Next Article in Journal
The Uptake of Ivermectin and Its Effects in Roots, Leaves and Seeds of Soybean (Glycine max)
Next Article in Special Issue
Oxazolochlorins 21. Most Efficient Access to meso-Tetraphenyl- and meso-Tetrakis(pentafluorophenyl)porpholactones, and Their Zinc(II) and Platinum(II) Complexes
Previous Article in Journal
An Optimization of Liquid–Liquid Extraction of Urinary Volatile and Semi-Volatile Compounds and Its Application for Gas Chromatography-Mass Spectrometry and Proton Nuclear Magnetic Resonance Spectroscopy
Previous Article in Special Issue
Weak Interactions and Conformational Changes in Core-Protonated A2- and Ax-Type Porphyrin Dications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress in Conversion of CO2 to Valuable Fuels

Centre for Bioengineering and Biotechnology, China University of Petroleum, Qingdao 266580, China
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(16), 3653; https://doi.org/10.3390/molecules25163653
Submission received: 28 June 2020 / Revised: 3 August 2020 / Accepted: 5 August 2020 / Published: 11 August 2020

Abstract

:
Rapid growth in the world’s economy depends on a significant increase in energy consumption. As is known, most of the present energy supply comes from coal, oil, and natural gas. The overreliance on fossil energy brings serious environmental problems in addition to the scarcity of energy. One of the most concerning environmental problems is the large contribution to global warming because of the massive discharge of CO2 in the burning of fossil fuels. Therefore, many efforts have been made to resolve such issues. Among them, the preparation of valuable fuels or chemicals from greenhouse gas (CO2) has attracted great attention because it has made a promising step toward simultaneously resolving the environment and energy problems. This article reviews the current progress in CO2 conversion via different strategies, including thermal catalysis, electrocatalysis, photocatalysis, and photoelectrocatalysis. Inspired by natural photosynthesis, light-capturing agents including macrocycles with conjugated structures similar to chlorophyll have attracted increasing attention. Using such macrocycles as photosensitizers, photocatalysis, photoelectrocatalysis, or coupling with enzymatic reactions were conducted to fulfill the conversion of CO2 with high efficiency and specificity. Recent progress in enzyme coupled to photocatalysis and enzyme coupled to photoelectrocatalysis were specially reviewed in this review. Additionally, the characteristics, advantages, and disadvantages of different conversion methods were also presented. We wish to provide certain constructive ideas for new investigators and deep insights into the research of CO2 conversion.

Graphical Abstract

1. Introduction

Overreliance on fossil fuels have led to the discharge of more and more carbon dioxide (CO2) into the atmosphere accompanied with the rapid development of modern industry. This has resulted in serious environmental problems, including global warming and other related problems, such as rising sea levels, ocean acidification, ozone layer depletion, and extreme weather patterns [1]. Additionally, the excessive consumption of fossil fuels and deforestation have interrupted the carbon cycle on the earth and accelerated the environmental deterioration and resource depletion [2,3]. In order to mitigate the issue of global warming, most countries have signed the Paris Agreement, which aims to abate the net atmospheric CO2 levels by 2050. Therefore, efficient measures have to be made to reduce atmospheric CO2 levels, such as reducing the direct emission of CO2 by developing clean and sustainable energy resources to replace traditional fossil fuels, capturing and storing CO2, and converting the anthropogenic CO2 into useful chemicals or fuels by reducing it. Among them, the efficient conversion of CO2 has attracted great consideration because of the fact that it is the most abundant C1 compound and is the major greenhouse gas [4,5,6,7]. The conversion of CO2 to value-added fuels or chemicals is important in recycling carbon species and simultaneously solving environmental problems and energy crises.
CO2 is one of the most stable molecules in which carbon is in the highest valence state. It is difficult to have an electrophilic reaction because of its poor electron affinity. Hence, the conversion of CO2 depends on nucleophilic attack of the carbon atom. As is known, the dissociation energy for breaking the C=O bond in CO2 molecules is higher than 750 kJ mol−1 [8]. This is an uphill reaction from a thermodynamic point of view. To complete such a reaction, high temperature, high pressure environment, or highly efficient catalysts are typically required to provide the necessary energy. Till now, different strategies including thermal catalysis [9,10,11,12,13], photocatalysis [14,15,16,17], electrocatalysis [18,19,20,21], and photoelectrochemical (PEC) reactions [22,23,24,25] have been adopted to conduct the reduction of CO2, in which heat, light, or electricity were used to supply essential energy for the reaction. As is known, eight electrons are needed for each CO2 molecule to complete the conversion to hydrocarbon compounds. This leads to various products during the reduction process, resulting in complicated purification procedures and poor yield of desired products. Inspired by natural photosynthesis, highly efficient and specific enzymatic reactions were incorporated into the aforementioned reducing technologies to improve the efficiency and specificity of CO2 conversion [26,27].
In this review, research progress of CO2 conversion through different strategies were summarized and discussed with detailed comments of advantages and disadvantages (Scheme 1). In particular, using macrocycles as photosensitizers, current achievement, development, and catalytic activity in photocatalytic and photoelectrocatalytic reactions coupled with enzymatic reactions were highlighted. Finally, this article also presented the main challenges and certain future prospects for the conversion of CO2 to useful fuels.

2. Catalytic Reduction of CO2

2.1. Thermal Catalysis

Most thermal catalytic conversion of CO2 involves the hydrogenation reaction at relatively low temperatures (≤523 K) to produce useful fuels such as CO, methane, and methanol [28]. Since CO2 molecules are thermodynamically and chemically stable, large amounts of energy are required if CO2 is used as a single reactant. The introduction of other substances with higher Gibbs free energy (such as H2) as the co-reactant will make the thermodynamic process easier [29]. In the past few decades, great attention has been paid to the thermal catalysis of CO2 and significant progress has been made. Various catalysts, including metals (Cu, Co, etc.) and metal oxides (ZnO2, InO2, etc.) [30,31,32] as well as novel nano-sized catalyst metal–organic frameworks (MOFs) have also been designed, prepared, and gradually developed.
Rungtaweevoranit and coauthors [33] reported a catalyst where Cu nanocrystal (Cu NC) was encapsulated in a metal–organic framework UiO-66 [Zr6O4(OH)4(BDC)6, BDC = 1,4-benzenedicarboxylate] to form catalyst Cu⊂UiO-66, which could catalyze the generation of methanol via the hydrogenation of CO2. Figure 1a shows the synthetic process for an ordered UiO-66 crystal structure from the reaction of Zr oxide [Zr6O4(OH)4(−CO2)12] secondary building units (SBUs) and BDC. By using Zr(OPrn)4, one Cu nanocrystal was successfully encapsulated in a nanocrystal UiO-66, as shown in Figure 1b. Cu⊂UiO-66 was precisely placed on the Cu surface, yielding high interfacial contact between Cu NC and Zr oxide SBUs, which could ensure the reactants reach the active sites (Figure 1c). By comparing the methanol turnover frequency (TOF) of the Cu⊂UiO-66 catalyst with the Cu/ZnO/Al2O3 catalyst at different reaction temperature (Figure 1d), it was found that the performance of the catalyst exceeded that of the Cu/ZnO/Al2O3 catalyst at relative lower temperatures (<250 °C), which can steadily increase the conversion rate by 8 times and with 100% methanol selectivity.
Various carbon materials including carbon nanofibers [34], carbon nanotubes [35,36,37,38], biochar [39], and carbon felt [40] have also been employed as carriers for CO2 hydrogenation catalysts, taking advantage of their high hydrogen storage capacity, high thermal conductivity, and high specific surface area of carbon carriers. Nanosized materials are used to define nanoscale catalyst structures, in which the composition of catalysts and their surface structures can be adjusted and may bring to darewidespread applications.
Despite such achievements, the reactant H2 added in the thermal catalysis process is usually more valuable than the product methane and methanol. Considering the higher cost than that from fossil fuels, the direct hydrogenation reaction of CO2 was rarely used to produce methane or methanol [28]. There are still great challenges in developing catalysts with high catalytic performance and long-term stability, reducing the size of thermal catalytic reactors and decreasing the production costs. In addition, more effective and economical methods to produce H2 are urgently needed [13]. In this case, CO2 conversion to useful fuels are attempted by other methods such as photocatalysis, electrocatalysis, and photoelectrocatalysis.

2.2. Photocatalysis

Solar energy is as an ideal energy source to replace traditional fossil fuels because it is an abundant, cheap, clean, and sustainable energy source. Therefore, the use of photocatalysts for solar-driven fuels or chemicals from CO2 is a very attractive approach. Similar to natural photosynthesis, electron–hole pairs are generated when the photocatalysts are exposed to solar light. The photogenerated electrons induce CO2 to undergo a redox reaction that results in hydrocarbon formation. There are three crucial procedures during the photocatalytic conversion of CO2: (1) absorption of sunlight; (2) charge separation and transfer; and (3) catalytic reduction of CO2 and oxidation of H2O [17]. Each procedure during the conversion of CO2 is closely related with the photocatalysts. Until now, the photocatalysts were mainly from semiconductor materials which are abundant on earth and easy to obtain [41]. As for the reaction products, CO, methane, formic acid, and other chemicals containing one or two carbon atoms are usually involved.
Until now, many efforts have been made to optimize the structure and composition of photocatalysts or integrate them with other functional units to construct multifunctional catalysts. For example, integration of photocatalysts with metal–organic frameworks (MOFs) has been demonstrated to offer more adsorptive sites for CO2 uptake because of their extreme larger surface area and microporous structure [42,43,44], resulting in remarkable improvement in CO2 conversion. However, there is a wide gap between the photocatalytic performance of these complexes and the requirements for practical application [45].
The construction of multi-junctions are randomly distributed on the surface of photocatalysts, which improve interfacial electron–hole separation and migration, even though the separation efficiency remains to be raised to a higher level [46]. Meng et al. [16] deposited MnOx nanosheets and Pt nanoparticles on different facets of anatase TiO2 to form surface heterojunction. The results indicated that heterojunction with multiple nodes in the photocatalysts improved the conversion efficiency of CO2.
Two dimensional (2D) nanosheets are particularly promising in improving charge separation because the photogenerated electrons and holes will move to the interface with shorter distances. Wei and coauthors [47] synthesized a series of heterostructured CdS/BiVO4 composites by depositing different amounts of CdS on the surface of BiVO4 nanosheets with variable thickness. The results showed that CdS/BiVO4 nanocomposites had higher photocatalytic activity in CO2 reduction than that of pure BiVO4 and CdS. Furthermore, the content of CdS in the composites were responsible for the yield of CO and CH4. Enhancement of photocatalytic activity was attributed to the synergistic effect of forming Z-scheme herterojunction and reduced thickness of BiVO4. According to density functional theory (DFT), theoretical calculations have been made for 2D photocatalysts and other types of catalysts [48,49,50,51,52], such that the characteristics of materials and the role of different components in the catalytic mechanism or the entire reaction cycle can be explored in depth. Computational approaches provide a way for understanding the catalytic effects at the mesoscopic and micro level. The mechanism of photocatalytic CO2 reduction remains obscure and needs more in-depth investigations.
In addition to inorganic photocatalysts, certain organic photocatalysts including porphyrin-based photocatalysts were also involved in CO2 reduction. Porphyrin are planar macrocyclic molecules and widely distributed in nature. There are four pyrroles connected in ring fashion through four methylene carbons. Also, porphyrins have highly delocalized π electrons to form a planar conjugated framework. This endows them with strong absorption in the visible light region and unique electronic redox characteristics. Moreover, the NH protons inside the ring are easily deprotonated and therefore exhibit remarkable coordination characteristics toward metal ions. For example, chlorophyll α is a complex of magnesium and porphyrin, and plays an important role in light capturing and H2O oxidation [53]. Incorporation of porphyrin into MOF can further improve its photocatalytic performance. Wang et al. [54] synthesized an indium–porphyrin MOF framework (In(H2TCPP)(1-n)[Fe(TCPP)(H2O)](1-n)[DEA](1-n)(In−FenTCPP-MOF) by incorporating porphyrin into MOF. Porphyrin rings in the MOFs support the single-site iron. This can both support the iron center as a catalytic active site and absorb visible light for high-performance conversion of CO2 to CO due to the synergetic effect between the porphyrin and the high-performance single-center Fe catalytic center (Figure 2).
Due to strong chemical bonds of CO2 molecules and complex transformation path involving multiple electrons [55,56], many problems in the CO2 photocatalytic conversion need to be resolved, such as low conversion efficiency, poor selectivity of products, competition for the generation of H2, and rapid recombination of photogenerated electrons and holes [57,58]. Furthermore, most commercial photocatalysts used for CO2 conversion contain toxic and expensive transition metal elements, resulting in cost increase and waste disposal. Therefore, it is necessary to design and prepare new photocatalytic materials that can effectively increase the activity, suppress competitive reactions, improve conversion efficiency, and even develop a new CO2 catalytic reduction system. Researchers have shown that electric field can effectively inhibit charge recombination [59], so electrocatalytic reduction of CO2 has been extensively studied.

2.3. Electrocatalysis

Certain renewable resources such as solar energy and wind energy are usually intermittent and limited by the season and weather, so energy storage technology is necessary for uninterrupted energy supply [60]. The electrocatalytic conversion of CO2 to valuable chemicals is an attractive solution for reducing atmospheric CO2 and storing energy. Using an external electric field as an energy source and water as the proton donor, various catalysts are applied to catalyze the reduction of CO2. Compared with thermocatalysis, the electrocatalytic conversion is a higher cost-effective method because water replacing H2 is used as the proton donor. Electrocatalytic CO2 reduction has attracted great attention due to its mild operating conditions (normal temperature and pressure), controllable reaction process conditions and reaction rate, recyclable catalyst and electrolyte, high energy utilization, simple equipment, and achievable conversion efficiency [61,62,63,64]. In the past few years, researchers have explored electrocatalytic reduction of CO2 using different electrode materials, such as metals [65,66], transition metal oxides [19,67], transition metal chalcogenides [68,69], metal-free 2D materials [70,71], metal–organic frameworks (MOFs) [72,73,74], and various reduction products including CO, methane, formic acid, ethanol, and other compounds were obtained.
Hu et al. [75] investigated the electrocatalytic performance of cobalt meso-tetraphenylporphyrin (CoTPP) and its complex with carbon materials under both homogeneous and heterogeneous conditions. Their catalytic ability for CO2 reduction was significantly increased by the strong π–π interactions between CoTPP and carbon materials when CoTPP was incorporated with carbon nanotubes (CNTs) or similar carbon materials (Figure 3).
Wang and coauthors [76] designed and synthesized a series of stable reductive polyoxometalate-metalloporphyrin organic frameworks (M-PMOF, M = Co, Fe, Ni, zinc, as shown in Figure 4) by using reductive polyoxometalates (POMs, such as Zn-ε-Keggin cluster, as electron donor) as building block and metalloporphyrin as linker. Metalloporphyrin is helpful for electron mobility for its inherent macrocycle conjugated π-electron structure. Connection of Zn-ε-Keggin and M-TCPP might create an oriented electron transportation pathway by which multiple electron transfer processes in electrocatalytic CO2 reduction were completed. The electrocatalytic performance of M-PMOFs was measured by linear sweep voltammetry. The total current density of Co-PMOF at −1.1 V was 38.9 mA cm−2, higher than that of Fe-PMOF (25.1 mA cm−2), Ni-PMOF (20.02 mA cm−2), and Zn-PMOF (16 mA cm−2). These PMOFs, especially Co-PMOF, exhibited excellent electrocatalytic performance in CO2 reduction.
Davethu et al. [77] studied the electrochemical reduction of CO2 to CO on an iron–porphyrin center using a computational modeling. The results showed that a ligand, rather than metal reduction, resulted in stable binding of CO2 as an [FeIII (CO22−) (TPP)]2− complex during the reduction process. Subsequent proton transfer from phenol was considered as a proton-coupled electron-transfer process, and the second proton transfer does not change the electronic configuration of the metal complex. It was demonstrated that iron porphyrin was an effective catalyst and could efficiently transform CO2 to CO. In addition, the results indicated that CO2 binding was the rate-determining step in the reaction cycle, providing a promising direction for further optimization. A series of skeletons based on porous metal–porphyrin triazine were synthesized by trimming porphyrin units under ionothermal conditions [78]. The skeletons have high specific surface areas and homogenously dispersed transition metals. This facilitates the adsorption of CO2 molecules and exposure of larger number of active sites, thereby improving the performance of CO2 reduction.
Reducing CO2 is, energetically, an uphill process. Many effective and selective homogeneous metal complex catalysts have been developed to promote CO2 conversion. For example, a number of pincer complexes can reduce CO2 into CO, CH4, or other compounds [79,80,81,82]. The ruthenium catalysts prepared by different methods are usually used in such systems [83,84,85,86] in which the turnover number (TON) of CO2 to methanol reached 9900 at an optimal condition. All of this shows the good conductivity and catalytic performance of pincer complexes, but these catalysts often rely on precious metals or supported ligands, such as bipyridines. Most catalysts used in CO2 reduction have the ability of hydrogen evolution, resulting in H2 generation during the conversion of CO2. This will suppress the formation of the desired products because of competition in electron capturing with the H2 evolution reaction. It is difficult to find a suitable electrocatalyst to selectively improve the conversion efficiency of CO2 [87]. Furthermore, multi-electron transfer process and electron transfer rate are the critical limiting factors in the electrocatalytic reaction. The main challenge is to find a highly selective electrocatalyst with high catalytic activity and long-term stability to overcome the thermodynamic stability of CO2 molecules. If electrocatalysis and photocatalysis are combined to take full advantage of their respective advantages, seem to be a better method to reduce CO2.

2.4. Photoelectrocatalysis

Since there is an inexhaustible solar energy supply in nature, it should be fully utilized in different ways. Photoelectrocatalysis, which combines the advantages of photocatalysis and electrocatalysis, is considered to be an ideal strategy for the selective conversion of CO2 into gaseous (such as CO, methane, etc.) and liquid products (such as formic acid, methanol, ethanol, etc.) under sunlight irradiation, and has therefore attracted great attention [88,89,90].
Photoelectrocatalysis makes the best use of solar energy to produce photoelectrons. The photogenerated electrons are transferred to the electrode surface under the action of an applied electric field, and finally obtained by CO2 for catalytic reduction. The applied electric field can effectively facilitate charge separation in the photocatalytic process [59], promote electron migration, and significantly improve the intrinsic activity and energy efficiency of CO2 molecules [91]. The efficient utilization of solar energy in photoelectrocatalysis can effectively overcome the problem of high energy consumption in the electrocatalysis of CO2.
In order to promote rapid charge transfer and improve the performance of photoelectrocatalysis, Ding and coauthors [92] patterned a photocathode through photolithography to expose a third of the surface, which is an effective and robust Si–Bi interface formed by Bi3+-assisted chemical etching of Si wafers and completed the reduction of CO2. This method increased the current density and facilitated the reduction of CO2 based on high product selectivity.
TiO2 is one of the most employed semiconductor in photo-assisted processes. Castro et al. [93] loaded different amounts of TiO2 on the photoanode using a Cu plate as the photocathode to build a photoelectric chemical device, and combined this with an electrochemical filter-press cell. This device was employed to continuously convert CO2 into alcohol with reducing energy consumption due to less external energy demand. Comparing the alcohol produced under different conditions, the TiO2 photoanode system exhibited enhanced alcohol production and reduced energy consumption under ultraviolet light irradiation.
Different photocathodes have different light absorption capabilities, which essentially depend on the optical characteristics of the semiconductor. Table 1 lists and compares the performance of different photoelectrochemical systems of CO2 reduction from the latest literature.
The efficiency of CO2 conversion is a criterion of photoelectric conversion efficiency, which can be calculated by the following equation.
Faradaic Efficiency (FE): FE can be understood as the percentage of actual product/theoretical product.
  F E ( % ) = e output e input × 100 = n ( mol ) × m Q ( Coulomb ) F ( Coulomb / mol ) × 100
In the above equation, n is the actual moles of product, m is the number of reaction electrons, Q is the calculated electric charge, and F is the Faraday constant (96,485 C/mol).
Applied Bias Photon-to-Current Efficiency (ABPE): ABPE is used to measure the efficiency when an external voltage ( V bias ) is applied.
A B P E = J ph ( mA / cm 2 ) × [ E ° ( V ) V bias ( V ) ] × F E P solar ( mW / cm 2 )
where J ph is the photocurrent detected under the external voltage, E ° is the thermodynamic energy stored in the PEC reactor, and P solar is the power density of light.

2.5. Enzyme

As is known, the catalyst is one of the key components in CO2 reduction systems including thermal catalysis, electrocatalysis, photocatalysis, and photoelectrocatalysis. However, a prominent problem associated with most catalytic systems is low product selectivity, where more than one product, including CO, formate, methane, ethylene, and other components, are usually observed in one catalysis reaction. By contrast, the reduction of CO2 via biocatalytic processes received particular attention because of their special substrate and product selectivity as well as high conversion efficiency.
Enzymes are biocatalysts renowned for their high efficiency and selectivity. In living cells, different enzymes often work together or in a specific order to catalyze multi-step biochemical reactions, playing crucial roles in the synthesis of natural products and metabolism [103]. Inspired by the biocatalytic reaction, enzymes including enzyme cascades were explored in vitro to complete the conversion of CO2 to certain chemicals via a one-step or multi-step process. Figure 5 shows the approximate number of papers published in the past two decades using enzymes as catalysts in CO2 conversion. It is obvious that the research has presented an increasing tendency, especially in the recent 10 years, suggesting more and more attention was paid to the biocatalytic conversion of CO2.
In 1993 and 1994, Yoneyam et al. [104,105] demonstrated that CO2 can be biocatalyzed into CH3OH in a CO2-saturated phosphate buffer solution, in which pyrroloquinoline quinone (or methyl viologen) was used as an electron mediator, and formate dehydrogenase, formaldehyde dehydrogenase, and alcohol dehydrogenase were used as biocatalysts. Subsequently, Obert [106] presented the reduction of CO2 to methanol using three different dehydrogenases in three consequent reductions, in which reduced nicotinamide adenine dinucleotide (NADH) molecules were required at each step. Such a multi-enzyme system was composed of three different dehydrogenases (Figure 6) that catalyze the conversion of CO2 to CH3OH in the presence of NADH. In this enzyme cascade, formate dehydrogenase (FDH) catalyzes the conversion of CO2 to formate, formaldehyde dehydrogenase (FaldDH) then catalyzes the formate to formaldehyde and, finally, alcohol dehydrogenase (ADH) catalyzes the formaldehyde to CH3OH. Each enzymatic step in the reduction cascade proceeds in the opposite direction of the natural (reversible) enzyme-catalyzed reaction and requires NADH as the electron donor for the reaction.
Researchers [107] also compared CO2 reduction from different sources of FDH, FaldDH, and ADH to gain an in-depth understanding of the multi-enzyme cascade reaction. The formate dehydrogenase (ClFDH), formaldehyde dehydrogenase (BmFaldDH), and alcohol dehydrogenase (YADH) were from Clostridium ljungdahlii, Burkholderia multivorans, and Saccharomyces cerevisiae, respectively. A 500-fold increase in total turnover number was observed for the ClFDH–BmFaldDH–YADH cascade system compared to the Candida boidinii FDH–Pseudomonas putida FaldDH–YADH system. This is conducive to develop an enzyme cascade reaction with higher conversion efficiency. The three dehydrogenases can not only be combined to convert CO2 into methanol but also can be used individually to convert CO2 to corresponding products such as formate or formaldehyde.
Up to now, dehydrogenases including formate (FDH), formaldehyde (FaldDH), and alcohol dehydrogenase (ADH) that are usually used as biocatalysts in CO2 reduction are NAH(P)H dependent. However, NADH is expensive, and the extensive use of NADH increased the cost of enzymatic reaction. The regeneration efficiency of NAD(P)H became an important criterion for evaluating the biocatalytic reaction and more efforts should be devoted to improving the NAD(P)H yield and reducing the production cost.
The combination of enzymes and photocatalysts for CO2 conversion has attracted increasing attention because it makes full use of the abundant energy supply of solar light and high specificity of enzyme catalysis [108,109,110]. These reactions can be conducted at mild conditions similar to the photosynthesis that occurs in plants or certain bacteria. This is also called artificial photosynthesis. The combination of electrodes with suitable biological enzymes could minimize the requirement of high overpotentials to excite the electrocatalytic reaction of CO2. Such an electroenzymatic reaction can be carried out at a low overpotential, or even without external bias [111,112,113].

2.6. Enzyme Coupled to Photocatalysis

The photosynthesis that occurs in green plants and certain bacteria converts solar energy into chemical energy that can be well utilized by organisms and, at the same time, absorbs carbon dioxide and produces oxygen to maintain the carbon–oxygen cycle on the earth. This inspired people to explore the intrinsic mechanism of the photosynthesis process and to construct artificial analogues via biomimetic mythologies for alternative sustainable energy carriers instead of traditional fossil fuels. Conversion of solar energy to chemical energy consists of hydrogen production, oxygen evolution, and carbon dioxide reduction as well as nitrogen fixation. Among them, the conversion of the human-made greenhouse gas CO2 into valuable fuels/chemicals using solar energy is considered a promising and compelling approach to solar energy utilization because it aims to simultaneously solve problems regarding global energy and environment [114]. In particular, photoreaction coupled with enzymes provides a highly efficient, specific, and energy saving strategy for CO2 conversion and has attracted special attention in recent years [115].
Yadav and coauthors [116] developed a graphene-based visible light active photocatalyst–FDH coupled system in which CO2 was specifically converted to formic acid. The chromophore (multianthraquinone-substituted porphyrin, MAQSP) with chemically modified graphene (CCG) were covalently combined to form a new catalyst (CCGMAQSP), through which the light-harvesting efficiency can be enhanced. The chromophore absorbs sunlight and acts as an electron donor. The light-generated electrons are transferred to the organometallic rhodium complex through graphene (electron acceptor). The rhodium complex accepts electrons from graphene and thus is reduced and further extracts H+ from water. NAD+ accepts electrons and H+ to form NADH [89]. Formate dehydrogenase converts CO2 to formate in the presence of NADH, as shown in Figure 7.
In addition to rhodium complexes, methyl viologen (MV) was also used as an electron mediator. Kumar and coauthors [117] designed a photocatalyst of graphene oxide modified with cobalt metallized aminoporphyrin (GO-Co-ATPP) for conversion of CO2 to formic acid under visible light. Porphyrin captures photons and generates electrons and transfers to methyl viologen (MV) complex via graphene. The organometallic MV complex can easily obtain electrons and exist in its reduced form. It further extracts protons from the aqueous solution, and transfers electrons and hydrogen ions to NAD+, which finally transforms to NADH, which is used for CO2 reduction.
However, this system complete the photocatalytic reaction and enzymatic reaction in the same environment, which will have a certain impact on the stability and activity of enzyme. For example, FDH has attracted much attention in recent years because it can directly reduce CO2 to formate without any other byproducts. FDH is divided into two types according to cofactor requirements: NADH-dependent FDH and metal-dependent FDH. Although metal-containing FDHs have a higher catalytic activity for CO2 reduction, these NADH-independent FDHs contain extremely unstable oxygen components, such as metal ions (tungsten or molybdenum), iron–sulfur clusters, and selenocysteine. The oxidation reaction of water may occur during photocatalysis to generate oxygen, which will affect the enzyme catalytic activity in the system, affect the stability of FDH, and then affect the final product of formic acid. The low compatibility of the photocatalysis and the biocatalysis in the system hindered its development.
As is known, thylakoids in chloroplast were employed to couple the photoreaction and the biological reaction system by which the enzymatic reaction was separated from the water oxidation reaction to protect enzymes from inactivation. In order to achieve cooperation between photocatalysis and biocatalysis and improve compatibility, Zhan et al. [118] developed an artificial thylakoid by decorating the inner wall of protamine–titania (PTi) microcapsules with cadmium sulfide quantum dots (CdS QDs), and coupled with biocatalysis to form an artificial photosynthesis system. Cds QDs absorb visible light and generate electrons and holes. The electrons are transferred to the outer surface of the capsule through the heterostructure of Cds QDs and amorphous TiO2. Through the intermediate transfer of the rhodium complex, formate dehydrogenase converts CO2 to formic acid in the presence of NADH. The size-selective capsule wall separates photocatalytic oxidation and enzymatic reduction of CO2, thereby protecting the enzyme from inactivation that usually caused by photogenerated holes and active oxygen.
Most enzymes are powdered reagents, which makes them difficult to separate from the substrate and cannot be recycled. It can effectively reduce the costs and simplify the product purification process through enzyme immobilization. Zeolite imidazolate framework (ZIF) is a type of MOF material that possesses well-defined pore structure, excellent chemical–thermal stability, extremely high surface area, and other excellent properties [119]. Moreover, ZIF is easy to prepare and has little effect on enzyme activity because the preparation is usually conducted at mild conditions. It has become one of the common methods for enzyme immobilization [120]. Zhou et al. [121] combined ZIF and TCPP to construct a photocatalytic multi-enzyme cascade biomimetic carbon sequestration system (TCPP&FF@ZIF-8 (FF = FateDH and FaldDH)). TCPP was used as the photocatalyst and ZIF-8 as the multi-enzyme immobilized carrier for FateDH and FaldDH. The catalytic system was then used to absorb CO2 and transform it to chemicals. Interestingly, the repeated stability investigation of the composite system showed that the residual activity of 3% TCPP and FF@ZIF-8 remained at 52.93 % after 10 batches of repeated use, suggesting that the system had excellent structural stability, light stability, and cycling stability.
Using different photocatalysts, enzymes, and cofactors, various products and yields were obtained. Table 2 provides a simple comparison of the performance of different coupled photocatalytic/enzymatic CO2 reduction systems in recent years.
The combination of photocatalysis and biocatalysis showed higher efficiency in CO2 reduction than that of photocatalysis. However, the prominent problem is the interference between the photocatalysis and biocatalysis, resulting in corrosion of the photocatalyst and inactivation of the enzyme. Additionally, the enzyme used in the biocatalysis is reversible and it is easy to perform the reverse reaction and charge reorganization [128]. To solve this problem, researchers adopted an electric field on the basis of photoenzymatic catalysis, which can effectively promote charge separation and improve the conversion efficiency of CO2.

2.7. Enzyme Coupled to Photoelectrocatalysis

Photocatalysis and electrocatalysis are combined to form a photoelectric cooperative catalysis, and then combined with biocatalysis to form a photoelectrochemical (PEC) cell similar to maintain the optimal conditions of enzymes and improve conversion efficiency [129,130]. Conducting wire was used to ensure the oriented transfer of reducing equivalents (primarily electrons, H+) from the photoelectrode to biocatalysis. Nam et al. [131] imagined that photoelectrochemical (PEC) cells could inhibit biocatalytic charge recombination and reverse reactions because photocatalytic and biocatalytic reactions can be separated in the anode and cathode compartments, respectively. An anode compartment with cobalt phosphate (Co-Pi) deposited hematite (Fe2O3) photoanode for photocatalytic water oxidation and a cathode compartment with formate dehydrogenase for NADH regeneration and CO2 reduction were designed (Figure 8). The co-catalyst (Co-Pi) in the photoelectrode can reduce the activation energy quickly, improve the quantum efficiency by promoting charge separation, and consume the photogenerated charges in time to improve the stability of the photoelectrode [132]. As can be seen from Figure 8, the PEC system is divided into two compartments in which the oxidation reaction of water can be well separated from the enzymatic reaction. The two compartments do not affect each other, enabling the formate dehydrogenase to work at its optimal pH conditions.
Generally, enzymes are easily affected by the reaction environment, making the enzyme electrode unstable [133]. For example, the enzyme cannot perform its maximum activity at a non-optimal pH solution. Eun-Gyu Choi and coauthors [134] studied the effect of pH on a coupled photoelectric–enzyme system. RcFDH-driven CO2 reduction was predominant at an acidic pH, whereas formate oxidation was favorable at basic conditions. pH = 6.5 is the most suitable condition for CO2 reduction to formic acid in their photoenzyme system by the comparison of the results at different pH values. For the stability and reusability of the enzyme, appropriate fixation methods can be adopted. Lee et al. [112] reported that a tightly organized biophotocathode (EC-PDA)-electrochemically synthesized polydopamine (PDA) film was copolymerized with FDH (E) and NADH (C) in which CO2 can be reduced to formic acid with high selectivity. The PDA was chosen as the substrate for enzyme immobilization because of its excellent biocompatibility and charge transfer ability [135]. The PDA layer on the electrode can fulfill the requirement of electron transfer and enzyme stabilization and extend the service life of the enzyme [136]. A similar photoelectrochemical cell was constructed by the EC-incorporated PDA bioelectrode and cobalt phosphate/bismuth vanadate (CoPi/BiVO4) photoanode by which the reduction of CO2 can be achieved without external bias.
In another study, a voltage was applied to the constructed PEC cell-Co-Pi/Fe2O3 photoanode and BiFeO3 photocathode [137]. The polarization treatment drove surface charge accumulation and accelerated the transfer of electrons to the electrolyte, therefore resulting in an improvement in CO2 conversion efficiency. The tandem PEC cell with an integrated enzyme cascade (TPIEC) system mimics the natural photosynthetic Z-scheme for the biocatalytic reduction of CO2 to methanol. The rate of methanol production per unit of reaction volume and the rate of methanol production per unit mass of photocatalyst can reach 220 μM h−1 and 220 μM gcat−1 h−1, respectively. This device exhibited significantly higher rate of methanol than those of other studies. Due to the relatively high price of the metal rhodium complex [138], researchers further improved the structure of the photoelectrochemical cell to reduce the cost. Neutral red was used as an alternative electron mediator to replace the metal rhodium complex and was conducive to electron recycling between the electrode and NAD+ [139]. Sokol and coauthors [140] adopted a semi-artificial design. The cathode containing formate dehydrogenase was connected to the photoanode containing photosynthetic water oxidase (photosystem II) to achieve the metabolic pathway of formic acid that is formed by light fixation of CO2 in the absence of precious metal catalyst (Figure 9). This demonstrated the feasibility of the nonmetal catalysts for the conversion of CO2 to formic acid and provides a novel method for CO2 photoelectrochemical reduction. Table 3 lists the performance parameters of different photoelectrochemical/enzyme systems.
Researchers not only improved the noble metal mediator, but also further studied the coenzyme NADH. NADH is necessary in most biocatalytic reductions of CO2 to provide electrons for dehydrogenase. NADH is expensive and easily forms enzymatically inactive dimers, NAD2, resulting in reduction of enzyme activity [143]. It is therefore necessary to develop economical methods for NADH cofactor regeneration. To achieve this goal, chemical, photochemical, and electrochemical regeneration of NADH has been developed over the last few decades [122,144,145,146]. Electrochemically mediated electron injection into the enzyme not only bypasses the requirement of NADH but also simplifies product separation [147]. Amao and Shuto [148] have demonstrated the use of viologen-modified FDH immobilized on an ITO electrode to electrochemically convert CO2 to formic acid. This study showed evidence of electrons directly transferring to FDH active sites from electrodes and represents a new strategy for CO2 reduction in the absence of NADH. However, finding alternative compounds to replace NADH is also an attractive research direction. Hence, researchers have tried to use cheaper electron mediators, such as methyl viologen (MV2+), instead of NADH. Miyatani et al. [149] developed a system using MV which combined the synthesis of formic acid from CO2 (bicarbonate ion) with FDH and MV2+, and photoreduction with ZnTMPyP as photosensor and TEOA as electron donor. Formic acid was successfully formed from bicarbonate ions and FDH in the absence of NADH.
Moreover, investigations have demonstrated that the oxidation of formate to carbon dioxide does not occur readily with MV2+. Therefore, the overall yield of formate would be preserved without loss from reoxidation [150]. Ishibashi et al. [151] used methyl viologen (MV2+) instead of NADH. A photoelectrochemical system was composed of TiO2 nanoparticles as photocatalyst, MV2+ as electron carrier, and FDH as biocatalyst, in which CO2 was successfully reduced to formic acid without sacrificing reagent, external bias, and NADH.

3. Conclusions and Outline

Thermal catalysis, photocatalysis, electrocatalysis, photoelectrocatalysis, and enzyme catalysis can effectively alleviate the greenhouse gas CO2. The catalyst is the key component of different CO2 reducing systems. A suitable catalyst can not only reduce energy consumption but also facilitate generation and transfer of electrons. On one side, from the perspective of energy consumption, we wish to complete the conversion of CO2 while consuming as little energy source as possible to save energy and the environment. Therefore, photocatalysis and photoelectrocatalysis that adopt clean and sustainable solar energy to drive the conversion are of interest, in which porphyrin-based macrocycles or their combination with other components presented promising properties of light-capturing and charge transferring. On the other side, the selectivity of products has challenged the development of CO2 conversion. Specific product generation from CO2 will greatly reduce the cost of subsequent product separation and purification. This has inspired researchers to pay particular attention to biocatalysts because of their high specificity and efficiency in catalyzing biochemical reactions. Among the aforementioned methods, enzyme coupled to photocatalysis and enzyme coupled to photoelectrocatalysis has integrated the two sides successfully, showing great potential in solar energy utilization and specific conversion of CO2. They are worthy of more investigation to make biocatalysis compatible with photocatalysis or photoelectrocatalysis. In general, the conversion of CO2 to valuable fuels or chemicals appears to have a bright future, and continuous efforts are needed to improve the catalytic efficiency, conversion rate, and product selectivity.

Author Contributions

Conceptualization, S.W. and L.X.; software, L.X. and Y.X.; validation, S.W.; resources, S.W.; date curation, F.L. and Y.L.; writing—original draft preparation, L.X. and Y.X.; writing—review and editing, S.W.; visualization, S.W. and F.L.; supervision, S.W.; project administration, S.W.; funding acquisition, S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (21773310) and Key R. & D. Program of Shandong Province (2019GGX103047).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hansen, J.; Sato, M.; Ruedy, R.; Lo, K.; Lea, D.W.; Medina-Elizade, M. Global temperature change. Proc. Natl. Acad. Sci. USA 2006, 103, 14288–14293. [Google Scholar] [CrossRef] [Green Version]
  2. Zachos, J.C.; Dickens, G.; Zeebe, R. An Early Cenozoic perspective on Greenhouse warming and carbon cycle dynamics. Nature 2008, 451, 279–283. [Google Scholar] [CrossRef]
  3. Hepp, S.; Jetter, M.; Portalupi, S.L.; Michler, P. Semiconductor Quantum Dots for Integrated Quantum Photonics. Adv. Quantum Technol. 2019, 2, 1900020. [Google Scholar] [CrossRef]
  4. Li, L.; Zhao, N.; Wei, W.; Sun, Y. A review of research progress on CO2 capture, storage, and utilization in Chinese Academy of Sciences. Fuel 2013, 108, 112–130. [Google Scholar] [CrossRef]
  5. Yu, K.M.K.; Curcic, I.; Gabriel, J.; Tsang, S.C.E. Recent Advances in CO2 Capture and Utilization. ChemSusChem 2008, 1, 893–899. [Google Scholar] [CrossRef] [PubMed]
  6. Arellano-Treviño, M.A.; Kanani, N.; Jeong-Potter, C.W.; Farrauto, R.J. Bimetallic catalysts for CO2 capture and hydrogenation at simulated flue gas conditions. Chem. Eng. J. 2019, 375. [Google Scholar] [CrossRef]
  7. Kar, S.; Goeppert, A.; Prakash, G.K.S. Combined CO2 Capture and Hydrogenation to Methanol: Amine Immobilization Enables Easy Recycling of Active Elements. ChemSusChem 2019, 12, 3172–3177. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, N.; Long, R.; Gao, C.; Xiong, Y. Recent progress on advanced design for photoelectrochemical reduction of CO2 to fuels. Sci. China Mater. 2018, 61, 771–805. [Google Scholar] [CrossRef] [Green Version]
  9. Kunkes, E.L.; Studt, F.; Abild-Pedersen, F.; Schlögl, R.; Behrens, M. Hydrogenation of CO2 to methanol and CO on Cu/ZnO/Al2O3: Is there a common intermediate or not? J. Catal. 2015, 328, 43–48. [Google Scholar] [CrossRef] [Green Version]
  10. Rui, N.; Wang, Z.; Sun, K.; Ye, J.; Ge, Q.; Liu, C.-J. CO2 hydrogenation to methanol over Pd/In2O3: Effects of Pd and oxygen vacancy. Appl. Catal. B Environ. 2017, 218, 488–497. [Google Scholar] [CrossRef]
  11. Song, Y.; Zhang, X.; Xie, K.; Wang, G.; Bao, X. High-Temperature CO2 Electrolysis in Solid Oxide Electrolysis Cells: Developments, Challenges, and Prospects. Adv. Mater. 2019, 31, 1902033. [Google Scholar] [CrossRef] [PubMed]
  12. Currie, R.; Mottaghi-Tabar, S.; Zhuang, Y.; Simakov, D.S.A. Design of an Air-Cooled Sabatier Reactor for Thermocatalytic Hydrogenation of CO2: Experimental Proof-of-Concept and Model-Based Feasibility Analysis. Ind. Eng. Chem. Res. 2019, 58, 12964–12980. [Google Scholar] [CrossRef]
  13. Jiang, X.; Nie, X.; Guo, X.; Song, C.; Chen, J.G. Recent Advances in Carbon Dioxide Hydrogenation to Methanol via Heterogeneous Catalysis. Chem. Rev. 2020. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, T.; Meng, X.; Li, P.; Ouyang, S.; Chang, K.; Liu, G.; Mei, Z.; Ye, J. Photoreduction of CO2 over the well-crystallized ordered mesoporous TiO2 with the confined space effect. Nano Energy 2014, 9, 50–60. [Google Scholar] [CrossRef]
  15. Li, K.; Peng, B.; Peng, T. Recent Advances in Heterogeneous Photocatalytic CO2 Conversion to Solar Fuels. ACS Catal. 2016, 6, 7485–7527. [Google Scholar] [CrossRef]
  16. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. TiO2–MnOx–Pt Hybrid Multiheterojunction Film Photocatalyst with Enhanced Photocatalytic CO2-Reduction Activity. ACS Appl. Mater. Interfaces 2019, 11, 5581–5589. [Google Scholar] [CrossRef]
  17. Ran, J.; Jaroniec, M.; Qiao, S.-Z. Cocatalysts in Semiconductor-based Photocatalytic CO2 Reduction: Achievements, Challenges, and Opportunities. Adv. Mater. 2018, 30, 1704649. [Google Scholar] [CrossRef]
  18. 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] [Green Version]
  19. Gao, S.; Lin, Y.; Jiao, X.; Sun, Y.; Luo, Q.; Zhang, W.; Li, D.; Yang, J.; Xie, Y. Partially oxidized atomic cobalt layers for carbon dioxide electroreduction to liquid fuel. Nature 2016, 529, 68–71. [Google Scholar] [CrossRef]
  20. Huang, J.; Buonsanti, R. Colloidal Nanocrystals as Heterogeneous Catalysts for Electrochemical CO2 Conversion. Chem. Mater. 2019, 31, 13–25. [Google Scholar] [CrossRef] [Green Version]
  21. Wang, J.; Ji, Y.; Shao, Q.; Yin, R.; Guo, J.; Li, Y.; Huang, X. Phase and structure modulating of bimetallic CuSn nanowires boosts electrocatalytic conversion of CO2. Nano Energy 2019, 59, 138–145. [Google Scholar] [CrossRef]
  22. Kim, C.; Cho, K.M.; Al-Saggaf, A.; Gereige, I.; Jung, H.-T. Z-scheme Photocatalytic CO2 Conversion on Three-Dimensional BiVO4/Carbon-Coated Cu2O Nanowire Arrays under Visible Light. ACS Catal. 2018, 8, 4170–4177. [Google Scholar] [CrossRef]
  23. Leung, J.J.; Warnan, J.; Ly, K.H.; Heidary, N.; Nam, D.H.; Kuehnel, M.F.; Reisner, E. Solar-driven reduction of aqueous CO2 with a cobalt bis(terpyridine)-based photocathode. Nat. Catal 2019, 2, 354–365. [Google Scholar] [CrossRef]
  24. Chen, J.; Yin, J.; Zheng, X.; Ait Ahsaine, H.; Zhou, Y.; Dong, C.; Mohammed, O.F.; Takanabe, K.; Bakr, O.M. Compositionally Screened Eutectic Catalytic Coatings on Halide Perovskite Photocathodes for Photoassisted Selective CO2 Reduction. ACS Energy Lett. 2019, 4, 1279–1286. [Google Scholar] [CrossRef] [Green Version]
  25. Kalamaras, E.; Belekoukia, M.; Tan, J.Z.Y.; Xuan, J.; Maroto-Valer, M.M.; Andresen, J.M. A microfluidic photoelectrochemical cell for solar-driven CO2 conversion into liquid fuels with CuO-based photocathodes. Faraday Discuss. 2019, 215, 329–344. [Google Scholar] [CrossRef]
  26. Sultana, S.; Sahoo, P.C.; Martha, S.; Parida, K. A review of harvesting clean fuels from enzymatic CO2 reduction. RSC Adv. 2016, 6, 44170–44194. [Google Scholar] [CrossRef]
  27. Rudroff, F.; Mihovilovic, M.D.; Groger, H.; Snajdrova, R.; Iding, H.; Bornscheuer, U.T. Opportunities and challenges for combining chemo- and biocatalysis. Nat. Catal. 2018, 1, 12–22. [Google Scholar] [CrossRef]
  28. Tackett, B.M.; Gomez, E.; Chen, J.G. Net reduction of CO2 via its thermocatalytic and electrocatalytic transformation reactions in standard and hybrid processes. Nat. Catal. 2019, 2, 381–386. [Google Scholar] [CrossRef]
  29. Song, C. Global challenges and strategies for control, conversion and utilization of CO2 for sustainable development involving energy, catalysis, adsorption and chemical processing. Catal. Today 2006, 115, 2–32. [Google Scholar] [CrossRef]
  30. Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjær, C.F.; Hummelshøj, J.S.; Dahl, S.; Chorkendorff, I.; Nørskov, J.K. Discovery of a Ni-Ga catalyst for carbon dioxide reduction to methanol. Nat. Chem. 2014, 6, 320–324. [Google Scholar] [CrossRef]
  31. Martin, O.; Martín, A.J.; Mondelli, C.; Mitchell, S.; Segawa, T.F.; Hauert, R.; Drouilly, C.; Curulla-Ferré, D.; Pérez-Ramírez, J. Indium Oxide as a Superior Catalyst for Methanol Synthesis by CO2 Hydrogenation. Angew. Chem. Int. Ed. 2016, 55, 6261–6265. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, J.; Li, G.; Li, Z.; Tang, C.; Feng, Z.; An, H.; Liu, H.; Liu, T.; Li, C. A highly selective and stable ZnO-ZrO2 solid solution catalyst for CO2 hydrogenation to methanol. Sci. Adv. 2017, 3, e1701290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Rungtaweevoranit, B.; Baek, J.; Araujo, J.R.; Archanjo, B.S.; Choi, K.M.; Yaghi, O.M.; Somorjai, G.A. Copper Nanocrystals Encapsulated in Zr-based Metal–Organic Frameworks for Highly Selective CO2 Hydrogenation to Methanol. Nano Lett. 2016, 16, 7645–7649. [Google Scholar] [CrossRef] [PubMed]
  34. Roldán, L.; Marco, Y.; García-Bordejé, E. Origin of the Excellent Performance of Ru on Nitrogen-Doped Carbon Nanofibers for CO2 Hydrogenation to CH4. ChemSusChem 2017, 10, 1139–1144. [Google Scholar] [CrossRef] [Green Version]
  35. O’Byrne, J.P.; Owen, R.E.; Minett, D.R.; Pascu, S.I.; Plucinski, P.K.; Jones, M.D.; Mattia, D. High CO2 and CO conversion to hydrocarbons using bridged Fe nanoparticles on carbon nanotubes. Catal. Sci. Technol. 2013, 3, 1202–1207. [Google Scholar] [CrossRef] [Green Version]
  36. Feng, Y.; Yang, W.; Chen, S.; Chu, W. Cerium Promoted Nano Nickel Catalysts Ni-Ce/CNTs and Ni-Ce/Al2O3 for CO2 Methanation. Integr. Ferroelectr. 2014, 151, 116–125. [Google Scholar] [CrossRef]
  37. Li, J.; Zhou, Y.; Xiao, X.; Wang, W.; Wang, N.; Qian, W.; Chu, W. Regulation of Ni–CNT Interaction on Mn-Promoted Nickel Nanocatalysts Supported on Oxygenated CNTs for CO2 Selective Hydrogenation. ACS Appl. Mater. Interfaces 2018, 10, 41224–41236. [Google Scholar] [CrossRef] [Green Version]
  38. Wang, W.; Duong Viet, C.; Housseinou, B.; Baaziz, W.; Tuci, G.; Caporali, S.; Nguyen, D.-L.; Ersen, O.; Giambastiani, G.; Pham-Huu, C. Nickel Nanoparticles Decorated Nitrogen-Doped Carbon Nanotubes (Ni/N-CNT); A Robust Catalyst for the Efficient and Selective CO2 Methanation. ACS Appl. Energy Mater. 2018. [Google Scholar] [CrossRef]
  39. Wang, X.; Liu, Y.; Zhu, L.; Li, Y.; Wang, K.; Qiu, K.; Tippayawong, N.; Aggarangsi, P.; Reubroycharoen, P.; Wang, S. Biomass derived N-doped biochar as efficient catalyst supports for CO2 methanation. J. CO2 Util. 2019, 34, 733–741. [Google Scholar] [CrossRef]
  40. Wang, W.; Duong-Viet, C.; Xu, Z.; Ba, H.; Tuci, G.; Giambastiani, G.; Liu, Y.; Truong-Huu, T.; Nhut, J.-M.; Pham-Huu, C. CO2 methanation under dynamic operational mode using nickel nanoparticles decorated carbon felt (Ni/OCF) combined with inductive heating. Catal. Today 2019. [Google Scholar] [CrossRef]
  41. Khalil, M.; Gunlazuardi, J.; Ivandini, T.A.; Umar, A. Photocatalytic conversion of CO2 using earth-abundant catalysts: A review on mechanism and catalytic performance. Renew. Sustain. Energy Rev. 2019, 113, 109246. [Google Scholar] [CrossRef]
  42. Zhang, T.; Lin, W. Metal–organic frameworks for artificial photosynthesis and photocatalysis. Chem. Soc. Rev. 2014, 43, 5982–5993. [Google Scholar] [CrossRef] [PubMed]
  43. Shi, L.; Wang, T.; Zhang, H.; Chang, K.; Ye, J. Electrostatic Self-Assembly of Nanosized Carbon Nitride Nanosheet onto a Zirconium Metal–Organic Framework for Enhanced Photocatalytic CO2 Reduction. Adv. Funct. Mater. 2015, 25, 5360–5367. [Google Scholar] [CrossRef]
  44. Dhakshinamoorthy, A.; Asiri, A.M.; García, H. Metal–Organic Framework (MOF) Compounds: Photocatalysts for Redox Reactions and Solar Fuel Production. Angew. Chem. Int. Ed. 2016, 55, 5414–5445. [Google Scholar] [CrossRef] [PubMed]
  45. Maina, J.W.; Pozo-Gonzalo, C.; Kong, L.; Schütz, J.; Hill, M.; Dumée, L.F. Metal organic framework based catalysts for CO2 conversion. Mater. Horiz. 2017, 4, 345–361. [Google Scholar] [CrossRef]
  46. Wang, P.; Zheng, J.Y.; Zhang, D.; Kang, Y.S. Selective construction of junctions on different facets of BiVO4 for enhancing photo-activity. New J. Chem. 2015, 39, 9918–9925. [Google Scholar] [CrossRef]
  47. Wei, Z.-H.; Wang, Y.-F.; Li, Y.-Y.; Zhang, L.; Yao, H.-C.; Li, Z.-J. Enhanced photocatalytic CO2 reduction activity of Z-scheme CdS/BiVO4 nanocomposite with thinner BiVO4 nanosheets. J. CO2 Util. 2018, 28, 15–25. [Google Scholar] [CrossRef]
  48. Zhao, Y.; Zhang, S.; Shi, R.; Waterhouse, G.I.N.; Tang, J.; Zhang, T. Two-dimensional photocatalyst design: A critical review of recent experimental and computational advances. Mater. Today 2020, 34, 78–91. [Google Scholar] [CrossRef]
  49. Mao, X.; Tang, C.; He, T.; Wijethunge, D.; Yan, C.; Zhu, Z.; Du, A. Computational screening of MN4 (M = Ti–Cu) based metal organic frameworks for CO2 reduction using the d-band centre as a descriptor. Nanoscale 2020, 12, 6188–6194. [Google Scholar] [CrossRef]
  50. Zhu, S.; Guo, L.; Li, P.; Zhang, B.; Zhao, G.; He, T. A computational study on linear and bent adsorption of CO2 on different surfaces for its photoreduction. Catal. Today 2019, 335, 278–285. [Google Scholar] [CrossRef]
  51. Guharoy, U.; Le Saché, E.; Cai, Q.; Reina, T.R.; Gu, S. Understanding the role of Ni-Sn interaction to design highly effective CO2 conversion catalysts for dry reforming of methane. J. CO2 Util. 2018, 27, 1–10. [Google Scholar] [CrossRef]
  52. Li, X.; Guo, T.; Zhu, L.; Ling, C.; Xue, Q.; Xing, W. Charge-modulated CO2 capture of C3N nanosheet: Insights from DFT calculations. Chem. Eng. J. 2018, 338, 92–98. [Google Scholar] [CrossRef]
  53. Jiang, Y.-B.; Sun, Z. Self-assembled porphyrin and macrocycle derivatives: From synthesis to function. MRS Bull. 2019, 44, 167–171. [Google Scholar] [CrossRef] [Green Version]
  54. Wang, S.-S.; Huang, H.-H.; Liu, M.; Yao, S.; Guo, S.; Wang, J.-W.; Zhang, Z.-M.; Lu, T.-B. Encapsulation of Single Iron Sites in a Metal–Porphyrin Framework for High-Performance Photocatalytic CO2 Reduction. Inorg. Chem. 2020, 59, 6301–6307. [Google Scholar] [CrossRef] [PubMed]
  55. Li, X.; Wen, J.; Low, J.; Fang, Y.; Yu, J. Design and fabrication of semiconductor photocatalyst for photocatalytic reduction of CO2 to solar fuel. Sci. China Mater. 2014, 57, 70–100. [Google Scholar] [CrossRef] [Green Version]
  56. Ma, Y.; Wang, Z.; Xu, X.; Wang, J. Review on porous nanomaterials for adsorption and photocatalytic conversion of CO2. Chin. J. Catal. 2017, 38, 1956–1969. [Google Scholar] [CrossRef]
  57. Kočí, K.; Matějů, K.; Obalová, L.; Krejčíková, S.; Lacný, Z.; Plachá, D.; Čapek, L.; Hospodková, A.; Šolcová, O. Effect of silver doping on the TiO2 for photocatalytic reduction of CO2. Appl. Catal. B Environ. 2010, 96, 239–244. [Google Scholar] [CrossRef]
  58. Xu, Q.; Yu, J.; Zhang, J.; Zhang, J.; Liu, G. Cubic anatase TiO2 nanocrystals with enhanced photocatalytic CO2 reduction activity. Chem. Commun. 2015, 51, 7950–7953. [Google Scholar] [CrossRef]
  59. Jang, J.-W.; Cho, S.; Magesh, G.; Jang, Y.J.; Kim, J.Y.; Kim, W.Y.; Seo, J.K.; Kim, S.; Lee, K.-H.; Lee, J.S. Aqueous-Solution Route to Zinc Telluride Films for Application to CO2 Reduction. Angew. Chem. Int. Ed. 2014, 53, 5852–5857. [Google Scholar] [CrossRef]
  60. Nesbitt, N.T.; Ma, M.; Trześniewski, B.J.; Jaszewski, S.; Tafti, F.; Burns, M.J.; Smith, W.A.; Naughton, M.J. Au Dendrite Electrocatalysts for CO2 Electrolysis. J. Phys. Chem. C 2018, 122, 10006–10016. [Google Scholar] [CrossRef]
  61. Lin, S.; Diercks, C.S.; Zhang, Y.-B.; Kornienko, N.; Nichols, E.M.; Zhao, Y.; Paris, A.R.; Kim, D.; Yang, P.; Yaghi, O.M.; et al. Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water. Science 2015, 349, 1208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Duan, X.; Xu, J.; Wei, Z.; Ma, J.; Guo, S.; Wang, S.; Liu, H.; Dou, S. Metal-Free Carbon Materials for CO2 Electrochemical Reduction. Adv. Mater. 2017, 29, 1701784. [Google Scholar] [CrossRef] [PubMed]
  63. Jiao, Y.; Zheng, Y.; Chen, P.; Jaroniec, M.; Qiao, S.-Z. Molecular Scaffolding Strategy with Synergistic Active Centers To Facilitate Electrocatalytic CO2 Reduction to Hydrocarbon/Alcohol. J. Am. Chem. Soc. 2017, 139, 18093–18100. [Google Scholar] [CrossRef] [PubMed]
  64. Seh, Z.W.; Kibsgaard, J.; Dickens, C.F.; Chorkendorff, I.; Nørskov, J.K.; Jaramillo, T.F. Combining theory and experiment in electrocatalysis: Insights into materials design. Science 2017, 355, eaad4998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Lu, Q.; Rosen, J.; Jiao, F. Nanostructured Metallic Electrocatalysts for Carbon Dioxide Reduction. ChemCatChem 2015, 7, 38–47. [Google Scholar] [CrossRef]
  66. Zhang, W.; Hu, Y.; Ma, L.; Zhu, G.; Wang, Y.; Xue, X.; Chen, R.; Yang, S.; Jin, Z. Progress and Perspective of Electrocatalytic CO2 Reduction for Renewable Carbonaceous Fuels and Chemicals. Adv. Sci. 2018, 5, 1700275. [Google Scholar] [CrossRef]
  67. Zhang, L.; Zhao, Z.-J.; Gong, J. Nanostructured Materials for Heterogeneous Electrocatalytic CO2 Reduction and their Related Reaction Mechanisms. Angew. Chem. Int. Ed. 2017, 56, 11326–11353. [Google Scholar] [CrossRef]
  68. Asadi, M.; Kim, K.; Liu, C.; Addepalli, A.V.; Abbasi, P.; Yasaei, P.; Phillips, P.; Behranginia, A.; Cerrato, J.M.; Haasch, R.; et al. Nanostructured transition metal dichalcogenide electrocatalysts for CO2 reduction in ionic liquid. Science 2016, 353, 467. [Google Scholar] [CrossRef] [Green Version]
  69. Zhu, D.D.; Liu, J.L.; Qiao, S.Z. Recent Advances in Inorganic Heterogeneous Electrocatalysts for Reduction of Carbon Dioxide. Adv. Mater. 2016, 28, 3423–3452. [Google Scholar] [CrossRef]
  70. Zhang, H.; Ma, Y.; Quan, F.; Huang, J.; Jia, F.; Zhang, L. Selective electro-reduction of CO2 to formate on nanostructured Bi from reduction of BiOCl nanosheets. Electrochem. Commun. 2014, 46, 63–66. [Google Scholar] [CrossRef]
  71. Liu, J.; Guo, C.; Vasileff, A.; Qiao, S. Nanostructured 2D Materials: Prospective Catalysts for Electrochemical CO2 Reduction. Small Methods 2017, 1, 1600006. [Google Scholar] [CrossRef]
  72. Li, J.-R.; Ma, Y.; McCarthy, M.C.; Sculley, J.; Yu, J.; Jeong, H.-K.; Balbuena, P.B.; Zhou, H.-C. Carbon dioxide capture-related gas adsorption and separation in metal-organic frameworks. Coord. Chem. Rev. 2011, 255, 1791–1823. [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. Shen, J.-Q.; Liao, P.-Q.; Zhou, D.-D.; He, C.-T.; Wu, J.-X.; Zhang, W.-X.; Zhang, J.-P.; Chen, X.-M. Modular and Stepwise Synthesis of a Hybrid Metal–Organic Framework for Efficient Electrocatalytic Oxygen Evolution. J. Am. Chem. Soc. 2017, 139, 1778–1781. [Google Scholar] [CrossRef] [PubMed]
  75. Hu, X.-M.; Rønne, M.H.; Pedersen, S.U.; Skrydstrup, T.; Daasbjerg, K. Enhanced Catalytic Activity of Cobalt Porphyrin in CO2 Electroreduction upon Immobilization on Carbon Materials. Angew. Chem. Int. Ed. 2017, 56, 6468–6472. [Google Scholar] [CrossRef] [PubMed]
  76. 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]
  77. Davethu, P.A.; de Visser, S.P. CO2 Reduction on an Iron-Porphyrin Center: A Computational Study. J. Phys. Chem. A 2019, 123, 6527–6535. [Google Scholar] [CrossRef]
  78. Feng, S.; Zheng, W.; Zhu, J.; Li, Z.; Yang, B.; Wen, Z.; Lu, J.; Lei, L.; Wang, S.; Hou, Y. Porous metal-porphyrin triazine-based frameworks for efficient CO2 electroreduction. Appl. Catal. B Environ. 2020, 270, 118908. [Google Scholar] [CrossRef]
  79. Elgrishi, N.; Chambers, M.B.; Artero, V.; Fontecave, M. Terpyridine complexes of first row transition metals and electrochemical reduction of CO2 to CO. Phys. Chem. Chem. Phys. 2014, 16, 13635–13644. [Google Scholar] [CrossRef]
  80. Rao, G.K.; Pell, W.; Korobkov, I.; Richeson, D. Electrocatalytic reduction of CO2 using Mn complexes with unconventional coordination environments. Chem. Commun. 2016, 52, 8010–8013. [Google Scholar] [CrossRef]
  81. Liu, F.-W.; Bi, J.; Sun, Y.; Luo, S.; Kang, P. Cobalt Complex with Redox-Active Imino Bipyridyl Ligand for Electrocatalytic Reduction of Carbon Dioxide to Formate. ChemSusChem 2018, 11, 1656–1663. [Google Scholar] [CrossRef] [PubMed]
  82. Myren, T.H.T.; Lilio, A.M.; Huntzinger, C.G.; Horstman, J.W.; Stinson, T.A.; Donadt, T.B.; Moore, C.; Lama, B.; Funke, H.H.; Luca, O.R. Manganese N-Heterocyclic Carbene Pincers for the Electrocatalytic Reduction of Carbon Dioxide. Organometallics 2019, 38, 1248–1253. [Google Scholar] [CrossRef]
  83. Wesselbaum, S.; vom Stein, T.; Klankermayer, J.; Leitner, W. Hydrogenation of Carbon Dioxide to Methanol by Using a Homogeneous Ruthenium–Phosphine Catalyst. Angew. Chem. Int. Ed. 2012, 51, 7499–7502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Kothandaraman, J.; Goeppert, A.; Czaun, M.; Olah, G.A.; Prakash, G.K.S. Conversion of CO2 from Air into Methanol Using a Polyamine and a Homogeneous Ruthenium Catalyst. J. Am. Chem. Soc. 2016, 138, 778–781. [Google Scholar] [CrossRef] [PubMed]
  85. Everett, M.; Wass, D.F. Highly productive CO2 hydrogenation to methanol—A tandem catalytic approach via amide intermediates. Chem. Commun. 2017, 53, 9502–9504. [Google Scholar] [CrossRef] [PubMed]
  86. Kar, S.; Sen, R.; Kothandaraman, J.; Goeppert, A.; Chowdhury, R.; Munoz, S.B.; Haiges, R.; Prakash, G.K.S. Mechanistic Insights into Ruthenium-Pincer-Catalyzed Amine-Assisted Homogeneous Hydrogenation of CO2 to Methanol. J. Am. Chem. Soc. 2019, 141, 3160–3170. [Google Scholar] [CrossRef]
  87. Sun, Z.; Ma, T.; Tao, H.; Fan, Q.; Han, B. Fundamentals and Challenges of Electrochemical CO2 Reduction Using Two-Dimensional Materials. Chem 2017, 3, 560–587. [Google Scholar] [CrossRef] [Green Version]
  88. Sahara, G.; Kumagai, H.; Maeda, K.; Kaeffer, N.; Artero, V.; Higashi, M.; Abe, R.; Ishitani, O. Photoelectrochemical Reduction of CO2 Coupled to Water Oxidation Using a Photocathode with a Ru(II)–Re(I) Complex Photocatalyst and a CoOx/TaON Photoanode. J. Am. Chem. Soc. 2016, 138, 14152–14158. [Google Scholar] [CrossRef]
  89. Song, J.T.; Ryoo, H.; Cho, M.; Kim, J.; Kim, J.-G.; Chung, S.-Y.; Oh, J. Nanoporous Au Thin Films on Si Photoelectrodes for Selective and Efficient Photoelectrochemical CO2 Reduction. Adv. Energy Mater. 2017, 7, 1601103. [Google Scholar] [CrossRef]
  90. DuChene, J.S.; Tagliabue, G.; Welch, A.J.; Cheng, W.-H.; Atwater, H.A. Hot Hole Collection and Photoelectrochemical CO2 Reduction with Plasmonic Au/p-GaN Photocathodes. Nano Lett. 2018, 18, 2545–2550. [Google Scholar] [CrossRef] [Green Version]
  91. Neyts, E.C.; Ostrikov, K.; Sunkara, M.K.; Bogaerts, A. Plasma Catalysis: Synergistic Effects at the Nanoscale. Chem. Rev. 2015, 115, 13408–13446. [Google Scholar] [CrossRef] [PubMed]
  92. Ding, P.; Hu, Y.; Deng, J.; Chen, J.; Zha, C.; Yang, H.; Han, N.; Gong, Q.; Li, L.; Wang, T.; et al. Controlled chemical etching leads to efficient silicon–bismuth interface for photoelectrochemical CO2 reduction to formate. Mater. Today Chem. 2019, 11, 80–85. [Google Scholar] [CrossRef]
  93. Castro, S.; Albo, J.; Irabien, A. Continuous conversion of CO2 to alcohols in a TiO2 photoanode-driven photoelectrochemical system. J. Chem. Technol. Biotechnol. 2020, 95, 1876–1882. [Google Scholar] [CrossRef]
  94. Gurudayal; Beeman, J.W.; Bullock, J.; Wang, H.; Eichhorn, J.; Towle, C.; Javey, A.; Toma, F.M.; Mathews, N.; Ager, J.W. Si photocathode with Ag-supported dendritic Cu catalyst for CO2 reduction. Energy Environ. Sci. 2019, 12, 1068–1077. [Google Scholar] [CrossRef] [Green Version]
  95. Choi, S.K.; Kang, U.; Lee, S.; Ham, D.J.; Ji, S.M.; Park, H. Sn-Coupled p-Si Nanowire Arrays for Solar Formate Production from CO2. Adv. Energy Mater. 2014, 4, 1301614. [Google Scholar] [CrossRef]
  96. Rajeshwar, K.; de Tacconi, N.R.; Ghadimkhani, G.; Chanmanee, W.; Janáky, C. Tailoring Copper Oxide Semiconductor Nanorod Arrays for Photoelectrochemical Reduction of Carbon Dioxide to Methanol. ChemPhysChem 2013, 14, 2251–2259. [Google Scholar] [CrossRef]
  97. Shan, B.; Vanka, S.; Li, T.-T.; Troian-Gautier, L.; Brennaman, M.K.; Mi, Z.; Meyer, T.J. Binary molecular-semiconductor p–n junctions for photoelectrocatalytic CO2 reduction. Nat. Energy 2019, 4, 290–299. [Google Scholar] [CrossRef]
  98. Rao, K.R.; Pishgar, S.; Strain, J.; Kumar, B.; Atla, V.; Kumari, S.; Spurgeon, J.M. Photoelectrochemical reduction of CO2 to HCOOH on silicon photocathodes with reduced SnO2 porous nanowire catalysts. J. Mater. Chem. A 2018, 6, 1736–1742. [Google Scholar] [CrossRef]
  99. Huang, X.; Shen, Q.; Liu, J.; Yang, N.; Zhao, G. A CO2 adsorption-enhanced semiconductor/metal-complex hybrid photoelectrocatalytic interface for efficient formate production. Energy Environ. Sci. 2016, 9, 3161–3171. [Google Scholar] [CrossRef] [Green Version]
  100. Liu, J.; Guo, C.; Hu, X.; Zhao, G. Bio-proton coupled semiconductor/metal-complex hybrid photoelectrocatalytic interface for efficient CO2 reduction. Green Chem. 2019, 21, 339–348. [Google Scholar] [CrossRef]
  101. Li, P.; Jing, H.; Xu, J.; Wu, C.; Peng, H.; Lu, J.; Lu, F. High-efficiency synergistic conversion of CO2 to methanol using Fe2O3 nanotubes modified with double-layer Cu2O spheres. Nanoscale 2014, 6, 11380–11386. [Google Scholar] [CrossRef] [PubMed]
  102. Yang, X.; Fugate, E.A.; Mueanngern, Y.; Baker, L.R. Photoelectrochemical CO2 Reduction to Acetate on Iron–Copper Oxide Catalysts. ACS Catal. 2017, 7, 177–180. [Google Scholar] [CrossRef]
  103. Ardao, I.; Hwang, E.T.; Zeng, A.-P. In Vitro Multienzymatic Reaction Systems for Biosynthesis. In Fundamentals and Application of New Bioproduction Systems; Zeng, A.-P., Ed.; Springer: Berlin/Heidelberg, Germany, 2013; pp. 153–184. [Google Scholar] [CrossRef]
  104. Kuwabata, S.; Tsuda, R.; Nishida, K.; Yoneyama, H. Electrochemical Conversion of Carbon Dioxide to Methanol with Use of Enzymes as Biocatalysts. Chem. Lett. 1993, 22, 1631–1634. [Google Scholar] [CrossRef]
  105. Kuwabata, S.; Tsuda, R.; Yoneyama, H. Electrochemical conversion of carbon dioxide to methanol with the assistance of formate dehydrogenase and methanol dehydrogenase as biocatalysts. J. Am. Chem. Soc. 1994, 116, 5437–5443. [Google Scholar] [CrossRef]
  106. Obert, R.; Dave, B.C. Enzymatic Conversion of Carbon Dioxide to Methanol: Enhanced Methanol Production in Silica Sol−Gel Matrices. J. Am. Chem. Soc. 1999, 121, 12192–12193. [Google Scholar] [CrossRef]
  107. Singh, R.K.; Singh, R.; Sivakumar, D.; Kondaveeti, S.; Kim, T.; Li, J.; Sung, B.H.; Cho, B.-K.; Kim, D.R.; Kim, S.C.; et al. Insights into Cell-Free Conversion of CO2 to Chemicals by a Multienzyme Cascade Reaction. ACS Catal. 2018, 8, 11085–11093. [Google Scholar] [CrossRef]
  108. Appel, A.M.; Bercaw, J.E.; Bocarsly, A.B.; Dobbek, H.; DuBois, D.L.; Dupuis, M.; Ferry, J.G.; Fujita, E.; Hille, R.; Kenis, P.J.A.; et al. Frontiers, Opportunities, and Challenges in Biochemical and Chemical Catalysis of CO2 Fixation. Chem. Rev. 2013, 113, 6621–6658. [Google Scholar] [CrossRef] [Green Version]
  109. Schlager, S.; Dibenedetto, A.; Aresta, M.; Apaydin, D.H.; Dumitru, L.M.; Neugebauer, H.; Sariciftci, N.S. Biocatalytic and Bioelectrocatalytic Approaches for the Reduction of Carbon Dioxide using Enzymes. Energy Technol. 2017, 5, 812–821. [Google Scholar] [CrossRef]
  110. Fu, Q.; Xiao, S.; Li, Z.; Li, Y.; Kobayashi, H.; Li, J.; Yang, Y.; Liao, Q.; Zhu, X.; He, X.; et al. Hybrid solar-to-methane conversion system with a Faradaic efficiency of up to 96%. Nano Energy 2018, 53, 232–239. [Google Scholar] [CrossRef]
  111. Aresta, M.; Dibenedetto, A.; Pastore, C. Biotechnology to develop innovative syntheses using CO2. Environ. Chem. Lett. 2005, 3, 113–117. [Google Scholar] [CrossRef] [Green Version]
  112. Lee, S.Y.; Lim, S.Y.; Seo, D.; Lee, J.-Y.; Chung, T.D. Light-Driven Highly Selective Conversion of CO2 to Formate by Electrosynthesized Enzyme/Cofactor Thin Film Electrode. Adv. Energy Mater. 2016, 6, 1502207. [Google Scholar] [CrossRef]
  113. Kuk, S.K.; Ham, Y.; Gopinath, K.; Boonmongkolras, P.; Lee, Y.; Lee, Y.W.; Kondaveeti, S.; Ahn, C.; Shin, B.; Lee, J.-K.; et al. Continuous 3D Titanium Nitride Nanoshell Structure for Solar-Driven Unbiased Biocatalytic CO2 Reduction. Adv. Energy Mater. 2019, 9, 1900029. [Google Scholar] [CrossRef]
  114. El-Khouly, M.E.; El-Mohsnawy, E.; Fukuzumi, S. Solar energy conversion: From natural to artificial photosynthesis. J. Photochem. Photobiol. C Photochem. Rev. 2017, 31, 36–83. [Google Scholar] [CrossRef]
  115. Zhang, S.; Shi, J.; Chen, Y.; Huo, Q.; Li, W.; Wu, Y.; Sun, Y.; Zhang, Y.; Wang, X.; Jiang, Z. Unraveling and Manipulating of NADH Oxidation by Photogenerated Holes. ACS Catal. 2020, 10, 4967–4972. [Google Scholar] [CrossRef]
  116. Yadav, R.K.; Baeg, J.-O.; Oh, G.H.; Park, N.-J.; Kong, K.-J.; Kim, J.; Hwang, D.W.; Biswas, S.K. A Photocatalyst–Enzyme Coupled Artificial Photosynthesis System for Solar Energy in Production of Formic Acid from CO2. J. Am. Chem. Soc. 2012, 134, 11455–11461. [Google Scholar] [CrossRef]
  117. Kumar, S.; Yadav, R.K.; Ram, K.; Aguiar, A.; Koh, J.; Sobral, A.J.F.N. Graphene oxide modified cobalt metallated porphyrin photocatalyst for conversion of formic acid from carbon dioxide. J. CO2 Util. 2018, 27, 107–114. [Google Scholar] [CrossRef]
  118. Zhang, S.; Shi, J.; Sun, Y.; Wu, Y.; Zhang, Y.; Cai, Z.; Chen, Y.; You, C.; Han, P.; Jiang, Z. Artificial Thylakoid for the Coordinated Photoenzymatic Reduction of Carbon Dioxide. ACS Catal. 2019, 9, 3913–3925. [Google Scholar] [CrossRef]
  119. Jiang, H.-L.; Xu, Q. Porous metal–organic frameworks as platforms for functional applications. Chem. Commun. 2011, 47, 3351–3370. [Google Scholar] [CrossRef]
  120. Nadar, S.S.; Vaidya, L.; Rathod, V.K. Enzyme embedded metal organic framework (enzyme–MOF): De novo approaches for immobilization. Int. J. Biol. Macromol. 2020, 149, 861–876. [Google Scholar] [CrossRef]
  121. Zhou, J.; Yu, S.; Kang, H.; He, R.; Ning, Y.; Yu, Y.; Wang, M.; Chen, B. Construction of multi-enzyme cascade biomimetic carbon sequestration system based on photocatalytic coenzyme NADH regeneration. Renew. Energy 2020, 156, 107–116. [Google Scholar] [CrossRef]
  122. Yadav, R.K.; Oh, G.H.; Park, N.-J.; Kumar, A.; Kong, K.-J.; Baeg, J.-O. Highly Selective Solar-Driven Methanol from CO2 by a Photocatalyst/Biocatalyst Integrated System. J. Am. Chem. Soc. 2014, 136, 16728–16731. [Google Scholar] [CrossRef] [PubMed]
  123. Liu, J.; Cazelles, R.; Chen, Z.P.; Zhou, H.; Galarneau, A.; Antonietti, M. The bioinspired construction of an ordered carbon nitride array for photocatalytic mediated enzymatic reduction. Phys. Chem. Chem. Phys. 2014, 16, 14699–14705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Amao, Y.; Kataoka, R. Methanol production from CO2 with the hybrid system of biocatalyst and organo-photocatalyst. Catal. Today 2018, 307, 243–247. [Google Scholar] [CrossRef]
  125. Yadav, D.; Yadav, R.K.; Kumar, A.; Park, N.-J.; Kim, J.Y.; Baeg, J.-O. Fullerene polymer film as a highly efficient photocatalyst for selective solar fuel production from CO2. J. Appl. Polym. Sci. 2020, 137, 48536. [Google Scholar] [CrossRef]
  126. Gu, F.; Wang, Y.; Meng, Z.; Liu, W.; Qiu, L. A coupled photocatalytic/enzymatic system for sustainable conversion of CO2 to formate. Catal. Commun. 2020, 136, 105903. [Google Scholar] [CrossRef]
  127. Tian, Y.; Zhou, Y.; Zong, Y.; Li, J.; Yang, N.; Zhang, M.; Guo, Z.; Song, H. Construction of Functionally Compartmental Inorganic Photocatalyst–Enzyme System via Imitating Chloroplast for Efficient Photoreduction of CO2 to Formic Acid. ACS Appl. Mater. Interfaces 2020. [Google Scholar] [CrossRef]
  128. Ma, K.; Yehezkeli, O.; Park, E.; Cha, J.N. Enzyme Mediated Increase in Methanol Production from Photoelectrochemical Cells and CO2. ACS Catal. 2016, 6, 6982–6986. [Google Scholar] [CrossRef]
  129. King, P.W. Designing interfaces of hydrogenase–nanomaterial hybrids for efficient solar conversion. Biochim. Biophys. Acta (BBA) Bioenerg. 2013, 1827, 949–957. [Google Scholar] [CrossRef] [Green Version]
  130. Sakimoto, K.K.; Kornienko, N.; Cestellos-Blanco, S.; Lim, J.; Liu, C.; Yang, P. Physical Biology of the Materials–Microorganism Interface. J. Am. Chem. Soc. 2018, 140, 1978–1985. [Google Scholar] [CrossRef] [Green Version]
  131. Nam, D.H.; Kuk, S.K.; Choe, H.; Lee, S.; Ko, J.W.; Son, E.J.; Choi, E.-G.; Kim, Y.H.; Park, C.B. Enzymatic photosynthesis of formate from carbon dioxide coupled with highly efficient photoelectrochemical regeneration of nicotinamide cofactors. Green Chem. 2016, 18, 5989–5993. [Google Scholar] [CrossRef]
  132. Yang, J.; Wang, D.; Han, H.; Li, C. Roles of Cocatalysts in Photocatalysis and Photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. [Google Scholar] [CrossRef] [PubMed]
  133. Cooney, M.J.; Svoboda, V.; Lau, C.; Martin, G.; Minteer, S.D. Enzyme catalysed biofuel cells. Energy Environ. Sci. 2008, 1, 320–337. [Google Scholar] [CrossRef]
  134. Choi, E.-G.; Yeon, Y.J.; Min, K.; Kim, Y.H. Communication—CO2 Reduction to Formate: An Electro-Enzymatic Approach Using a Formate Dehydrogenase from Rhodobacter capsulatus. J. Electrochem. Soc. 2018, 165, H446–H448. [Google Scholar] [CrossRef]
  135. Dreyer, D.R.; Miller, D.J.; Freeman, B.D.; Paul, D.R.; Bielawski, C.W. Perspectives on poly(dopamine). Chem. Sci. 2013, 4, 3796–3802. [Google Scholar] [CrossRef]
  136. Shi, J.; Yang, C.; Zhang, S.; Wang, X.; Jiang, Z.; Zhang, W.; Song, X.; Ai, Q.; Tian, C. Polydopamine Microcapsules with Different Wall Structures Prepared by a Template-Mediated Method for Enzyme Immobilization. ACS Appl. Mater. Interfaces 2013, 5, 9991–9997. [Google Scholar] [CrossRef]
  137. Kuk, S.K.; Singh, R.K.; Nam, D.H.; Singh, R.; Lee, J.-K.; Park, C.B. Photoelectrochemical Reduction of Carbon Dioxide to Methanol through a Highly Efficient Enzyme Cascade. Angew. Chem. Int. Ed. 2017, 56, 3827–3832. [Google Scholar] [CrossRef]
  138. Zhang, L.; Vilà, N.; Kohring, G.-W.; Walcarius, A.; Etienne, M. Covalent Immobilization of (2,2′-Bipyridyl) (Pentamethylcyclopentadienyl)-Rhodium Complex on a Porous Carbon Electrode for Efficient Electrocatalytic NADH Regeneration. ACS Catal. 2017, 7, 4386–4394. [Google Scholar] [CrossRef]
  139. Srikanth, S.; Alvarez-Gallego, Y.; Vanbroekhoven, K.; Pant, D. Enzymatic Electrosynthesis of Formic Acid through Carbon Dioxide Reduction in a Bioelectrochemical System: Effect of Immobilization and Carbonic Anhydrase Addition. ChemPhysChem 2017, 18, 3174–3181. [Google Scholar] [CrossRef]
  140. Sokol, K.P.; Robinson, W.E.; Oliveira, A.R.; Warnan, J.; Nowaczyk, M.M.; Ruff, A.; Pereira, I.A.C.; Reisner, E. Photoreduction of CO2 with a Formate Dehydrogenase Driven by Photosystem II Using a Semi-artificial Z-Scheme Architecture. J. Am. Chem. Soc. 2018, 140, 16418–16422. [Google Scholar] [CrossRef] [Green Version]
  141. Amao, Y.; Fujimura, M.; Miyazaki, M.; Tadokoro, A.; Nakamura, M.; Shuto, N. A visible-light driven electrochemical biofuel cell with the function of CO2 conversion to formic acid: Coupled thylakoid from microalgae and biocatalyst immobilized electrodes. New J. Chem. 2018, 42, 9269–9280. [Google Scholar] [CrossRef]
  142. Srikanth, S.; Maesen, M.; Dominguez-Benetton, X.; Vanbroekhoven, K.; Pant, D. Enzymatic electrosynthesis of formate through CO2 sequestration/reduction in a bioelectrochemical system (BES). Bioresour. Technol. 2014, 165, 350–354. [Google Scholar] [CrossRef]
  143. Ali, I.; Gill, A.; Omanovic, S. Direct electrochemical regeneration of the enzymatic cofactor 1,4-NADH employing nano-patterned glassy carbon/Pt and glassy carbon/Ni electrodes. Chem. Eng. J. 2012, 188, 173–180. [Google Scholar] [CrossRef]
  144. Xiu, Y.; Zhang, X.; Feng, Y.; Wei, R.; Wang, S.; Xia, Y.; Cao, M.; Wang, S. Peptide-mediated porphyrin based hierarchical complexes for light-to-chemical conversion. Nanoscale 2020, 12, 15201–15208. [Google Scholar] [CrossRef]
  145. Wang, S.; Zhang, D.; Zhang, X.; Yu, D.; Jiang, X.; Wang, Z.; Cao, M.; Xia, Y.; Liu, H. Short peptide-regulated aggregation of porphyrins for photoelectric conversion. Sustain. Energy Fuels 2019, 3, 529–538. [Google Scholar] [CrossRef]
  146. Wang, S.; Li, M.; Patil, A.J.; Sun, S.; Tian, L.; Zhang, D.; Cao, M.; Mann, S. Design and construction of artificial photoresponsive protocells capable of converting day light to chemical energy. J. Mater. Chem. A 2017, 5, 24612–24616. [Google Scholar] [CrossRef] [Green Version]
  147. Rabaey, K.; Rozendal, R.A. Microbial electrosynthesis—Revisiting the electrical route for microbial production. Nat. Rev. Microbiol. 2010, 8, 706–716. [Google Scholar] [CrossRef]
  148. Amao, Y.; Shuto, N. Formate dehydrogenase–viologen-immobilized electrode for CO2 conversion, for development of an artificial photosynthesis system. Res. Chem. Intermed. 2014, 40, 3267–3276. [Google Scholar] [CrossRef]
  149. Miyatani, R.; Amao, Y. Photochemical synthesis of formic acid from CO2 with formate dehydrogenase and water-soluble zinc porphyrin. J. Mol. Catal. B Enzym. 2004, 27, 121–125. [Google Scholar] [CrossRef]
  150. Jayathilake, B.S.; Bhattacharya, S.; Vaidehi, N.; Narayanan, S.R. Efficient and Selective Electrochemically Driven Enzyme-Catalyzed Reduction of Carbon Dioxide to Formate using Formate Dehydrogenase and an Artificial Cofactor. Acc. Chem. Res. 2019, 52, 676–685. [Google Scholar] [CrossRef]
  151. Ishibashi, T.; Higashi, M.; Ikeda, S.; Amao, Y. Photoelectrochemical CO2 Reduction to Formate with the Sacrificial Reagent Free System of Semiconductor Photocatalysts and Formate Dehydrogenase. ChemCatChem 2019, 11, 6227–6235. [Google Scholar] [CrossRef]
Scheme 1. Schematic showing the strategies that are usually used in CO2 conversion.
Scheme 1. Schematic showing the strategies that are usually used in CO2 conversion.
Molecules 25 03653 sch001
Figure 1. (a) Crystal structure of UiO-66; (b) TEM images of Cu⊂UiO-66 (single Cu NC inside UiO-66); (c) Cu NC⊂UiO-66 catalyst. Atom labeling scheme: Cu, brown; C, black; O, red; Zr, blue polyhedra. H atoms are omitted for clarity; (d) turnover frequencies (TOFs) of product formation over Cu⊂UiO-66 catalyst and Cu/ZnO/Al2O3 catalyst at various reaction temperatures [33].
Figure 1. (a) Crystal structure of UiO-66; (b) TEM images of Cu⊂UiO-66 (single Cu NC inside UiO-66); (c) Cu NC⊂UiO-66 catalyst. Atom labeling scheme: Cu, brown; C, black; O, red; Zr, blue polyhedra. H atoms are omitted for clarity; (d) turnover frequencies (TOFs) of product formation over Cu⊂UiO-66 catalyst and Cu/ZnO/Al2O3 catalyst at various reaction temperatures [33].
Molecules 25 03653 g001
Figure 2. Schematic illustration of the indium–porphyrin framework for CO2 conversion with high CO selectivity [49].
Figure 2. Schematic illustration of the indium–porphyrin framework for CO2 conversion with high CO selectivity [49].
Molecules 25 03653 g002
Figure 3. Immobilization of the catalyst–carbon composite on GC (M = Co or Fe-Cl). Inset: Chemical structure of CoTPP [75].
Figure 3. Immobilization of the catalyst–carbon composite on GC (M = Co or Fe-Cl). Inset: Chemical structure of CoTPP [75].
Molecules 25 03653 g003
Figure 4. Schematic illustration of the structures of M-PMOFs (M = Co, Fe, Ni, Zn). M-TCPP: tetrakis [4-carboxyphenyl]-porphyrin-M (M-TCPP) linkers [76].
Figure 4. Schematic illustration of the structures of M-PMOFs (M = Co, Fe, Ni, Zn). M-TCPP: tetrakis [4-carboxyphenyl]-porphyrin-M (M-TCPP) linkers [76].
Molecules 25 03653 g004
Figure 5. Statistical graph of papers using enzymes as catalyst in CO2 conversion published during the past two decades.
Figure 5. Statistical graph of papers using enzymes as catalyst in CO2 conversion published during the past two decades.
Molecules 25 03653 g005
Figure 6. Biocatalytic transformation pathway of CO2 to CH3OH via stepwise reverse enzymatic catalysis by FDH, FaldDH, and ADH [106].
Figure 6. Biocatalytic transformation pathway of CO2 to CH3OH via stepwise reverse enzymatic catalysis by FDH, FaldDH, and ADH [106].
Molecules 25 03653 g006
Figure 7. Graphene-based photocatalyst catalyzed artificial photosynthesis of formic acid from CO2 under visible light [116].
Figure 7. Graphene-based photocatalyst catalyzed artificial photosynthesis of formic acid from CO2 under visible light [116].
Molecules 25 03653 g007
Figure 8. Schematic illustration of the biocatalytic PEC platform [131].
Figure 8. Schematic illustration of the biocatalytic PEC platform [131].
Molecules 25 03653 g008
Figure 9. Schematic representation of the semi-artificial photosynthetic tandem PEC cell coupling CO2 reduction to water oxidation. A blend of POs and PSII adsorbed on a dpp-sensitized photoanode (IO-TiO2|dpp|POs-PSII) is wired to an IO-TiO2|FDH cathode [140].
Figure 9. Schematic representation of the semi-artificial photosynthetic tandem PEC cell coupling CO2 reduction to water oxidation. A blend of POs and PSII adsorbed on a dpp-sensitized photoanode (IO-TiO2|dpp|POs-PSII) is wired to an IO-TiO2|FDH cathode [140].
Molecules 25 03653 g009
Table 1. Performance comparison of different photoelectrochemical CO2 reduction systems from recent literature.
Table 1. Performance comparison of different photoelectrochemical CO2 reduction systems from recent literature.
Photocathode aCondition bEfficiency cRef.
p+-n-n+-Si/TiO2 + Cu/Ag100 mW cm2, 0.1 M CsHCO3C2H4, 10–25%, −8 mA cm2 at −0.4 V vs. reversible hydrogen electrode (RHE) for 20 days[94]
p-Si NWs + Sn100 mW cm2, 0.1 M KHCO3HCOOH, 88%, 18.9 μmol h1 cm2, −0.875 V vs. RHE for 3 h[95]
CuO + Cu2O70 mW cm2, 0.1 M NaHCO3CH3OH, 95%, 85 mM at −0.2 V vs. standard hydrogen electrode (SHE) after 1.5 h[96]
Si/GaN-NPhN4-Ru(CP ) 2 2 + RuCt100 mW cm2, 0.05 M NaHCO3HCOOH, 35–64%, −1.1 mA cm2 at −0.25 V vs. RHE for 20 h[97]
p-n+-Si + SnO2 NW100 mW cm−2, 0.1 M KHCO3HCOOH, 59.2%, −18 mA cm2 at −0.4 V vs. RHE for 3 h[98]
Co3O4/CA + Ru(bpy)2dppz9 mW cm−2, 0.1 M NaHCO3HCOOH, 86%, 110 μmol h−1 cm2 at −0.60 V vs. normal hydrogen electrode (NHE) for 8 h[99]
FTO/TiO2/Cu2O + Ru-BNAH100 mW cm−2, 0.1 M KClHCOOH, NA, 409.5 umol at −0.9 V vs. NHE after 8 h[100]
p-Si + Bi50 mW cm−2, 0.5 M KHCO3HCOOH, 70–95%, ~−4 mA cm2 at −0.32 V vs. RHE for 7 h[92]
Fe2O3 NTs + Cu2O100 mW cm−2, 0.1 M KHCO3CH3OH, 93%, 6 h, 4.94 mmol L−1 cm−2 at −1.3Vvs. SCE for 6 h[101]
FTO/CuFeO2 + CuO100 mW cm−2, 0.1 M NaHCO3CH3COOH, 80%, 142 μM at −0.4 V vs. Ag/AgCl after 2 h[102]
a The configuration is described as “semiconductor + cocatalyst”. b The reaction conditions for photoelectrochemical (PEC) measurements include the light intensity of solar simulator and the electrolyte. c The PEC efficiency parameters include the product, faradaic efficiency/photocurrent density/production rate or yield/stability at a certain working potential.
Table 2. Performance comparison of different coupled photocatalytic/enzymatic CO2 reduction systems.
Table 2. Performance comparison of different coupled photocatalytic/enzymatic CO2 reduction systems.
PhotocatalystEnzymeCofactorsEfficiency aRef.
CCG-IPFateDH, FaldDH, ADHNADH + [Cp*Rh(bpy)H2O]2+Rh + TEOACH3OH, 11.21 μM after 1 h[122]
CNAFateDH, FaldDH, YADHNADH + [Cp*Rh(bpy)H2O]2+Rh + TEOACH3OH, 0.21 mM min−1[123]
H2TPPSFDH, AldDH, ADHMV2+CH3OH, 6.8 μM after 100 min[124]
C60 polymer filmFDHNADH + TEOAHCOOH, 239.46 μM after 2 h[125]
TiO2FDHNADHHCOOH, 1.634 mM after 4.5 h[126]
C3N4(TPE-C3N4)MAF-7@FDHNADH + Rh + TEOAHCOOH, 16.75 mM after 9 h[127]
CCGCMAQSPFateDH, FaldDH, ADHNADH + [Cp*Rh(bpy)H2O]2+ Rh + TEOACH3OH, 110.55 μM after 2 h[116]
CdS QDs+PTiClFDHNADH + RhHCOOH, 1500 μM h−1[118]
a The efficiency includes the products and rates.
Table 3. Performance comparison of different coupled photoelectrocatalytic/enzymatic CO2 reduction systems.
Table 3. Performance comparison of different coupled photoelectrocatalytic/enzymatic CO2 reduction systems.
PhotoanodePhotocathodeEfficiency aRef.
Co-Pi/Fe2O3ITO/FDHHCOOH, 6.4 μM h−1[131]
CoPi/BiVO4EC-PDAHCOOH, FE: 99.18%[112]
Co-Pi/αoFe2O3BiFeO3-CcFDH/PcFaldDH/YADHCH3OH, 220 μM h−1[137]
FTO/IO-TiO2/dPP/POs-PSIIFTO/IO-TiO2/FDHHCOOH, 0.185 μM cm–2[140]
FTO/FeOOH/BiVO4FTO/3D TiN-ClFDHHCOOH, 0.78 μM h−1, FE: 77.3%[113]
TK/TiO2FDH-CH3V(CH2)9COOHHCOOH, 30.0 nM after 3 h[141]
Plain graphite rodPt-FDHHCOOH, 15.49 μM mg Enzyme−1 min−1[142]
a The efficiency includes the product and rates.

Share and Cite

MDPI and ACS Style

Xu, L.; Xiu, Y.; Liu, F.; Liang, Y.; Wang, S. Research Progress in Conversion of CO2 to Valuable Fuels. Molecules 2020, 25, 3653. https://doi.org/10.3390/molecules25163653

AMA Style

Xu L, Xiu Y, Liu F, Liang Y, Wang S. Research Progress in Conversion of CO2 to Valuable Fuels. Molecules. 2020; 25(16):3653. https://doi.org/10.3390/molecules25163653

Chicago/Turabian Style

Xu, Luyi, Yang Xiu, Fangyuan Liu, Yuwei Liang, and Shengjie Wang. 2020. "Research Progress in Conversion of CO2 to Valuable Fuels" Molecules 25, no. 16: 3653. https://doi.org/10.3390/molecules25163653

Article Metrics

Back to TopTop