2(L) on Cu foam incorporates the uncoupled Cu(II) active site and shows high conductivity and stability (Figure 3). This electrode shows excellent activity for the electro-reduction of CO2 to formic acid, with a low onset potential of −1.45 V vs. Ag/Ag+, and the FEHCOOH reaches 90.5% at −1.80 V vs. Ag/Ag+ with a current density of 65.8 mA·cm−2. Experimental (EPR spectroscopy) and theoretical (DFT) methods confirm that the reaction is driven by defects within the structure of the decorated electrode.

3.4.4. Other Metal as the Active Centers

MOFs using metallic nickel as the catalytic active site have also been extensively investigated. For example, Jiang’s group [75] has proposed a post-ion exchange strategy to successfully obtain low-coordinated, single-atom Ni catalysts based on Zn-based MOFs (ZIF-8) using a two-step ion-exchange strategy (Figure 4). The material morphology can be well maintained from the ZIF-8 precursor to Zn-N3-C and then to the Ni-N3-C material obtained by ion exchange. The MOFs with metallic nickel as the catalytic active centre were experimentally verified to have a Faraday efficiency of 95% in catalyzing the formation of CO from CO2. Cao’s group has constructed nickel-based MOFs with phthalocyanine as the organic ligand; the Faraday efficiency of the catalytic CO generation can also reach 96.4%, and the material can be electrolysed for 10 h with almost constant activity [76].

3.4.5. Application of Porphyrin-Based Metal–Organic Frameworks in Electrocatalytic CO2–RR

MOF-based catalysts play a good catalytic role in electrocatalytic CO2–RR as a class of porous materials, where the key influence is the choice of organic ligands and metal nodes. As an important organic group in nature, porphyrins are heterocyclic molecules consisting of four pyrroles linked by hypomethyles that can carry different metals in their centres and thus have a variety of properties. Metal porphyrins and their derivatives are commonly found in nature and in organelles associated with the human body, and play an important role (Table 2). The use of metalloporphyrins as molecular catalysts has been widely studied for small molecule activation reactions and has the following advantages in terms of coordination chemistry. First, the rigid and robust porphyrin backbone can be systematically designed and synthesized based on the porphyrin framework; second, many functional groups can be on the porphyrin backbone, and these play a key role not only in improving catalysis but also in controlling the framework structure and morphology; third, porphyrins can be used as structural units and catalytic sites in frameworks based on MOFs. Bimetallic and polymetallic porphyrin-based frameworks can be used to construct catalysts for synergistic catalytic effects. Thus, the application of porphyrins as organic ligands to MOF materials has led to important advances in electrocatalytic CO2–RR.
PCN-222 showed excellent electrocatalytic CO2–RR performance. For example, Professor Jiang’s group chose a MOF (PCN-222) constructed from porphyrin tetracarboxylic acid ligands and zirconium ions for electrocatalytic CO2–RR and found that the Faraday efficiency of catalytic CO2 to CO reached 87%, while the catalytic performance did not decrease after 10 h of continuous electrolysis [77]. Lan and co-workers [78] have proposed a new strategy to insert metal clusters into porphyrin-based backbones by chemical vapour deposition (CVD) to achieve electrocatalytic CO2–RR selectivity. By optimizing the experimental conditions, the resulting complexes have a Faraday efficiency of 97% for CO at −0.7 V (vs. RHE) using MOF-545-Co as the backbone. Not only is the introduction of metal clusters an electron donor and carrier, but also the strong binding interactions between the metal clusters and the metal active site of the porphyrin centre reduce the CO2 adsorption energy.
Table 2. MOFs with active metal nodes for CO2 reduction.
Table 2. MOFs with active metal nodes for CO2 reduction.
MOFMetal NodeProductReference
NH2-UiO-66 (Zr/Ti)(Zr/Ti)6O4(OH)4formate[79]
MIL-101 (Fe)Fe–O clusterformate[80]
NNU-13Zn-ε-Keggin clusterCH4[81]
Co-ZIF-9Co–O clusterCO[82]
MOF-NiNi clusterCO[83]
BIF-29Cu clusterCO[84]
Ni MOLsNi2(OH)2CH3OH[85]

4. Photocatalytic CO2 Reduction

Photocatalysis is an efficient, low-energy, clean, and non-secondary pollution-free technology [86,87,88,89,90]. The three important parameters of a photocatalyst are: (1) the spectral response range; (2) the effective separation efficiency of electrons and holes; (3) the redox reaction efficiency of the charge carrier. Organic ligands provide the highest occupied molecular orbital (HOMO), and metal centres provide the lowest unoccupied molecular orbital (LUMO), so that metallic crystalline materials can exhibit similar semiconductor behavior [91,92,93]. Irradiation of MOFs by a certain light source can lead to metal charge transfer (LMCT) of the organic ligands, where electrons from the organic ligands are transferred to the metal centres and electron and hole separation occurs [94,95,96,97]. The photocatalytic activity of MOFs is further modulated by improving the photo-response range of the ligand. As a result, MOFs and MOF derivatives are rapidly being developed for photocatalytic applications, widely used in the photocatalytic reduction of CO2 [82,98,99].

4.1. Metal Porphyrin Ligands for MOF Photocatalysts

Metalloporphyrins are widely used in photocatalytic reactions due to their broad absorption spectra [100]. Ye et al. [101] prepared highly catalytically active MOF-525-Co by post-synthetic modification using MOF-525 as the parent body and found that under visible light irradiation (Figure 5), the porphyrins were excited by light absorption and photogenerated electrons were transferred to the Co2+ present in the MOF in the form of single atoms, reducing it to form the true catalytic centre; Co+ is subsequently oxidised to Co2+ by CO2 molecules, completing a catalytic cycle in which CO2 is reduced to carbon monoxide and methane; The intermediates of the whole process can be captured by electron paramagnetic resonance [101]. The introduction of Co2+ was found to promote the separation of photogenerated charges and thus enhance the catalytic activity, with an increase of about six-fold in the rate of methane production and about three-fold in the rate of carbon monoxide production for MOF-52-Co compared to the parent.
Zhang et al. [63] used cobalt porphyrin derivatives containing polyphenolic groups to coordinate with zirconium oxygen clusters to generate ZrPP-1 and ZrPP-2. Unlike other carboxy porphyrin MOFs, this MOF utilises -OH for coordination and is therefore highly acid and base resistant. The carbon dioxide adsorption of ZrPP-1 was found to be as high as 90 cm3g−1, with a carbon monoxide production rate of 14 μmol g−1 h−1 under visible light (λ > 420 nm) irradiation. Based on theoretical calculations and experimental data, The researchers concluded that ZrPP-1 is excited by the absorption of light by the porphyrin ligand under visible light irradiation, and the excitation electrons are transferred to Co2+, followed by a continuous quenching process that generates Co0, which can activate carbon dioxide to form Co2+-CO and eventually release carbon monoxide, completing a catalytic cycle. Jiang et al. [77] reported the use of a MOF (PCN-222) formed by zirconium oxygen clusters and tetracarboxy porphyrins for the photocatalytic reduction of carbon dioxide to formic acid. The ultrafast transient absorption spectra showed that the zirconium oxygen clusters acted as a catalytic centre to inhibit the complexation of electrons and holes, thus enhancing the catalytic activity.

4.2. Bipyridyl Metal Complex Ligands for MOF Photocatalysts

Bipyridyl metal complexes are widely used in the study of photo(electro)-catalytic CO2 reduction, but the stability of molecular catalysts is poor, and making them into MOFs can significantly improve the stability and catalytic activity. Yang et al. used dicarboxylated bipyridyl rhenium complexes (H2ReTC) to replace some of the biphenyldicarboxylic acid (H2BPDC) to generate the hybrid UIO-67 MOF (Figure 6). The presence of too many H2ReTC ligands in the MOF cell increases the collision between ligands, which is detrimental to the catalytic activity and reduces the specific surface area of the MOF, leading to the obstruction of the diffusion of reaction substrates and products; when silver nanoparticles with surface plasmon resonance effect are encapsulated in Re3-MOF, the catalytic activity can be increased by 7 times [102].
Lin et al. [103] grafted the catalytically active M(bpy)(CO)3X (M = Re or Mn) onto a Hf12-Ru monolayer MOF with photosensitive properties by means of carboxylic acid ligand exchange to form a Hf12-Ru-M hybrid MOF with both photosensitivity and catalytic activity, and since the distance between the photosensitive and catalytic centres is only 1.3 nm, photogenerated electrons can be quickly transferred to the catalytic site to participate in the reaction, so this MOF shows excellent catalytic activity: the conversion number of Hf12-RuReMOF is 3500 and that of Hf12-Ru-MnMOF is 1000 after 25 h of irradiation by artificial light, and more excitingly, the conversion number of Hf12-RuReMOF reaches 670 after 6 h of irradiation by natural light.

4.3. MOF Photocatalysts Containing the -NH2 Functional Group

Li et al. [104] formed NH2-MIL-125(Ti) from amino terephthalic acid (NH2-BDC) and titanium oxide clusters. Since amine sites are highly effective for CO2 adsorption and can act as light-absorption chromophores, NH2-MIL-125(Ti) produced 8 μmol of formic acid, while MIL-125(Ti) produced almost no product. The UV-vis absorption spectra indicated that the introduction of amino groups in the ligand could broaden the absorption band of MOF to 550 nm, thus increasing the energy uptake of the reaction system. Meanwhile, the introduction of amino groups on the benzene ring could significantly increase the carbon dioxide adsorption of MOF, thus leading to the high catalytic activity of NH2-MIL-125(Ti) [105]. When the amount of amino groups introduced on the benzene ring increases, the reaction activity is further enhanced. The group partially replaced 2-amino-terephthalic acid with 2,5-diamino-terephthalic acid to form a mixed ligand NH2-UIO-66 (Zr) (mixed NH2-UIO-66 (Zr)), and experiments revealed that mixed NH2-UIO-66 (Zr) catalyzed the reduction of carbon dioxide formic acid, which was found to be more active than NH2-UIO-66 (Zr), which was analysed to indicate that this was due to the introduction of more amino groups to broaden the absorption spectrum of MOF while further increasing the amount of CO2 adsorbed, leading to a stronger catalytic activity.

4.4. Others

Ligands with special characteristics, such as large conjugated group and high visible-light harvesting abilities, have also been applied for constructing MOFs with photocatalytic CO2–RR properties. Recently Lan et al. [106] reported a two-in-one strategy for efficient CO2–RR. Two eco-friendly biomimetic MOFs were prepared for their study ([Co2(AD)4H2(BDA)] (ADMOF-1) and [Co2(AD)4H2(IBA)2] (AD-MOF-2)). First, the -NH2 of HAD ligands acted as auxochromic groups to make MOFs responsive to visible light. Second, the uncoordinated aromatic nitrogen of HAD (on the o-position of -NH2) acted as an active position for CO2–RR. Photocatalytic CO2–RR was carried out in pure aqueous solution without any other photosensitizer. Under the visible-light irradiation for 5 h, HCOOH production over AD-MOF-2 was detected to be as high as 443.2 mol g−1 h−1. These results indicated that the two-in-one strategy efficiently improved the photocatalytic CO2-RR performance of the MOFs.

4.5. The Dynamic Role/Mechanism of MOFs Works as a Catalyst

Up to now, the examples of CO2 reduction based on MOF-based photocatalysts are still insufficient to elucidate the structure–activity relationship and reaction mechanism in the photocatalytic process. The active centers, such as open metal sites, are required to be chosen judiciously to achieve efficient activation of CO2 molecules. For heterogeneous reactions, CO2 molecules generally need to diffuse into the channels of catalysts to increase the utilization of the active sites within the solid.

5. Conclusions

In summary, MOF-based materials have important applications in the conversion of CO2, as the MOF catalysts contain transition metal sites in the porous structures, giving them a rich arrangement of electrons outside the nucleus, which can be used to change the reaction path during the reaction and thus reduce the activation energy. The porous framework of the MOF material can be used as a good support structure, allowing it to be compounded with other materials to form a wider variety of new materials. Due to the regular structure of MOF materials, calcination of MOF materials as precursors can lead to inorganic nanomaterials with special morphology and chemical composition, and the carbon elements in MOF can also form carbon coating films during calcination to obtain non-metal doped composites.
Although great progress has been made in the application of MOF-based materials in the catalytic conversion of CO2, there are certain limitations. First, the separation, transport, and complexation of photogenerated carriers (electrons and holes) in the photocatalytic process has always been a problem across all types of materials. A good photocatalyst should have the advantage of a high number of photogenerated carriers, fast transport rates, and less complexation during the transport of carriers from the catalyst interior to the surface. Therefore, there is room for further development of MOF-based materials. The doping of MOF-based materials with photosensitive components can effectively improve the response of the material to light. For example, ligands containing strong conjugated groups have a certain absorption in the visible region, and the introduction of such ligands or molecules can lead to photocatalytic activity of the material under visible light irradiation.
Constructing heterojunctions can improve the separation and transport efficiency of photogenerated carriers. Changing the morphology of the material and shortening the transport distance of photogenerated carriers can effectively control the compounding of electrons and holes during the transport process. Changes in the chemical composition of the catalyst surface can enhance the interaction between the material and CO2, and the fabrication of more pore structures can expand the specific surface area of the material and expose more catalytic active sites to the reaction system, thus improving the catalytic efficiency of the material. Secondly, the stability and recyclability of the catalyst material is also an important indicator of the quality of the catalyst, especially in liquid phase reaction systems where the MOF structure is somewhat unstable in the presence of other Lewis acids and bases. For photocatalytic CO2 reduction, many current systems require the use of organic sacrificial agents (e.g., TEOA) to provide electrons, whereas a more ideal system would use water as the electron donor. Once again, MOF-based materials have a relatively low product generation rate and are not sufficiently selective to be used industrially. Finally, the economic value and environmental friendliness of the catalyst material itself are also issues that need to be considered for the industrial application of MOF-based materials.
There are many important issues to be solved in the field of MOF-based materials for catalytic CO2 conversion, and the future development directions are mainly in the following two areas:
(1)
design and construct MOF-based materials with high activity, high selectivity, high stability, low production cost, and low toxicity to achieve efficient and green conversion of CO2 under mild conditions;
(2)
further broaden the types of reactions for CO2 conversion by MOF-based materials, develop new ways of CO2 conversion, and convert CO2 into a variety of high value-added chemicals.

Author Contributions

Writing—original draft preparation, Y.Q. and X.Y.; writing—review and editing, F.H. and Z.D.; resources, project administration, funding acquisition, Z.D. All authors have read and agreed to the published version of the manuscript.

Funding

Thanks for the financial support from the National Natural Science Foundation of China (No. 22205163), Natural Science Foundation of Shandong Province (No. ZR2021QB078), State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, and the Doctoral Program Foundation of Tianjin Normal University (043135202-XB1704).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Elhenawy, S.E.M.; Khraisheh, M.; AlMomani, F.; Walker, G. Metal-Organic Frameworks as a Platform for CO2 Capture and Chemical Processes: Adsorption, Membrane Separation, Catalytic-Conversion, and Electrochemical Reduction of CO2. Catalysts 2020, 10, 1293. [Google Scholar] [CrossRef]
  2. Yang, Q.; Xu, Q.; Jiang, H.L. Metal-organic frameworks meet metal nanoparticles: Synergistic effect for enhanced catalysis. Chem. Soc. Rev. 2017, 46, 4774–4808. [Google Scholar] [CrossRef] [PubMed]
  3. Zhou, Z.; He, C.; Xiu, J.; Yang, L.; Duan, C. Metal-Organic Polymers Containing Discrete Single-Walled Nanotube as a Heterogeneous Catalyst for the Cycloaddition of Carbon Dioxide to Epoxides. J. Am. Chem. Soc. 2015, 137, 15066–15069. [Google Scholar] [CrossRef] [PubMed]
  4. Li, W.; Wang, H.; Jiang, X.; Zhu, J.; Liu, Z.; Guo, X.; Song, C. A short review of recent advances in CO2 hydrogenation to hydrocarbons over heterogeneous catalysts. RSC Adv. 2018, 8, 7651–7669. [Google Scholar] [CrossRef] [Green Version]
  5. Kang, X.; Wang, B.; Hu, K.; Lyu, K.; Han, X.; Spencer, B.F.; Frogley, M.D.; Tuna, F.; McInnes, E.J.L.; Dryfe, R.A.W.; et al. Quantitative Electro-Reduction of CO2 to Liquid Fuel over Electro-Synthesized Metal-Organic Frameworks. J. Am. Chem. Soc. 2020, 142, 17384–17392. [Google Scholar] [CrossRef]
  6. Agatemor, C.; Quintana, A.A.; Sztapka, L.M.; Ebinuma, V.C.S. Enabling Sustainable Chemistry with Ionic Liquids and Deep Eutectic Solvents: A Fad or the Future? Angew. Chem. Int. Ed. 2022, 134, e202205609. [Google Scholar]
  7. Wang, S.; Guan, B.Y.; Lu, Y.; Lou, X.W.D. Formation of Hierarchical In2S3-CdIn2S4 Heterostructured Nanotubes for Efficient and Stable Visible Light CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 17305–17308. [Google Scholar] [CrossRef]
  8. Xu, Q.; Zhang, L.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Jaroniec, M. Direct Z-scheme photocatalysts: Principles, synthesis, and applications. Mater. Today 2018, 21, 1042–1063. [Google Scholar] [CrossRef]
  9. Whang, H.S.; Lim, J.; Choi, M.S.; Lee, J.; Lee, H. Heterogeneous catalysts for catalytic CO2 conversion into value-added chemicals. BMC Chem. Eng. 2019, 1, 9. [Google Scholar] [CrossRef]
  10. Liu, S.; Chen, F.; Li, S.; Peng, X.; Xiong, Y. Enhanced photocatalytic conversion of greenhouse gas CO2 into solar fuels over g-C3N4 nanotubes with decorated transparent ZIF-8 nanoclusters. Appl. Catal. B Environ. 2017, 211, 1–10. [Google Scholar] [CrossRef]
  11. Chen, S.; Yu, J.; Zhang, J. Enhanced photocatalytic CO2 reduction activity of MOF-derived ZnO/NiO porous hollow spheres. J. CO2 Util. 2018, 24, 548–554. [Google Scholar] [CrossRef]
  12. Dang, S.; Gao, P.; Liu, Z.; Chen, X.; Yang, C.; Wang, H.; Zhong, L.; Li, S.; Sun, Y. Role of zirconium in direct CO2 hydrogenation to lower olefins on oxide/zeolite bifunctional catalysts. J. Catal. 2018, 364, 382–393. [Google Scholar] [CrossRef]
  13. Fujiwara, M.; Ando, H.; Tanaka, M.; Souma, Y. Hydrogenation of carbon dioxide over Cu-Zn-chromate/zeolite composite catalyst: The effects of reaction behavior of alkenes on hydrocarbon synthesis. Appl. Catal. A Gen. 1995, 130, 105–116. [Google Scholar] [CrossRef]
  14. Di, Z.; Pang, J.; Hu, F.; Wu, M.; Hong, M. An ultra-stable microporous supramolecular framework with highly selective adsorption and separation of water over ethanol. Nano. Res. 2021, 14, 2584–2588. [Google Scholar] [CrossRef]
  15. Goswami, L.; Kushwaha, A.; Kafle, S.R.; Kim, B.-S. Surface Modification of Biochar for Dye Removal from Wastewater. Catalysts 2022, 12, 817. [Google Scholar] [CrossRef]
  16. Ren, F.; Ji, P. Recent Advances in the Application of Metal–Organic Frameworks for Polymerization and Oligomerization Reactions. Catalysts 2020, 10, 1141. [Google Scholar] [CrossRef]
  17. Fujlwara, M.; Kieffer, R.; Ando, H.; Xu, Q.; Souma, Y. Change of catalytic properties of Fe-ZnO/zeolite composite catalyst in the hydrogenation of carbon dioxide. Appl. Catal. A Gen. 1997, 154, 87–101. [Google Scholar] [CrossRef]
  18. Gao, J.; Jia, C.; Liu, B. Direct and selective hydrogenation of CO2 to ethylene and propene by bifunctional catalysts. Catal. Sci. Technol. 2017, 7, 5602–5607. [Google Scholar] [CrossRef]
  19. Gao, P.; Dang, S.; Li, S.; Bu, X.; Liu, Z.; Qiu, M.; Yang, C.; Wang, H.; Zhong, L.; Han, Y.; et al. Direct Production of Lower Olefins from CO2 Conversion via Bifunctional Catalysis. ACS Catal. 2017, 8, 571–578. [Google Scholar] [CrossRef]
  20. Goyal, S.; Shaharun, M.; Kait, C.; Abdullah, B.; Ameen, M. Photoreduction of Carbon Dioxide to Methanol over Copper Based Zeolitic Imidazolate Framework-8: A New Generation Photocatalyst. Catalysts 2018, 8, 581. [Google Scholar] [CrossRef] [Green Version]
  21. Jia, S.; Ma, X.; Sun, X.; Han, B. Electrochemical Transformation of CO2 to Value-Added Chemicals and Fuels. CCS Chem. 2022, 4, 3213–3229. [Google Scholar] [CrossRef]
  22. Jin, S.; Hao, Z.; Zhang, K.; Yan, Z.; Chen, J. Advances and Challenges for the Electrochemical Reduction of CO2 to CO: From Fundamentals to Industrialization. Angew. Chem. Int. Ed. 2021, 60, 20627–20648. [Google Scholar] [CrossRef] [PubMed]
  23. Di, Z.; Zheng, X.; Qi, Y.; Yuan, H.; Li, C.-P. Recent Advances in C2 Gases Separation and Purification by Metal-Organic Frameworks. Chin. J. Struct. Chem. 2022, 41, 2211031–2211044. [Google Scholar]
  24. Fang, Z.; Ren, R.; Wang, Y.; Hu, Y.; Dong, M.; Ye, Z.; He, Q.; Peng, X. Solar-driven all-in-one MOFs-based catalyst for highly efficient CO2 conversion. Appl. Catal. B Environ. 2022, 318, 121878. [Google Scholar] [CrossRef]
  25. Zhang, A.A.; Si, D.; Huang, H.; Xie, L.; Fang, Z.B.; Liu, T.F.; Cao, R. Partial Metalation of Porphyrin Moieties in Hydrogen-Bonded Organic Frameworks Provides Enhanced CO2 Photoreduction Activity. Angew. Chem. Int. Ed. 2022, 61, e202203955. [Google Scholar]
  26. Han, Y.; Xu, H.; Su, Y.; Xu, Z.L.; Wang, K.; Wang, W. Noble metal (Pt, Au@Pd) nanoparticles supported on metal organic framework (MOF-74) nanoshuttles as high-selectivity CO2 conversion catalysts. J. Catal. 2019, 370, 70–78. [Google Scholar] [CrossRef]
  27. Helal, A.; Cordova, K.E.; Arafat, M.E.; Usman, M.; Yamani, Z.H. Defect-engineering a metal–organic framework for CO2 fixation in the synthesis of bioactive oxazolidinones. Inorg. Chem. Front. 2020, 7, 3571–3577. [Google Scholar] [CrossRef]
  28. Luo, H.; Gu, Y.; Liu, D.; Sun, Y. Advances in Oxidative Desulfurization of Fuel Oils over MOFs-Based Heterogeneous Catalysts. Catalysts 2021, 11, 1557. [Google Scholar] [CrossRef]
  29. Qiu, X.F.; Huang, J.R.; Yu, C.; Zhao, Z.H.; Zhu, H.L.; Ke, Z.; Liao, P.Q.; Chen, X.M. A Stable and Conductive Covalent Organic Framework with Isolated Active Sites for Highly Selective Electroreduction of Carbon Dioxide to Acetate. Angew. Chem. Int. Ed. 2022, 61, e202206470. [Google Scholar]
  30. Wang, H.; Dong, J.; Allard, L.F.; Lee, S.; Oh, S.; Wang, J.; Li, W.; Shen, M.; Yang, M. Single-site Pt/La-Al2O3 stabilized by barium as an active and stable catalyst in purifying CO and C3H6 emissions. Appl. Catal. B Environ. 2019, 244, 327–339. [Google Scholar] [CrossRef]
  31. Overa, S.; Crandall, B.S.; Shrimant, B.; Tian, D.; Ko, B.H.; Shin, H.; Bae, C.; Jiao, F. Enhancing acetate selectivity by coupling anodic oxidation to carbon monoxide electroreduction. Nat. Catal. 2022, 5, 738–745. [Google Scholar] [CrossRef]
  32. Di, Z.; Liu, C.; Pang, J.; Zou, S.; Ji, Z.; Hu, F.; Chen, C.; Yuan, D.; Hong, M.; Wu, M. A Metal-Organic Framework with Nonpolar Pore Surfaces for the One-step Acquisition of C2H4 from a C2H4 and C2H6 Mixture. Angew. Chem. Int. Ed. 2022, 61, e202210343. [Google Scholar] [CrossRef] [PubMed]
  33. Li, S.-L.; Xu, Q. Metal–organic frameworks as platforms for clean energy. Energy Environ. Sci. 2013, 6, 1656. [Google Scholar] [CrossRef]
  34. Hu, S.; Liu, M.; Ding, F.; Song, C.; Zhang, G.; Guo, X. Hydrothermally stable MOFs for CO2 hydrogenation over iron-based catalyst to light olefins. J. CO2 Util. 2016, 15, 89–95. [Google Scholar] [CrossRef]
  35. Valdebenito, G.; Gonzaléz-Carvajal, M.; Santibañez, L.; Cancino, P. Metal–Organic Frameworks (MOFs) and Materials Derived from MOFs as Catalysts for the Development of Green Processes. Catalysts 2022, 12, 136. [Google Scholar] [CrossRef]
  36. Nitopi, S.; Bertheussen, E.; Scott, S.B.; Liu, X.; Engstfeld, A.K.; Horch, S.; Seger, B.; Stephens, I.E.L.; Chan, K.; Hahn, C.; et al. Progress and Perspectives of Electrochemical CO2 Reduction on Copper in Aqueous Electrolyte. Chem. Rev. 2019, 119, 7610–7672. [Google Scholar] [CrossRef] [Green Version]
  37. Maina, J.W.; Schutz, J.A.; Grundy, L.; Des Ligneris, E.; Yi, Z.; Kong, L.; Pozo-Gonzalo, C.; Ionescu, M.; Dumee, L.F. Inorganic Nanoparticles/Metal Organic Framework Hybrid Membrane Reactors for Efficient Photocatalytic Conversion of CO2. ACS Appl. Mater. Interfaces 2017, 9, 35010–35017. [Google Scholar] [CrossRef]
  38. Porosoff, M.D.; Yan, B.; Chen, J.G. Catalytic reduction of CO2 by H2 for synthesis of CO, methanol and hydrocarbons: Challenges and opportunities. Energy Environ. Sci. 2016, 9, 62–73. [Google Scholar] [CrossRef]
  39. Ziemba, M.; Radtke, M.; Schumacher, L.; Hess, C. Elucidating CO2 Hydrogenation over In2O3 Nanoparticles using Operando UV-vis and Impedance Spectroscopies. Angew. Chem. Int. Ed. 2022, 61, e202209388. [Google Scholar] [CrossRef]
  40. Zang, Y.; Liu, T.; Wei, P.; Li, H.; Wang, Q.; Wang, G.; Bao, X. Selective CO2 Electroreduction to Ethanol over Carbon-Coated CuOx Catalyst. Angew. Chem. Int. Ed. 2022, 61, e202209629. [Google Scholar] [CrossRef]
  41. Wang, G.; Chen, Z.; Wang, T.; Wang, D.; Mao, J. P and Cu Dual Sites on Graphitic Carbon Nitride for Photocatalytic CO2 Reduction to Hydrocarbon Fuels with High C2H6 Evolution. Angew. Chem. Int. Ed. 2022, 61, e202210789. [Google Scholar]
  42. Santibáñez, L.; Escalona, N.; Torres, J.; Kremer, C.; Cancino, P.; Spodine, E. CuII- and CoII-Based MOFs: {[La2Cu3(µ-H2O)(ODA)6(H2O)3]∙3H2O}n and {[La2Co3(ODA)6(H2O)6]∙12H2O}n. The Relevance of Physicochemical Properties on the Catalytic Aerobic Oxidation of Cyclohexene. Catalysts 2020, 10, 589. [Google Scholar] [CrossRef]
  43. Tarasov, A.L.; Isaeva, V.I.; Tkachenko, O.P.; Chernyshev, V.V.; Kustov, L.M. Conversion of CO2 into liquid hydrocarbons in the presence of a Co-containing catalyst based on the microporous metal-organic framework MIL-53(Al). Fuel. Proc. Tech. 2018, 176, 101–106. [Google Scholar] [CrossRef]
  44. Sun, X.; Sun, L.; Li, G.; Tuo, Y.; Ye, C.; Yang, J.; Low, J.; Yu, X.; Bitter, J.H.; Lei, Y.; et al. Phosphorus Tailors the d-Band Center of Copper Atomic Sites for Efficient CO2 Photoreduction under Visible-Light Irradiation. Angew. Chem. Int. Ed. 2022, 61, e202207677. [Google Scholar] [CrossRef] [PubMed]
  45. Gomez-Oliveira, E.P.; Reinares-Fisac, D.; Aguirre-Diaz, L.M.; Esteban-Betegon, F.; Pintado-Sierra, M.; Gutierrez-Puebla, E.; Iglesias, M.; Monge, A.; Gandara, F. Framework Adaptability and Concerted Structural Response in a Bismuth Metal-Organic Framework Catalyst. Angew. Chem. Int. Ed. 2022, 61, e202209335. [Google Scholar] [CrossRef]
  46. An, B.; Zhang, J.; Cheng, K.; Ji, P.; Wang, C.; Lin, W. Confinement of Ultrasmall Cu/ZnO Nanoparticles in Metal-Organic Frameworks for Selective Methanol Synthesis from Catalytic Hydrogenation of CO. J. Am. Chem. Soc. 2017, 139, 3834–3840. [Google Scholar] [CrossRef]
  47. Becerra, J.; Nguyen, D.-T.; Gopalakrishnan, V.-N.; Do, T.-O. Plasmonic Au Nanoparticles Incorporated in the Zeolitic Imidazolate Framework (ZIF-67) for the Efficient Sunlight-Driven Photoreduction of CO2. ACS Appl. Energy Mater. 2020, 3, 7659–7665. [Google Scholar] [CrossRef]
  48. Cancino, P.; Santibañez, L.; Stevens, C.; Fuentealba, P.; Audebrand, N.; Aravena, D.; Torres, J.; Martinez, S.; Kremer, C.; Spodine, E. Influence of the channel size of isostructural 3d–4f MOFs on the catalytic aerobic oxidation of cycloalkenes. New J. Chem. 2019, 43, 11057–11064. [Google Scholar] [CrossRef]
  49. Lu, S.; Lou, F.; Yu, Z. Recent Progress in Two-Dimensional Materials for Electrocatalytic CO2 Reduction. Catalysts 2022, 12, 228. [Google Scholar] [CrossRef]
  50. Mahlaule-Glory, L.M.; Hintsho-Mbita, N.C. Green Derived Zinc Oxide (ZnO) for the Degradation of Dyes from Wastewater and Their Antimicrobial Activity: A Review. Catalysts 2022, 12, 833. [Google Scholar] [CrossRef]
  51. Lv, J.J.; Yin, R.; Zhou, L.; Li, J.; Kikas, R.; Xu, T.; Wang, Z.J.; Jin, H.; Wang, X.; Wang, S. Microenvironment Engineering for the Electrocatalytic CO2 Reduction Reaction. Angew. Chem. Int. Ed. 2022, 61, e202207252. [Google Scholar] [CrossRef] [PubMed]
  52. Grammatico, D.; Bagnall, A.J.; Riccardi, L.; Fontecave, M.; Su, B.L.; Billlon, L. Heterogenised molecular catalysts for sustainable electrochemical CO2 reduction. Angew. Chem. Int. Ed. 2022, 61, e202206399. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, T.; Shi, L.; Tang, J.; Malgras, V.; Asahina, S.; Liu, G.; Zhang, H.; Meng, X.; Chang, K.; He, J.; et al. A Co3O4 embedded porous ZnO rhombic dodecahedron prepared using zeolitic imidazolate frameworks as precursors for CO2 photoreduction. Nanoscale 2016, 8, 6712–6720. [Google Scholar] [CrossRef]
  54. Wei, J.; Ge, Q.; Yao, R.; Wen, Z.; Fang, C.; Guo, L.; Xu, H.; Sun, J. Directly converting CO2 into a gasoline fuel. Nat. Commun. 2017, 8, 15174. [Google Scholar] [CrossRef] [Green Version]
  55. Zhang, G.; Wang, Z.; Wu, J. Construction of a Z-scheme heterojunction for high-efficiency visible-light-driven photocatalytic CO2 reduction. Nanoscale 2021, 13, 4359–4389. [Google Scholar] [CrossRef] [PubMed]
  56. Zhang, H.; Wang, T.; Wang, J.; Liu, H.; Dao, T.D.; Li, M.; Liu, G.; Meng, X.; Chang, K.; Shi, L.; et al. Surface-Plasmon-Enhanced Photodriven CO2 Reduction Catalyzed by Metal-Organic-Framework-Derived Iron Nanoparticles Encapsulated by Ultrathin Carbon Layers. Adv. Mater. 2016, 28, 3703–3710. [Google Scholar] [CrossRef]
  57. Zhao, Z.-W.; Zhou, X.; Liu, Y.-N.; Shen, C.-C.; Yuan, C.-Z.; Jiang, Y.-F.; Zhao, S.-J.; Ma, L.-B.; Cheang, T.-Y.; Xu, A.-W. Ultrasmall Ni nanoparticles embedded in Zr-based MOFs provide high selectivity for CO2 hydrogenation to methane at low temperatures. Catal. Sci. Technol. 2018, 8, 3160–3165. [Google Scholar] [CrossRef]
  58. Yi, J.-D.; Xie, R.; Xie, Z.-L.; Chai, G.-L.; Liu, T.-F.; Chen, R.-P.; Huang, Y.-B.; Cao, R. Highly Selective CO2 Electroreduction to CH4 by In Situ Generated Cu2O Single-Type Sites on a Conductive MOF: Stabilizing Key Intermediates with Hydrogen Bonding. Angew. Chem. Int. Ed. 2020, 59, 23641–23648. [Google Scholar] [CrossRef]
  59. Zhang, M.; Lin, Q.; Wu, W.; Ye, Y.; Yao, Z.; Ma, X.; Xiang, S.; Zhang, Z. Isostructural MOFs with Higher Proton Conductivity for Improved Oxygen Evolution Reaction Performance. ACS Appl. Mater. Interfaces 2020, 12, 16367–16375. [Google Scholar] [CrossRef]
  60. Zhou, Z.; He, C.; Yang, L.; Wang, Y.; Liu, T.; Duan, C. Alkyne Activation by a Porous Silver Coordination Polymer for Heterogeneous Catalysis of Carbon Dioxide Cycloaddition. ACS Catal. 2017, 7, 2248–2256. [Google Scholar] [CrossRef]
  61. Gong, Y.N.; Jiao, L.; Qian, Y.; Pan, C.Y.; Zheng, L.; Cai, X.; Liu, B.; Yu, S.H.; Jiang, H.L. Regulating the Coordination Environment of MOF-Templated Single-Atom Nickel Electrocatalysts for Boosting CO2 Reduction. Angew. Chem. Int. Ed. 2020, 59, 2705–2709. [Google Scholar] [CrossRef] [PubMed]
  62. Li, X.; Zhu, Q.-L. MOF-based materials for photo- and electrocatalytic CO2 reduction. EnergyChem 2020, 2, 100033. [Google Scholar] [CrossRef]
  63. Luo, Y.-H.; Dong, L.-Z.; Liu, J.; Li, S.-L.; Lan, Y.-Q. From molecular metal complex to metal-organic framework: The CO2 reduction photocatalysts with clear and tunable structure. Coord. Chem. Rev. 2019, 390, 86–126. [Google Scholar] [CrossRef]
  64. Liu, C.; Wang, W.; Liu, B.; Qiao, J.; Lv, L.; Gao, X.; Zhang, X.; Xu, D.; Liu, W.; Liu, J.; et al. Recent Advances in MOF-based Nanocatalysts for Photo-Promoted CO2 Reduction Applications. Catalysts 2019, 9, 658. [Google Scholar] [CrossRef] [Green Version]
  65. Bhugun, I.; Lexa, D.; Saveant, J.-M. Ultraefficient Selective Homogeneous Catalysis of the Electrochemical Reduction of Carbon Dioxide by an Iron(0) Porphyrin Associated with a Weak Bronsted Acid Cocatalyst. J. Am. Chem. Soc. 1994, 116, 5015–5016. [Google Scholar] [CrossRef]
  66. Hod, I.; Sampson, M.D.; Deria, P.; Kubiak, C.P.; Farha, O.K.; Hupp, J.T. Fe-Porphyrin-Based Metal–Organic Framework Films as High-Surface Concentration, Heterogeneous Catalysts for Electrochemical Reduction of CO2. ACS Catal. 2015, 5, 6302–6309. [Google Scholar] [CrossRef]
  67. Wu, Q.; Xie, R.-K.; Mao, M.-J.; Chai, G.-L.; Yi, J.-D.; Zhao, S.-S.; Huang, Y.-B.; Cao, R. Integration of Strong Electron Transporter Tetrathiafulvalene into Metalloporphyrin-Based Covalent Organic Framework for Highly Efficient Electroreduction of CO2. ACS Energy Lett. 2020, 5, 1005–1012. [Google Scholar] [CrossRef]
  68. Ejsmont, A.; Jankowska, A.; Goscianska, J. Insight into the Photocatalytic Activity of Cobalt-Based Metal–Organic Frameworks and Their Composites. Catalysts 2022, 12, 110. [Google Scholar] [CrossRef]
  69. Kornienko, N.; Zhao, Y.; Kley, C.S.; Zhu, C.; Kim, D.; Lin, S.; Chang, C.J.; Yaghi, O.M.; Yang, P. Metal-organic frameworks for electrocatalytic reduction of carbon dioxide. J. Am. Chem. Soc. 2015, 137, 14129–14135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Sun, M.L.; Wang, Y.R.; He, W.W.; Zhong, R.L.; Liu, Q.Z.; Xu, S.; Xu, J.M.; Han, X.L.; Ge, X.; Li, S.L.; et al. Efficient Electron Transfer from Electron-Sponge Polyoxometalate to Single-Metal Site Metal-Organic Frameworks for Highly Selective Electroreduction of Carbon Dioxide. Small 2021, 17, e2100762. [Google Scholar] [CrossRef]
  71. Sun, Z.; Hu, Y.; Zhou, D.; Sun, M.; Wang, S.; Chen, W. Factors Influencing the Performance of Copper-Bearing Catalysts in the CO2 Reduction System. ACS Energy Lett. 2021, 6, 3992–4022. [Google Scholar] [CrossRef]
  72. Hinogami, R.; Yotsuhashi, S.; Deguchi, M.; Zenitani, Y.; Hashiba, H.; Yamada, Y. Electrochemical Reduction of Carbon Dioxide Using a Copper Rubeanate Metal Organic Framework. ECS Electrochem. Lett. 2012, 1, H17–H19. [Google Scholar] [CrossRef]
  73. Senthil Kumar, R.; Senthil Kumar, S.; Anbu Kulandainathan, M. Highly selective electrochemical reduction of carbon dioxide using Cu based metal organic framework as an electrocatalyst. Electrochem. Commun. 2012, 25, 70–73. [Google Scholar] [CrossRef]
  74. Kang, X.; Li, L.; Sheveleva, A.; Han, X.; Li, J.; Liu, L.; Tuna, F.; McInnes, E.J.L.; Han, B.; Yang, S.; et al. Electro-reduction of carbon dioxide at low over-potential at a metal-organic framework decorated cathode. Nat. Commun. 2020, 11, 5464. [Google Scholar] [CrossRef]
  75. Zhang, Y.; Jiao, L.; Yang, W.; Xie, C.; Jiang, H.L. Rational Fabrication of Low-Coordinate Single-Atom Ni Electrocatalysts by MOFs for Highly Selective CO2 Reduction. Angew. Chem. Int. Ed. 2021, 60, 7607–7611. [Google Scholar] [CrossRef]
  76. Zhang, M.-D.; Si, D.-H.; Yi, J.-D.; Yin, Q.; Huang, Y.-B.; Cao, R. Conductive phthalocyanine-based metal-organic framework as a highly efficient electrocatalyst for carbon dioxide reduction reaction. Sci China Chem. 2021, 64, 1332–1339. [Google Scholar] [CrossRef]
  77. Jiao, L.; Yang, W.; Wan, G.; Zhang, R.; Zheng, X.; Zhou, H.; Yu, S.H.; Jiang, H.L. Single-Atom Electrocatalysts from Multivariate Metal-Organic Frameworks for Highly Selective Reduction of CO2 at Low Pressures. Angew. Chem. Int. Ed. 2020, 59, 20589–20595. [Google Scholar] [CrossRef]
  78. Wang, Y.R.; Huang, Q.; He, C.T.; Chen, Y.; Liu, J.; Shen, F.C.; Lan, Y.Q. Oriented electron transmission in polyoxometalate-metalloporphyrin organic framework for highly selective electroreduction of CO2. Nat. Commun. 2018, 9, 4466. [Google Scholar] [CrossRef] [Green Version]
  79. Sun, D.; Liu, W.; Qiu, M.; Zhang, Y.; Li, Z. Introduction of a mediator for enhancing photocatalytic performance via post-synthetic metal exchange in metal-organic frameworks (MOFs). Chem. Commun. 2015, 51, 2056–2059. [Google Scholar] [CrossRef]
  80. Wang, D.; Huang, R.; Liu, W.; Sun, D.; Li, Z. Fe-Based MOFs for Photocatalytic CO2 Reduction: Role of Coordination Unsaturated Sites and Dual Excitation Pathways. ACS Catal. 2014, 4, 4254–4260. [Google Scholar] [CrossRef]
  81. Huang, Q.; Liu, J.; Feng, L.; Wang, Q.; Guan, W.; Dong, L.Z.; Zhang, L.; Yan, L.K.; Lan, Y.Q.; Zhou, H.C. Multielectron transportation of polyoxometalate-grafted metalloporphyrin coordination frameworks for selective CO2-to-CH4 photoconversion. Natl. Sci. Rev. 2020, 7, 53–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Wang, S.; Yao, W.; Lin, J.; Ding, Z.; Wang, X. Cobalt imidazolate metal-organic frameworks photosplit CO2 under mild reaction conditions. Angew. Chem. Int. Ed. 2014, 53, 1034–1038. [Google Scholar] [CrossRef] [PubMed]
  83. Wang, X.-K.; Liu, J.; Zhang, L.; Dong, L.-Z.; Li, S.-L.; Kan, Y.-H.; Li, D.-S.; Lan, Y.-Q. Monometallic Catalytic Models Hosted in Stable Metal–Organic Frameworks for Tunable CO2 Photoreduction. ACS Catal. 2019, 9, 1726–1732. [Google Scholar] [CrossRef]
  84. Zhang, H.X.; Hong, Q.L.; Li, J.; Wang, F.; Huang, X.; Chen, S.; Tu, W.; Yu, D.; Xu, R.; Zhou, T.; et al. Isolated Square-Planar Copper Center in Boron Imidazolate Nanocages for Photocatalytic Reduction of CO2 to CO. Angew. Chem. Int. Ed. 2019, 58, 11752–11756. [Google Scholar] [CrossRef] [PubMed]
  85. Han, B.; Ou, X.; Deng, Z.; Song, Y.; Tian, C.; Deng, H.; Xu, Y.J.; Lin, Z. Nickel Metal-Organic Framework Monolayers for Photoreduction of Diluted CO2: Metal-Node-Dependent Activity and Selectivity. Angew. Chem. Int. Ed. 2018, 57, 16811–16815. [Google Scholar] [CrossRef]
  86. Shinde, G.Y.; Mote, A.S.; Gawande, M.B. Recent Advances of Photocatalytic Hydrogenation of CO2 to Methanol. Catalysts 2022, 12, 94. [Google Scholar] [CrossRef]
  87. He, H.; Sun, Q.; Gao, W.; Perman, J.A.; Sun, F.; Zhu, G.; Aguila, B.; Forrest, K.; Space, B.; Ma, S. A Stable Metal-Organic Framework Featuring a Local Buffer Environment for Carbon Dioxide Fixation. Angew. Chem. Int. Ed. 2018, 57, 4657–4662. [Google Scholar] [CrossRef]
  88. Wang, S.; Wang, X. Photocatalytic CO2 reduction by CdS promoted with a zeolitic imidazolate framework. Appl. Catal. B Environ. 2015, 162, 494–500. [Google Scholar] [CrossRef]
  89. Li, J.; Luo, D.; Yang, C.; He, S.; Chen, S.; Lin, J.; Zhu, L.; Li, X. Copper(II) imidazolate frameworks as highly efficient photocatalysts for reduction of CO2 into methanol under visible light irradiation. J. Solid State Chem. 2013, 203, 154–159. [Google Scholar] [CrossRef]
  90. Li, N.; Liu, X.; Zhou, J.; Chen, W.; Liu, M. Encapsulating CuO quantum dots in MIL-125(Ti) coupled with g-C3N4 for efficient photocatalytic CO2 reduction. Chem. Eng. J. 2020, 399, 125782. [Google Scholar] [CrossRef]
  91. Wang, Z.; Jiao, X.; Chen, D.; Li, C.; Zhang, M. Porous Copper/Zinc Bimetallic Oxides Derived from MOFs for Efficient Photocatalytic Reduction of CO2 to Methanol. Catalysts 2020, 10, 1127. [Google Scholar] [CrossRef]
  92. Cardoso, J.C.; Stulp, S.; de Brito, J.F.; Flor, J.B.S.; Frem, R.C.G.; Zanoni, M.V.B. MOFs based on ZIF-8 deposited on TiO2 nanotubes increase the surface adsorption of CO2 and its photoelectrocatalytic reduction to alcohols in aqueous media. Appl. Catal. B Environ. 2018, 225, 563–573. [Google Scholar] [CrossRef] [Green Version]
  93. Helal, A.; Usman, M.; Arafat, M.E.; Abdelnaby, M.M. Allyl functionalized UiO-66 metal-organic framework as a catalyst for the synthesis of cyclic carbonates by CO2 cycloaddition. J. Ind. Eng. Chem. 2020, 89, 104–110. [Google Scholar] [CrossRef]
  94. Di, Z.; Liu, C.; Pang, J.; Chen, C.; Hu, F.; Yuan, D.; Wu, M.; Hong, M. Cage-like Porous Materials with Simultaneous High C2H2 Storage and Excellent C2H2/CO2 Separation Performance. Angew. Chem. Int. Ed. 2021, 60, 10828–10832. [Google Scholar] [CrossRef] [PubMed]
  95. Meng, J.; Chen, Q.; Lu, J.; Liu, H. Z-Scheme Photocatalytic CO2 Reduction on a Heterostructure of Oxygen-Defective ZnO/Reduced Graphene Oxide/UiO-66-NH2 under Visible Light. ACS Appl. Mater. Interfaces 2019, 11, 550–562. [Google Scholar] [CrossRef]
  96. Su, Y.; Zhang, Z.; Liu, H.; Wang, Y. Cd0.2Zn0.8S@UiO-66-NH2 nanocomposites as efficient and stable visible-light-driven photocatalyst for H2 evolution and CO2 reduction. Appl. Catal. B Environ. 2017, 200, 448–457. [Google Scholar] [CrossRef]
  97. Wang, G.; He, C.T.; Huang, R.; Mao, J.; Wang, D.; Li, Y. Photoinduction of Cu Single Atoms Decorated on UiO-66-NH2 for Enhanced Photocatalytic Reduction of CO2 to Liquid Fuels. J. Am. Chem. Soc. 2020, 142, 19339–19345. [Google Scholar] [CrossRef] [PubMed]
  98. Verma, P.; Stewart, D.J.; Raja, R. Recent Advances in Photocatalytic CO2 Utilisation Over Multifunctional Metal–Organic Frameworks. Catalysts 2020, 10, 1176. [Google Scholar] [CrossRef]
  99. Zhu, W.; Zhang, C.; Li, Q.; Xiong, L.; Chen, R.; Wan, X.; Wang, Z.; Chen, W.; Deng, Z.; Peng, Y. Selective reduction of CO2 by conductive MOF nanosheets as an efficient co-catalyst under visible light illumination. Appl. Catal. B Environ. 2018, 238, 339–345. [Google Scholar] [CrossRef]
  100. Garrido, M.; Volland, M.K.; Munich, P.W.; Rodriguez-Perez, L.; Calbo, J.; Orti, E.; Herranz, M.A.; Martin, N.; Guldi, D.M. Mono- and Tripodal Porphyrins: Investigation on the Influence of the Number of Pyrene Anchors in Carbon Nanotube and Graphene Hybrids. J. Am. Chem. Soc. 2020, 142, 1895–1903. [Google Scholar] [CrossRef]
  101. Zhang, H.; Wei, J.; Dong, J.; Liu, G.; Shi, L.; An, P.; Zhao, G.; Kong, J.; Wang, X.; Meng, X.; et al. Efficient Visible-Light-Driven Carbon Dioxide Reduction by a Single-Atom Implanted Metal-Organic Framework. Angew. Chem. Int. Ed. 2016, 55, 14310–14314. [Google Scholar] [CrossRef] [PubMed]
  102. Choi, K.M.; Kim, D.; Rungtaweevoranit, B.; Trickett, C.A.; Barmanbek, J.T.; Alshammari, A.S.; Yang, P.; Yaghi, O.M. Plasmon-Enhanced Photocatalytic CO Conversion within Metal-Organic Frameworks under Visible Light. J. Am. Chem. Soc. 2017, 139, 356–362. [Google Scholar] [CrossRef] [PubMed]
  103. Lan, G.; Li, Z.; Veroneau, S.S.; Zhu, Y.Y.; Xu, Z.; Wang, C.; Lin, W. Photosensitizing Metal-Organic Layers for Efficient Sunlight-Driven Carbon Dioxide Reduction. J. Am. Chem. Soc. 2018, 140, 12369–12373. [Google Scholar] [CrossRef] [PubMed]
  104. Sun, D.; Fu, Y.; Liu, W.; Ye, L.; Wang, D.; Yang, L.; Fu, X.; Li, Z. Studies on photocatalytic CO2 reduction over NH2 -Uio-66(Zr) and its derivatives: Towards a better understanding of photocatalysis on metal-organic frameworks. Chem. Eur. J. 2013, 19, 14279–14285. [Google Scholar] [CrossRef] [PubMed]
  105. Fu, Y.; Sun, D.; Chen, Y.; Huang, R.; Ding, Z.; Fu, X.; Li, Z. An amine-functionalized titanium metal-organic framework photocatalyst with visible-light-induced activity for CO2 reduction. Angew. Chem. Int. Ed. 2012, 51, 3364–3367. [Google Scholar] [CrossRef] [PubMed]
  106. Li, N.; Liu, J.; Liu, J.J.; Dong, L.Z.; Xin, Z.F.; Teng, Y.L.; Lan, Y.Q. Adenine Components in Biomimetic Metal-Organic Frameworks for Efficient CO2 Photoconversion. Angew. Chem. Int. Ed. 2019, 58, 5226–5231. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of CO2–RR toward artificial carbon cycle.
Figure 1. Schematic illustration of CO2–RR toward artificial carbon cycle.
Catalysts 12 01582 g001
Figure 2. Electrocatalytic CO2–RR performances of H−POM@PCN−222 (Co), H−POM@PCN−222 (Fe), H−POM@PCN−222 (Mn), H−POM@PCN−222 (Ni), PCN−222 (Co), and POM. (a) Linear sweep voltametric scans. (b) Faradaic efficiencies for CO with different metals and at different potentials. (c) Maximum FECO for six samples at −0.8 V (vs RHE). (d) CO current density at different potentials. (e) Tafel plots. (f) The turnover frequency for different metals and PCN−222 (Co). Re−printed with permission from Ref. [70]. Copyright 2021, Wiley–VCH.
Figure 2. Electrocatalytic CO2–RR performances of H−POM@PCN−222 (Co), H−POM@PCN−222 (Fe), H−POM@PCN−222 (Mn), H−POM@PCN−222 (Ni), PCN−222 (Co), and POM. (a) Linear sweep voltametric scans. (b) Faradaic efficiencies for CO with different metals and at different potentials. (c) Maximum FECO for six samples at −0.8 V (vs RHE). (d) CO current density at different potentials. (e) Tafel plots. (f) The turnover frequency for different metals and PCN−222 (Co). Re−printed with permission from Ref. [70]. Copyright 2021, Wiley–VCH.
Catalysts 12 01582 g002
Figure 3. Photograph (a) and SEM image (b) of a fresh Cu-foam electrode (0.5 × 1.0 cm2). (c) Photograph of the as-prepared electrode Cu2(L)-e/Cu (0.5 × 1.0 cm2). (df) SEM images of the as-prepared electrode Cu2(L)-e/Cu. (g,h) Views of the crystal structure of [Cu2(L)]. Hydrogen atoms are omitted for clarity. The scale bars of (b,df) are 300 µm, 300 µm, 300 nm and 100 nm, respectively. Reprinted with permission from Ref. [74]. Copyright 2020, Springer Nature.
Figure 3. Photograph (a) and SEM image (b) of a fresh Cu-foam electrode (0.5 × 1.0 cm2). (c) Photograph of the as-prepared electrode Cu2(L)-e/Cu (0.5 × 1.0 cm2). (df) SEM images of the as-prepared electrode Cu2(L)-e/Cu. (g,h) Views of the crystal structure of [Cu2(L)]. Hydrogen atoms are omitted for clarity. The scale bars of (b,df) are 300 µm, 300 µm, 300 nm and 100 nm, respectively. Reprinted with permission from Ref. [74]. Copyright 2020, Springer Nature.
Catalysts 12 01582 g003
Figure 4. Fabrication of low-coordination single-atom Ni electrocatalysts via a PSMS strategy. Reprinted with permission from Ref. [75]. Copyright 2021, Wiley-VCH.
Figure 4. Fabrication of low-coordination single-atom Ni electrocatalysts via a PSMS strategy. Reprinted with permission from Ref. [75]. Copyright 2021, Wiley-VCH.
Catalysts 12 01582 g004
Figure 5. (a) View of the 3D network of MOF-525-Co featuring a highly porous framework and incorporated active sites. (b) Fourier transform magnitudes of the experimental Co K-edge EXAFS spectra of samples (not corrected for phase shift). Key: Co foil (blue), Co@C (orange), MOF-525-Co (green). (c) Wavelet transform for the k3-weighted EXAFS signal of MOF-525-Co, based on Morlet wavelets with optimum resolution at the first (lower panel) and higher (upper panel) coordination shells. Vertical dashed lines denoting the k-space positions of the Co-N and Co-Co contributions are provided to guide the eye. (d) Comparison between the Co K-edge XANES experimental spectrum (orange) of MOF-525-Co and the theoretical spectrum calculated with the depicted structure. For clarity, the non-convoluted theoretical spectrum is also shown. The energy-dependent exchange–correlation potential was calculated in the real Hedin–Lundqvist scheme, and then the spectra are convoluted using a Lorentzian function with an energy-dependent width to account for the broadening due both to the core–hole width and to the final state width. (e) UV/Vis spectra of MOF-525 (green), MOF-525-Co (purple), and MOF-525-Zn (orange). Reprinted with permission from Ref. [101]. Copyright 2016, Wiley-VCH.
Figure 5. (a) View of the 3D network of MOF-525-Co featuring a highly porous framework and incorporated active sites. (b) Fourier transform magnitudes of the experimental Co K-edge EXAFS spectra of samples (not corrected for phase shift). Key: Co foil (blue), Co@C (orange), MOF-525-Co (green). (c) Wavelet transform for the k3-weighted EXAFS signal of MOF-525-Co, based on Morlet wavelets with optimum resolution at the first (lower panel) and higher (upper panel) coordination shells. Vertical dashed lines denoting the k-space positions of the Co-N and Co-Co contributions are provided to guide the eye. (d) Comparison between the Co K-edge XANES experimental spectrum (orange) of MOF-525-Co and the theoretical spectrum calculated with the depicted structure. For clarity, the non-convoluted theoretical spectrum is also shown. The energy-dependent exchange–correlation potential was calculated in the real Hedin–Lundqvist scheme, and then the spectra are convoluted using a Lorentzian function with an energy-dependent width to account for the broadening due both to the core–hole width and to the final state width. (e) UV/Vis spectra of MOF-525 (green), MOF-525-Co (purple), and MOF-525-Zn (orange). Reprinted with permission from Ref. [101]. Copyright 2016, Wiley-VCH.
Catalysts 12 01582 g005
Figure 6. Structures of Ren-MOF and Ag⊂Ren-MOF for plasmon-enhanced photocatalytic CO2 conversion. (a) Zr6O4(OH)4(−CO2)12 secondary building units are combined with 4,4′-biphenyldicarboxylate (BPDC) and ReTC linkers to form Ren-MOF. The structure of Re3-MOF is shown, as identified from single-crystal X-ray diffraction. The 12 coordinated Zr-based metal clusters are interconnected with 21 BPDC and 3 ReTC linkers in a face-centred cubic array. Atom labeling scheme: C, black; O, red; Zr, blue polyhedral; Re, yellow; Cl, green; H atoms are omitted for clarity. (b) Ren-MOF coated on an Ag nanocube for enhanced photocatalytic conversion of CO2. Reprinted with permission from [102]. Copyright 2017, American Chemical Society.
Figure 6. Structures of Ren-MOF and Ag⊂Ren-MOF for plasmon-enhanced photocatalytic CO2 conversion. (a) Zr6O4(OH)4(−CO2)12 secondary building units are combined with 4,4′-biphenyldicarboxylate (BPDC) and ReTC linkers to form Ren-MOF. The structure of Re3-MOF is shown, as identified from single-crystal X-ray diffraction. The 12 coordinated Zr-based metal clusters are interconnected with 21 BPDC and 3 ReTC linkers in a face-centred cubic array. Atom labeling scheme: C, black; O, red; Zr, blue polyhedral; Re, yellow; Cl, green; H atoms are omitted for clarity. (b) Ren-MOF coated on an Ag nanocube for enhanced photocatalytic conversion of CO2. Reprinted with permission from [102]. Copyright 2017, American Chemical Society.
Catalysts 12 01582 g006
Table 1. Standard electrochemical potentials for CO2 reduction.
Table 1. Standard electrochemical potentials for CO2 reduction.
ReactionE0/[V vs RHE (Reversible
Hydrogen Electrode)]
2H2O → O2 + 4H+ + 4e1.23
xCO2 + nH+ + ne → product + yH2O-
CO2 + 2H+ + 2e → HCOOH(aq)−0.12
CO2 + 2H+ + 2e → CO(g) + H2O−0.10
CO2 + 4H+ + 4e → C(s) + 2H2O0.21
CO2 + 6H+ + 6e → CH3OH(aq) + H2O0.03
CO2 + 8H+ + 8e → CH4(g) + 2H2O0.17
2CO2 + 12H+ + 12e → C2H4(g) + 4H2O0.08
2CO2 + 12H+ + 12e → C2H5OH(aq) + 3H2O0.09
2CO2 + 14H+ + 14e → C2H6(g) + 4H2O0.14
3CO2 + 18H+ + 18e → C3H7OH(aq) + 5H2O0.10
2CO2 + 2H+ + 2e → (COOH)2(s)−0.47
2CO2 + 8H+ + 8e → CH3COOH(aq) + 2H2O0.11
2CO2 + 10H+ + 10e → CH3CHO(aq) + 3H2O0.06
3CO2 + 16H+ + 16e → C2H5CHO(aq) + 5H2O0.09
xCO + nH+ + ne → product + yH2O-
CO + 6H+ + 6e → CH4(g) + H2O0.26
2CO + 8H+ + 8e → CH3CH2OH(aq) + H2O0.19
2CO + 8H+ + 8e → C2H4(g) + 2H2O0.17
2H+ + 2e → H20
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Di, Z.; Qi, Y.; Yu, X.; Hu, F. The Progress of Metal-Organic Framework for Boosting CO2 Conversion. Catalysts 2022, 12, 1582. https://doi.org/10.3390/catal12121582

AMA Style

Di Z, Qi Y, Yu X, Hu F. The Progress of Metal-Organic Framework for Boosting CO2 Conversion. Catalysts. 2022; 12(12):1582. https://doi.org/10.3390/catal12121582

Chicago/Turabian Style

Di, Zhengyi, Yu Qi, Xinxin Yu, and Falu Hu. 2022. "The Progress of Metal-Organic Framework for Boosting CO2 Conversion" Catalysts 12, no. 12: 1582. https://doi.org/10.3390/catal12121582

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop