Next Article in Journal
Tunable Production of Diesel Bio-Blendstock by Rhenium-Catalyzed Hydrogenation of Crude Hexanoic Acid from Grape Pomace Fermentation
Next Article in Special Issue
Research Progress of Tungsten Oxide-Based Catalysts in Photocatalytic Reactions
Previous Article in Journal
Synthesis and Visible Light Catalytic Performance of BiOI/Carbon Nanofibers Heterojunction
Previous Article in Special Issue
Recent Progress of Ga-Based Catalysts for Catalytic Conversion of Light Alkanes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic CO2 Conversion to Ethanol: A Concise Review

School of Chemical and Environmental Engineering, Liaoning University of Technology, Jinzhou 121001, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2022, 12(12), 1549; https://doi.org/10.3390/catal12121549
Submission received: 28 October 2022 / Revised: 27 November 2022 / Accepted: 29 November 2022 / Published: 1 December 2022
(This article belongs to the Special Issue Photocatalysis for Energy Transformation Reactions)

Abstract

:
Photo-catalytically converting the greenhouse gas CO2 into ethanol is an important avenue for the mitigation of climate issues and the utilization of renewable energies. Catalysts play critical roles in the reaction of photocatalytic CO2 conversion to ethanol, and a number of catalysts have been investigated, including semiconductors and plasmonic metal-based catalysts, as well as several other catalysts. In this review, the progress in the development of each category of catalysts is summarized, the current status is reviewed, the remaining challenges are pointed out, and the future research directions are prospected, with the aim being to pave pathways for the rational design of better catalysts.

1. Introduction

With the proposal of the concept of “emission peaking” and “carbon neutralization”, the conversion and utilization of CO2 have been put on the agenda [1,2,3,4,5]. Photo-catalytically converting CO2 into valuable fuels is a promising approach, since it could mitigate the climate issues caused by greenhouse gas CO2 and store the renewable solar energy as chemical energy simultaneously [6,7,8,9]. The products of CO2 photocatalytic conversion reactions include CO [10,11,12], CH4 [13,14], CH3OH [15,16], C2H5OH [17,18,19,20], HCOOH [21,22], etc. Noteworthily, C2H5OH (ethanol) is a chemical with wide applications in the chemical industry, medical and healthcare industries, food industry, agriculture production, and so on. Therefore, photocatalytic CO2 conversion to ethanol has recently become a research hotspot.
Catalysts play an essential role in the reaction of photocatalytic CO2 conversion to ethanol. Up to now, a great number of photocatalysts have been developed, such as TiO2 [23], Bi2MoO6 [24], g-C3N4 [25], Cu/TiO2 [26], and AuCu/g-C3N4 [27]. Based on the nature of the developed catalysts, they could roughly be divided into semiconductors, plasmonic metal-based catalysts, and several others (Scheme 1).
Up to the present, there have been many excellent reviews on CO2 photocatalytic reactions. However, some of them focused on a special type of catalyst, such as a semiconductor [28,29] or Mexene [30], while some of them focused on the conversion of CO2 to CH4 [31] or other products [32]. To the best of our knowledge, there have been no reviews on catalysts for photocatalytic CO2 reduction to ethanol. In this paper, the progress of each category of catalysts (Scheme 1) for photocatalytic CO2 conversion to ethanol is summarized, and the current research status and the future prospect are reviewed, with the aim being to give the readers a clear picture and to inspire more studies to further advance this research area.

2. Semiconductor Based Catalysts

Semiconductors describe a category of materials which can harness solar light. Upon the irradiation of solar light with photon energy, matches or exceeds the bandgap energy of the semiconductor, and an electron jumps from the valence band (VB) to the conduction band (CB), leaving a hole. The electrons and holes can combine and dissipate the input energy as heat or transfer it to the catalyst surface. In the case that the position of CB is lower than the potential required for CO2 conversion to ethanol, the electron reacts with the adsorbed species and participates in a CO2 reduction reaction to produce ethanol (Figure 1). According to this principle, several semiconductors have been verified to be active in photocatalytic CO2 conversion to ethanol.
There are several factors affecting the efficiency of photocatalytic CO2 conversion to ethanol, as follows: (1) Light absorption region and efficiency. Solar light mainly consists of a large amount of infrared light, visible light, and a small amount of ultraviolet light. Absorbing more light means more energy can be utilized to promote the reaction. The bandgap of the catalysts is an important factor influencing the light absorption region. Meanwhile, the energy levels of the catalysts should meet the requirement of the reaction. Therefore, upon suitable energy levels, the bandgap width of the catalysts should be as small as possible to absorb more sunlight to improve the photocatalytic activity and conversion efficiency. (2) The separation and transfer of photogenerated electron–hole pairs. As catalysis is a surface reaction process, photogenerated electrons and holes must be separated and transferred to the surface to react with the adsorbates. However, photogenerated electron–hole pairs are unstable and easy to recombine during transfer. Therefore, facilitating the electron–hole pairs’ separation and transfer is effective for accelerating the reaction. (3) Photogenerated electrons and holes react with the surface adsorbates, respectively, to give products. Not all the photogenerated electrons and holes can react with the adsorbates. ① The positions of the conduction band and valence band must correspond to the positions of the corresponding reaction levels in order to have sufficient redox capacity. ② For different reactants, their ability to adsorb electrons and holes is also different. In order to ensure the smooth progress of the surface reaction, it is usually required that the reactants have sufficient adsorption on the catalyst surface, and can receive the electrons and holes on the surface smoothly.

2.1. Pristine Semiconductors

Pristine semiconductors, such as TiO2, Bi2MoO6, BiOCl, and TaON, have been reported to be active in photocatalytic CO2 conversion to ethanol. For instance, self-organized TiO2 nanotube arrays could serve as a photocatalyst for the conversion of CO2 to alcohols under xenon lamp irradiation, with the methanol and ethanol formation rates are ~10.0 nmol cm−2 h−1 and ~9.0 nmol cm−2 h−1, respectively [23]. The large specific surface area and the one-dimensional nanotubular structure of TiO2 nanotube arrays accounted for catalytic activities (Figure 2a). The mechanism is proposed as follows (Figure 2b). Upon light irradiation, electrons and holes are generated over TiO2 nanotube arrays. The holes react with adsorbed H2O to form hydroxyl radicals and hydrogen ions. The interaction between hydrogen ions and the excited electrons leads to hydrogen radicals. Meanwhile, the photoexcited electrons transfer to the conduction band and react with CO2 to produce ·CO2. However, ·CO2 is not stable and will be transformed to chemisorbed ·CO, which is subsequently reduced to ·CH2 and eventually yields methanol and ethylidene by reaction with ·OH and ·H. Ethanol is formed via the reaction of ethylidene, ·OH, and ·H (Figure 2b).
Furthermore, Bi2MoO6 is another pristine semiconductor for photocatalytic CO2 conversion to ethanol. Dai et al. reported that hierarchical flower-like Bi2MoO6 exhibited high catalytic activity for the photocatalytic reduction of CO2 under visible light irradiation, with methanol and ethanol yields of 6.2 and 4.7 μmol g−1 h−1, respectively [34]. Ribeiro et al. fabricated Bi2MoO6 catalysts by a simple hydrothermal or solvothermal method and investigated the effects of synthesis parameters on their performance in CO2 photoreduction in an aqueous medium under visible light irradiation, with the aim to pave pathways for the rational design of better catalysts in the future [35]. It was discovered that the pH value of the precursor suspensions was a key factor in determining the properties (such as zeta potential, crystallinity, and morphology) and performance of Bi2MoO6 catalysts. The more acidic the pH values, the higher ethanol production rates. The Bi2MoO6 synthesized with H2O as the solvent and pH = 2 gave the highest ethanol yield, reaching 34.4 μmol g−1 h−1 [35].
Several other Bi-based pristine semiconductors, such as BiVO4 [36], Bi2WO6 [24], and BiOCl [37], have also been successfully applied in photocatalytic CO2 reduction to ethanol. Taking BiVO4 as an example, Huang et al. reported that a large number of C1 intermediates could be generated on the surface of BiVO4 under highly intensive light irradiation, which dimerized to produce ethanol [36]. Monoclinic BiVO4 was more efficient than tetragonal BiVO4 for ethanol production, recording an ethanol production rate of 2033.0 μmol g−1 h−1 under a 300 W Xe-arc lamp irradiation, without the detection of methanol as a byproduct [36].
Additionally, TaON [38] and SrZrO3 [39] are also promising in photocatalytic CO2 reduction to ethanol. Here, SrZrO3 is taken as a representative example for elaboration. He et al. prepared SrZrO3 nanoparticles via a sonochemical method and employed it in a photocatalytic CO2 reduction reaction [39]. Ethanol, methane, and carbon monoxide were detected as the main products, with an ethanol production rate of 10.2 μmol g−1 h−1 under the irradiation of a 300 W xenon lamp. Characterization results suggested that the position of CB of SrZrO3 was 1.37 eV vs. vacuum and −3.13 eV vs. NHE, which lies above the redox potential of methane, ethanol, and carbon monoxide, indicating all of them are possible products of CO2 reduction by SrZrO3. Upon light irradiation, electron–hole pairs were generated. The electrons activated CO2 on the catalyst surface to form ·CO2− and reacted with H+ in the solution to produce ·H. The interaction between ·CO2− and ·H gave CO. The resultant CO could also be converted into ·C, followed by the formation of ·CH, ·CH2, and ·CH3 through successive reactions, which then reacted with H2O, H+ or OH to produce ethanol or methane [39].
In spite of the fact that several pristine semiconductors have been successfully applied in the reaction of photocatalytic CO2 reduction to ethanol, their efficiencies are generally low, due to their weak light absorption capacity, low photon utilization efficiency, and so on [40,41]. In this regard, several approaches have been adopted to modify the semiconductors, for example, by delicately introducing vacancy sites and constructing a heterojunction or hybrid catalyst with another semiconductor or non-semiconductor material, with the aim being to further improve their catalytic performance. In the following several sub-sections, we will review the progress of modified semiconductors in photocatalytic CO2 reduction to ethanol.

2.2. Semiconductors with Vacancy Sites

Delicately introducing vacancy sites into semiconductors is an important approach to extend the light absorption spectrum, narrow the bandgap, and regulate the electronic structure of pristine semiconductors. Semiconductors with vacancy sites have also been studied in photocatalytic CO2 reduction to ethanol.
Yang et al.’s work is a typical example [42]. They synthesized a Bi2MoO6 catalyst by assembling two-dimensional ultra-thin Bi2MoO6 nanoflakes into three-dimensional nanospherical Bi2MoO6. During the assemble process, abundant oxygen vacancies were created, resulting in two primary sites, namely the oxygen vacancies and the exposed molybdenum atoms (Figure 3). The two primary sites served as dual binding sites to trap CO2 for its activation into electronic CO* species, which were subject to accepting electrons and holes, realizing the selective reduction of CO2 into methanol and ethanol. Under visible light irradiation, the as-prepared Bi2MoO6 catalyst afforded methanol and ethanol production rates of 26.6 μmol g−1 h−1 and 2.6 μmol g−1 h−1, respectively, far surpassing those of bulk Bi2MoO6 [42].
Do et al.’s work is another example that falls into this category [43]. The authors reduced a HCa2Ta3O10 nanosheet and used it as a catalyst for photocatalytic CO2 reduction with H2O. It was discovered that the reduction process induced a considerable amount of Ta4+ and oxygen vacancies, which significantly improved the visible light harvesting capacity of HCa2Ta3O10 [43]. Introducing CuO onto reduced HCa2Ta3O10 further enhanced its performance in photocatalytic CO2 reduction to alcohols, with ethanol and methanol production rates of 113.0 μmol g−1 h−1 and 7.4 μmol g−1 h−1, respectively. The enhanced performance was ascribed to the facilitated separation of photogenerated electron–hole pairs due to the formation of p–n junctions as well as the boosted CO2 adsorption and stabilization of C1 intermediates by CuO [43].

2.3. Heterojunctions

Heterojunctions constructed by two or more semiconductors generally exhibit stronger light absorption capacity and a narrower bandgap than their corresponding single semiconductor counterparts. A number of heterojunctions have been adopted as catalysts for photocatalytic CO2 reduction to ethanol, including g-C3N4/ZnTe [44], Cu2O/g-C3N4 [45], Co3O4/CeO2 [46], MoS2/Bi2WO6 [47], TiO2/Ni(OH)2 [48], Bi/Bi2MoO6 [49], TiO2/Ti3C2 [50], CuO/TiO2 [51], and AgBr/TiO2 [52].
Here, the applications of TiO2/Ti3C2 [50] and P25 (heterojunction between anatase and rutile TiO2) [53] in photocatalytic CO2 reduction to ethanol are chosen as representatives for elaboration. The TiO2/Ti3C2, synthesized by a facile hydrothermal oxidation method, exhibited a narrowed band gap and enhanced light harvesting capacity [50]. The ratio between TiO2 and Ti3C2 affected the optical properties and performance of the heterojunctions. After the functionalization by imine ligands and Pd nanoparticles, the performance of the catalysts in CO2 activation and water splitting was further promoted. The TiO2/Ti3C2 with an optimal TiO2:Ti3C2 ratio recorded an ethanol production rate of ~10.0 μmol cm−2 h−1 at -0.6 V [50]. In case that P25 was used as a photocatalyst for CO2 conversion with H2O, multiple products, including O2, H2, C1-C4 hydrocarbons, methanol, ethanol, and acetone were detected, with an ethanol yield of 0.14 μmol g−1 h−1, under the illumination of a 100 W UV-LED in a wavelength range of 355–385 nm and a light intensity of 120 mW cm−2 [53]. The specific structure and the intensive light illumination accounted for the high ethanol yield over P25 [53].
A Z-scheme is a special category that falls into the class of heterojunctions. Catalysts with a Z-scheme structure have also been investigated in photocatalytic CO2 reduction to ethanol. For instance, Seeharaj et al. constructed TiO2/rGO/CeO2 (rGO is reduced graphene oxide) catalysts by combining surface-modified TiO2 nanoparticles with rGO and CeO2 [54]. The TiO2 surface was initially modified via the sono-assisted exfoliation method in 10 M NaOH for 1 h, which led to increased specific surface area, enhanced light absorption, and a decreased recombination rate of photoinduced electron–hole pairs. The incorporation of rGO and CeO2 further boosted the separation and transfer of photogenerated charges, electron mobility, and CO2 absorptivity. The high interfacial contact area and strong interaction between modified TiO2, rGO, and CeO2 resulted in a high photocatalytic CO2 reduction rate, with methanol and ethanol production rates of 641.0 μmol g−1 h−1 and 271.0 μmol g−1 h−1, respectively [54]. The reaction mechanism is proposed with a schematic illustration in Figure 4. The photocatalytic CO2 reduction reaction is a two-step process, involving water splitting and CO2 photoreduction. Upon light irradiation, both modified TiO2 and CeO2 were excited, forming electrons in CB and holes in VB. Then, the holes from the modified TiO2 VB transferred to CeO2 VB and subsequently oxidized H2O into OH·, H+, and O2. Meanwhile, the electrons at CeO2 CB transferred to modified TiO2 CB and then to the rGO sheet. The multiple electrons were collected and transported along the rGO sheet to reduce the adsorbed CO2 to form intermediates, such as ·CO2 and ·CO. Eventually ·CO2 and ·CO reacted with H+ to obtain methanol and ethanol [54].

2.4. Hybrid Catalysts Constructed between a Semiconductor and a Non-Semiconductor Material

Fabricating a hybrid catalyst by combining a semiconductor with a non-semiconductor material is another avenue to tailor the physicochemical and optical properties of semiconductors. Quantum dots (QD), metal organic frameworks (MOFs), conducting materials, and isolators have been adopted as modifiers to construct this type of hybrid catalyst.
(1) QD–semiconductor hybrid catalysts. The QDs are nanoparticles of semiconductors and describe a category of nanoscale crystals that can transport electrons [55,56]. In this regard, QD–semiconductor hybrid catalysts generally exhibit extraordinary properties and performance. A number of QD–semiconductor hybrid catalysts have been constructed and applied in photocatalytic CO2 reduction to ethanol, such as WS2 QD/Bi2S3 [57], and Bi2MoO6 QD/rGO [58]. Taking WS2 QD/Bi2S3 as an example, WS2 QD/Bi2S3 constructed by decorating WS2 QD onto Bi2S3 nanotubes by seed-mediated strategy was sensitive to visible/near-infrared light and displayed an excellent CO2 photoreduction activity, with methanol and ethanol production rates of 9.6 μmol g−1 h−1 and 7.0 μmol g−1 h−1, respectively [57]. Characterization results revealed that in WS2 QD/Bi2S3, the exposed S atoms in WS2 QD coordinated to Bi3+ to form a Bi–S bond, enabling the sharing of S atoms between WS2 QD and Bi2S3 (Figure 5). The junction interface between WS2 QD and Bi2S3 facilitated the separation and transfer of electron–hole pairs and consequently accounted for its enhanced catalytic performance [57]. Cheng et al.’s study is another example [59]. They prepared a CdS-Cu2+/TiO2 nanorod array film photocatalyst, in which a TiO2 nanorod array was synthesized by a hydrothermal method, and CdS and Cu2+ were deposited on TiO2 by a cation adsorption method and successive ion layer adsorption reaction (SILAR). Its performance in photocatalytic reduction of CO2 under visible light irradiation was measured under visible-near infrared light. The results showed that the yield of ethanol reached the maximum value (109.1 μmol g-cat−1 h−1) when SILAR was deposited twice, at a flow rate of 4 mL min−1 and a reaction temperature of 80 °C. The high catalytic activity of CdS-Cu2+/TiO2 was attributed to the combination of one-dimensional nanostructure with Cu2+ ions and CdS quantum dots, which restrained the recombination of the electron–hole pairs and broadened the visible light responsive region [59].
(2) The MOF–semiconductor hybrid catalysts. The MOFs are a class of porous polymeric materials, in which metal ions are linked together by organic bridging ligands. These MOFs usually have the advantages of highly porous structure, large specific surface area, and adjustable pore size, which endow them special properties as modifiers or catalysts [60,61,62]. For instance, Liu et al. encapsulated CuO QDs in the pores of MIL-125(Ti) (MIL-125(Ti) is a type of MOF) and further combined it with g-C3N4 to fabricate a g-C3N4/CuO@MIL-125(Ti) catalyst, which exhibited a high catalytic activity for photocatalytic CO2 reduction in the presence of H2O, with yields of CO, methanol, acetaldehyde, and ethanol up to 60.0 μmol g−1 h−1, 332.4 μmol g−1 h−1, 177.2 μmol g−1 h−1, and 501.9 μmol g−1 h−1, respectively [63]. A mechanism study revealed that, under light irradiation, electrons and holes were generated and separated (Figure 6). Due to the positions of the energy levels of g-C3N4, CuO QDs, and MIL-125(Ti), the electrons remained at CB of CuO QDs, and the holes remained at VB of g-C3N4. The potential energy of electrons on CB of CuO QDs met the requirements for CO2 reduction to CO, methanol, acetaldehyde, and ethanol, and led to the generation of these products. The valence band of g-C3N4 was more positive than the oxidation potential of H2O, resulting in the oxidation of H2O to O2 [63]. Cardoso et al. prepared a hybrid catalyst via growing MOF-based nanoparticles (ZIF-8) on Ti/TiO2 nanotubes and adopted the as-prepared Ti/TiO2-ZIF-8 catalyst in the photocatalytic CO2 reduction reaction. The Ti/TiO2-ZIF-8 can produce ethanol up to 10.0 mmol L−1. The increased photocurrent (ZIF-8 acted as a cocatalyst to interact with Ti/TiO2 nanotubes) and promoted electron transfer accelerated CO2 photocatalytic reduction to ethanol [64].
(3) Conducting material–semiconductor hybrid catalysts. Integrating a semiconductor with a conducting material is an avenue to facilitate the electron transfer and prohibit the recombination of photoexcited electron–hole pairs [65,66]. For instance, modifying semiconductor Bi2WO6 with conducting polymers tailored the photoelectronic properties (band gap, charge mobility, etc.) and promoted the photocatalytic performance in photocatalytic CO2 reduction [65]. Under visible light irradiation, the as-fabricated catalyst demonstrated methanol and ethanol yields of 14.1 μmol g−1 h−1 and 5.1 μmol g−1 h−1, respectively [65]. Similarly, graphitic-supported multiple functionalized TiO2 nanowire (denoted as R-TiO2@Gs) recorded an ethanol yield of 124.2 μM in CO2 reduction with water after light irradiation for 6 h. The graphitic support accelerated the electron transfer, while the ligands in functionalized TiO2 enabled the catalyst to capture CO2 more efficiently and facilitated C–C coupling to produce ethanol [67].
(4) Isolator–semiconductor hybrid catalysts. Loading a semiconductor onto an isolator with a large specific surface area could increase the number of active sites and enhance the photocatalytic activity. Du and co-author’s work is representative of this [68]. They constructed a TPS/g-C3N4 (TPS is trimodal porous silica) composite catalyst via a two-step hydrothermal synthesis method. The TPS/g-C3N4 catalysts were of hollow tubular shapes, with a large specific surface area, high CO2 adsorption capacity, and more active sites. Consequently, TPS/g-C3N4 exhibited a high activity in photocatalytic CO2 reduction reaction to ethanol, with an ethanol yield of 196.0 μmol g−1 h−1 and an ethanol selectivity of ~100% [68].

2.5. Doped Semiconductors

Doped semiconductors generally exhibit engineered energy levels and bandgaps, which improve the light absorption and facilitate the separation and transfer of electron–hole pairs.
For example, Maimaitizi et al. prepared hollow-graded BiOCl microspheres co-doped with N and Pt by an in situ hydrothermal method and explored its performance in CO2 photoreduction to ethanol [69]. Under visible light irradiation, the ethanol yield reached 14.15 μmol gcat−1 h−1. Results suggested that the scattering effect and surface reflection caused by the special layered structure of the catalyst, the narrowing of the bandgap caused by N doping, and the Schottky barrier caused by the existence of Pt accelerated the charge separation and transfer, and consequently accounted for the high catalytic performance [69]. Li et al. successfully synthesized a Zn-doped g-C3N4 catalyst by a one-step calcination method and investigated the effects of operational conditions on its performance in CO2 photoreduction under ultraviolet or visible light irradiation [70]. Notably, the optimized 0.5%Ru/Zn-g-C3N4-1/20 catalyst gave the best photocatalytic activity, with the yield of ethanol reaching 1442.9 μmol g−1. A mechanism study revealed that electrons were transferred to Ru through Zn–N bonds and reacted with adsorbed CO2 during light irradiation. At the same time, CH4 combined with holes to form methyl, which can be attracted by Ru and connects with *CHO to form acetaldehyde intermediate. When some of the intermediates were converted to acetaldehyde, most of them were further hydrogenated to form ethanol [70].

3. Plasmonic Metal-Based Catalysts

Plasmonic metals, such as Cu, Ag, Au, and their alloys, are sensitive to visible light and could act as active sites for photocatalytic reactions [71,72]. Plasmonic metal-based catalysts have also been widely applied in photocatalytic CO2 reduction to ethanol [73,74]. Generally speaking, plasmonic metal-based catalysts give higher activities than semiconductors for CO2 photoreduction. In this section, the progress of plasmonic metals-based catalysts for the photocatalytic conversion of CO2 to ethanol is reviewed.

3.1. Cu-Based Catalysts

(1) Cu nanoparticle-based catalysts. These Cu nanoparticles are of plasmonic properties and have been studied in photocatalytic CO2 reduction to ethanol. For example, Xuan et al. took the advantages of the plasmonic effect of Cu and the chemical absorption capacity of CO2 by Cu@Ni to fabricate a SrTiO3/Cu@Ni/TiN catalyst [75]. The as-prepared SrTiO3/Cu@Ni/TiN could capture full-spectrum solar energy and activate CO2 efficiently, and consequently exhibited an ethanol evolution rate of 21.3 μmol g−1 h−1 and an ethanol selectivity of 79% under the irradiation of a 600 mW cm−2 Xe lamp [75]. Density functional theory calculation suggested that CO2 activation was the rate-determining step and that CO2* was easier to absorb on the interface of Cu (100) and Ni (111) (Figure 7). In addition, CO* was difficult to desorb at the interface of Cu (100) and Ni (111), which facilitated the dimerization of CO to produce ethanol (Figure 7) [75]. Similarly, Cu-TiO2/GO (GO = graphene oxide) synthesized via a one-step hydrothermal method was effective for photocatalytic CO2 reduction to ethanol, with an ethanol production rate of 233 μmol g−1 h−1 [76]. The high specific surface area, the narrowed band gap, and the plasmonic properties of Cu accounted for its performance [76].
(2) Cu ion-based catalysts. Here, CuI could selectively catalyze CO2 conversion to ethanol; however, the catalytic sites of CuI are not stable. Incorporating CuI into the cavities of MOFs or decorating Cu single atoms onto MOFs could retain the chemical state of CuI [77,78]. In this regard, several light responsive Cu–MOFs catalysts have been designed for photocatalytic CO2 reduction to ethanol. For instance, Lin et al. used low intensity light to activate an in situ CuII(HxPO4)y@Ru-Uio catalyst to generate CuI species in the cavities of Uio-67 [77]. Upon light irradiation, one single electron transferred from photoexcited [Ru(bpy)3]2+-based ligands on Uio-67 to CuII centers in the cavities and one single hole transferred from Cu0 to [Ru(bpy)3]2+-based ligands for the generation of CuI (Figure 8). The CuI then served as the active centers for photocatalytic CO2 reduction to ethanol, with an activity of 9650.0 μmol gCu−1 h−1 at 150 °C [77].
The Cu2+ incorporated into semiconductors can also serve as a catalyst to drive the reaction of photocatalytic CO2 reduction to ethanol, such as Cu doped into TiO2 [26,79]. The preparation method, as well as the morphology of TiO2, strongly affected the properties and performance of the as-prepared Cu-TiO2 catalysts. The Cu-doped TiO2 nanorod, which was synthesized via the combination of the hydrothermal method and ultrasonic assisted sequential adsorption method, exhibited improved photon transfer due to the one-dimensional nanostructure of TiO2 and the incorporation of Cu2+, and resulted in methanol and ethanol yields of 36.2 μmol g−1 h−1 and 79.1 μmol g−1 h−1 at 80 °C and UV light irradiation [26]. The Cu-TiO2 nanocatalyst fabricated by the sol-gel method possessed a large specific surface area, increased number of oxygen vacancies, and enhanced atomic mobility, which improved CO2 photoreduction by H2O, with methane, hydrogen, methanol, ethanol, and acetaldehyde as products [79].
(3) CuO semiconductor-based catalyst. In recent years, CuO has attracted extensive attention in the field of photocatalytic CO2 reduction due to its strong absorption capacity towards solar energy. In addition, the combination of CuO with other semiconductors could reduce the rapid recombination of photogenerated electron–hole pairs and produce ethanol under light irradiation. For example, Lu et al. prepared a Re-doped CuO/TiO2-NTs catalyst by doping rhenium into CuO/TiO2 nanotube arrays, which gave methanol and ethanol as the main products in photocatalytic CO2 reduction reaction. With the increase in Re, the proportion of ethanol in the product increased (Figure 9), with the optimized yield of ethanol reaching 7.5 µmol over 6wt% re-doped CuO/TiO2-NTs after applying an external voltage of 0.4 V under simulated solar light illumination. The remarkable result might have originated from the tuned interface characteristics of re-doped CuO/TiO2-NTs, which promoted the selectivity towards alcohols and accelerated the occurrence of the C–C coupling reaction [80].

3.2. Ag-Based Catalysts

Another plasmonic metal, Ag, has been utilized in photocatalytic CO2 reduction to ethanol. For example, Shu et al. synthesized an Ag@AgBr/carbon nanotubes (CNT) nanocomposite catalyst by anchoring Ag@AgBr nanoparticles onto the surface of CNT, and investigated the effects of CNT length on the performance of Ag@AgBr/CNT in photocatalytic CO2 reduction reaction under visible light irradiation [81]. It was discovered that CNT with longer length facilitated the separation of electron–hole pairs. Together with the plasmonic properties of Ag and the unique structure of Ag@AgBr/CNT nanocomposite, Ag@AgBr/CNT with longer CNT length exhibited a promoted activity in CO2 reduction to methane, CO, methanol, and ethanol, with an ethanol yield of ~5.0 μmol g−1 h−1 [81]. This was not only limited to Ag@AgBr/CNT, as Ag@AgBr/AgCl also showed activity for CO2 conversion to methanol and ethanol under visible light irradiation [82].

3.3. Au-Based Catalysts

The Au-based catalysts are another type of plasmonic metal for CO2 reduction to ethanol. Do et al. deposited plasmonic Au nanoparticles onto ZIF-67 (ZIF-67 is a type of MOF) and investigated the performance of Au/ZIF-67 in a photocatalytic CO2 reduction [83]. It was found that the loading of Au affected the size of Au nanoparticles, and Au nanoparticles with sizes in the range of 30–40 nm exhibited improved light harvesting capacity, enhanced charge separation, and played crucial roles in determining selectivity. Volcano relationships were obtained between the production rates of methanol/ethanol and the loading of Au, with an optimal ethanol production rate of 0.5 mmol g−1 h−1 (Figure 10a,b) [83]. The mechanism was proposed as follows: under light irradiation, plasmonic Au were excited and generated energetic electrons. These electrons overcame the Schottky barrier and injected into ZIF-67, which then participated in the activation and conversion of CO2 to methanol and ethanol, which had already been adsorbed on the surface of ZIF-67 (Figure 10c) [83]. Ramis et al. probed the key intermediates and products over Au/TiO2 in a photocatalytic CO2 reduction reaction [84]. They revealed that several different CO2 adsorption modes (i.e., CO2, bicarbonate, and carbonate) could be observed depending on the loading of Au. The presence of H2O promoted the formation of CO2 radicals. Methanol mainly adsorbed over TiO2 sites, forming methoxy-species, which could be converted into ethanol [84]. The probation of the intermediates and products provided insights for the mechanism study.

3.4. Plasmonic Alloy-Based Catalysts

Plasmonic alloys exhibit not only the plasmonic properties but also some specific properties [11], which empower their applicability in photocatalytic CO2 reduction to ethanol. For instance, AuCu/g-C3N4 was a very promising catalyst, affording an ethanol yield and selectivity of 0.9 mmol g−1 h−1 and 93.1%, respectively [27]. In addition to the plasmonic properties, the alloy structure of AuCu, as well as the interactions between AuCu and g-C3N4, contributed to its photocatalytic performance. Over AuCu alloy, Au was positively charged, and Cu was negatively charged due to their electronegativity difference. The positive charge on Au promoted CO2 adsorption and the negative charge on Cu facilitated the formation of the intermediates CO2· and *CO. The interaction between AuCu and g-C3N4 facilitated the transfer of photogenerated charges [27]. Similarly, Pd2Cu/TiO2 catalyst gave an ethanol production rate of 4.1 mmol g−1 h−1 at 150 °C under visible light irradiation [85]. The plasmonic properties of Pd2Cu, the CO2 adsorption capacity of Cu, and the oxidation and C-C bond formation competence of Pd accounted for its performance [85].

4. Other Catalysts

In addition to semiconductor- and plasmonic metal-based catalysts, Co-based catalysts and Pd-based catalysts have also been studied in photocatalytic CO2 reduction to ethanol. In this section, we will review the progress of these two types of photocatalysts for CO2 reduction reactions.

4.1. Co-Based Catalysts

In thermal-driven reaction systems, Co is one of the active metals for C-C coupling reactions [86,87]. The competence of Co active sites for the formation of C-C bond endows Co-based catalysts applicability in the photocatalytic CO2 reduction to ethanol [88,89]. For instance, Na-modified Co@C nanocomposite catalyst gave almost 100% selectivity to hydrocarbons and ~6% selectivity towards ethanol at 235 °C and under the irradiation of a solar simulator. Mechanism study revealed that, upon light irradiation, photoexcited charges were generated on Na-Co@C, which facilitated the formation of electron-rich carbon species. These species were further involved in CO2 activation to CO2δ− and promoted the dissociation of CO2 to CO. The CO was stabilized by the carbon layers on Na-Co@C, and produced ethanol via a CO insertion pathway [88].

4.2. Pd-Based Catalysts

The PdIn@N3-COF (N3-COF is a photosensitizing covalent organic framework) [90] and Pd/Mn-TiO2 [91] catalysts have also been successfully utilized in photocatalytic CO2 reduction to ethanol. Here, PdIn@N3-COF is taken as an example for elaboration. Lu et al. confined bimetallic PdIn nanoclusters in N3-COF to construct a PdIn@N3-COF composite, and investigated its performance in photocatalytic CO2 reduction to ethanol [90]. It revealed that PdIn@N3-COF gave a total yield toward alcohols of 33.3 μmol g−1 h−1 and a selectivity to ethanol of 26%. On the one hand, the interaction between PdIn and N3-COF facilitated the charge transfer; on the other hand, the bimetallic synergistic effect of PdIn-stabilized C1 intimidates C-C coupling while retaining some of the C-O bonds. Both of these two factors contributed to the high conversion of CO2 to ethanol [90].

5. Summary and Outlook

Up to now, a number of catalysts have been designed for photocatalytic CO2 reduction to ethanol, including semiconductors, plasmonic-metal based catalysts, and several other catalysts (a brief summary of some typical catalysts is shown in Table 1). Clearly, in spite of the rapid progress, several challenges remain.
(1) CO2 conversion rates over most of the catalysts are still low.
At present, semiconductors in photocatalytic CO2 conversion to ethanol mainly face the following challenges: ➀ most semiconductors have a relatively low response to light due to the limitation of their own electronic structures, and ➁ photogenerated electron–hole pairs in semiconductors combine relatively quickly. These challenges result in ineffective performance of semiconductors in CO2 photoreduction to ethanol. In terms of plasmonic metal-based catalysts, the absorption of light by plasmonic metal nanoparticles is mainly concentrated in the range of the ultraviolet light and visible light region, which makes the light utilization efficiency very low and leads to poor performance.
Due to these reasons, over most of the studied catalysts, CO2 conversion rates are in the magnitude of μmol g−1 h−1. Even though some catalysts could record CO2 conversion rates up to mmol g−1 h−1, this is still far away from what is required for industrialization applications. Therefore, there is still a long way to accelerate CO2 conversion rates. Developing catalysts with high efficiencies for CO2 activation or establishing photothermal catalytic systems to enhance CO2 conversion might be future research pathways. For instance, it has been reported that surface site engineering of semiconductors is beneficial to increase the absorption range of light and to enhance the separation of photogenerated electrons and holes, thereby promoting the surface redox reaction and improving the photocatalytic CO2 reduction to methanol, methane, CO, and others [92,93]. Adopting the surface site engineering strategy to develop suitable catalysts for CO2 photoreduction to ethanol might be promising approach to enhance the CO2 conversion rates.
(2) The selectivity towards ethanol needs improvement.
Photocatalytic CO2 reduction to ethanol requires multiple electrons of strong energies [94,95]. The requirements are more critical than CO2 photoreduction to CO, methane, methanol, and so on, which makes it difficult to realize 100% selectivity towards ethanol. Therefore, developing catalysts with tailored properties which could selectively produce ethanol is one of the challenges in this study. It has been reported that single atom catalysts are of specific geometric and electronic structures, which limits the absorption geometry of reactants on the catalytic active sites and is beneficial to providing product selectively. Therefore, delicately designing single atom catalysts or catalysts with specified sizes might be avenue to improve the selectivity towards ethanol.
(3) Some plasmonic metal-based catalysts are expensive.
Currently, the plasmonic metal-based catalysts used in photocatalytic CO2 conversion to ethanol are mainly focused on precious metals, such as gold and silver. These noble metals are expensive and deactivate easily due to sintering. Therefore, non-noble metal-based plasmonic catalysts should be developed, such as Cu, Al, and some transition oxides with plasmonic properties. With the help of the above two prospects to enhance the CO2 conversion rate and improve the selectivity towards ethanol, if these non-noble metal-based plasmonic catalysts are of high catalytic performance in CO2 photoreduction to ethanol, it will be of stronger practical significance.
(4) The reaction mechanism is still unclear.
Most of the current studies focus on catalyst design and the improvement of ethanol production, whereas the underlying reaction mechanisms are not extensively investigated. Making clear the reaction mechanism could provide guidance for the rational design of efficient catalysts in the future. Combining the advanced in situ techniques (such as high-angle annular dark field scanning transmission electron microscopy, extended X-ray absorption fine structure, X-ray absorption near-edge structure, diffuse reflectance infrared Fourier transform spectroscopy, atmospheric pressure X-ray photoelectron spectroscopy), and theoretical calculations (density functional theory), might be avenues to unravel the underlying mechanisms.

Author Contributions

Writing—original draft preparation, D.L.; writing—review and editing, C.H.; guidance, supervision and project administration, H.L.; visualization, Conceptualization, R.Z. and Y.L.; software, J.G. (Jiawen Guo); formal analysis, J.G. (Jiapeng Guo); resources, C.C.V.; All authors have read and agreed to the published version of the manuscript.

Funding

National Natural Science Foundation of China (21902116) and the Education Department of Liaoning Province (JQL202015401).

Data Availability Statement

We can provide the data upon requirements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Das, S.; Pérez-Ramírez, J.; Gong, J.; Dewangan, N.; Hidajat, K.; Gates, B.C.; Kawi, S. Core-shell structured catalysts for thermocatalytic, photocatalytic, and electrocatalytic conversion of CO2. Chem. Soc. Rev. 2020, 49, 2937–3004. [Google Scholar] [CrossRef] [PubMed]
  2. Singh, G.; Lee, J.; Karakoti, A.; Bahadur, R.; Yi, J.; Zhao, D.; AlBahily, K.; Vinu, A. Emerging trends in porous materials for CO2 capture and conversion. Chem. Soc. Rev. 2020, 49, 4360–4404. [Google Scholar] [CrossRef]
  3. Masel, R.I.; Liu, Z.; Yang, H.; Kaczur, J.; Carrillo, D.; Ren, S.; Salvatore, D.; Berlinguette, C. An industrial perspective on catalysts for low-temperature CO2 electrolysis. Nat. Nanotechnol. 2021, 16, 118–128. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, D.; Bi, Q.; Yin, G.; Wang, P.; Huang, F.; Xie, X.; Jiang, M. Photochemical preparation of anatase titania supported gold catalyst for ethanol synthesis from CO2 hydrogenation. Catal. Lett. 2018, 148, 11–22. [Google Scholar] [CrossRef]
  5. Wang, J.; Han, B.; Nie, R.; Xu, Y.; Yu, X.; Dong, Y.; Wang, J.; Jing, H. Photoelectrocatalytic reduction of CO2 to chemicals via ZnO@ nickel foam: Controlling C-C coupling by ligand or morphology. Top. Catal. 2018, 61, 1563–1573. [Google Scholar] [CrossRef]
  6. Wang, Z.J.; Song, H.; Liu, H.; Ye, J. Coupling of solar energy and thermal energy for carbon dioxide reduction: Status and prospects. Angew. Chem. Int. Ed. 2020, 59, 8016–8035. [Google Scholar] [CrossRef]
  7. Zhang, W.; Mohamed, A.R.; Ong, W.J. Z-Scheme photocatalytic systems for carbon dioxide reduction: Where are we now? Angew. Chem. Int. Ed. 2020, 59, 22894–22915. [Google Scholar] [CrossRef] [PubMed]
  8. Dong, W.-H.; Wu, D.-D.; Luo, J.-M.; Xing, Q.-J.; Liu, H.; Zou, J.-P.; Luo, X.-B.; Min, X.-B.; Liu, H.-L.; Luo, S.-L. Coupling of photodegradation of RhB with photoreduction of CO2 over rGO/SrTi0.95Fe0.005O3−delta catalyst: A strategy for one-pot conversion of organic pollutants to methanol and ethanol. J. Catal. 2017, 349, 218–225. [Google Scholar] [CrossRef]
  9. Jeyalakshmi, V.; Tamilmani, S.; Mahalakshmy, R.; Bhyrappa, P.; Krishnamurthy, K.R.; Viswanathan, B. Sensitization of La modified NaTaO3 with cobalt tetra phenyl porphyrin for photo catalytic reduction of CO2 by water with UV–visible light. J. Mol. Catal. A-Chem. 2016, 420, 200–207. [Google Scholar] [CrossRef]
  10. Liu, H.; Meng, X.; Dao, T.D.; Zhang, H.; Li, P.; Chang, K.; Wang, T.; Li, M.; Nagao, T.; Ye, J. Conversion of carbon dioxide by methane reforming under visible-light irradiation: Surface-plasmon-mediated nonpolar molecule activation. Angew. Chem. Int. Ed. 2015, 127, 11707–11711. [Google Scholar] [CrossRef]
  11. Liu, H.; Dao, T.D.; Liu, L.; Meng, X.; Nagao, T.; Ye, J. Light assisted CO2 reduction with methane over group VIII metals: Universality of metal localized surface plasmon resonance in reactant activation. Appl. Catal. B Environ. 2017, 209, 183–189. [Google Scholar] [CrossRef]
  12. Liu, H.; Li, M.; Dao, T.D.; Liu, Y.; Zhou, W.; Liu, L.; Meng, X.; Nagao, T.; Ye, J. Design of PdAu alloy plasmonic nanoparticles for improved catalytic performance in CO2 reduction with visible light irradiation. Nano Energy 2016, 26, 398–404. [Google Scholar] [CrossRef]
  13. Zhao, Y.; Wei, Y.; Wu, X.; Zheng, H.; Zhao, Z.; Liu, J.; Li, J. Graphene-wrapped Pt/TiO2 photocatalysts. with enhanced photogenerated charges separation and reactant adsorption for high selective photoreduction of CO2 to CH4. Appl. Catal. B Environ. 2018, 226, 360–372. [Google Scholar] [CrossRef]
  14. Li, X.; Sun, Y.; Xu, J.; Shao, Y.; Wu, J.; Xu, X.; Pan, Y.; Ju, H.; Zhu, J.; Xie, Y. Selective visible-light-driven photocatalytic CO2 reduction to CH4 mediated by atomically thin CuIn5S8 layers. Nat. Energy 2019, 4, 690–699. [Google Scholar] [CrossRef]
  15. Wu, Y.A.; McNulty, I.; Liu, C.; Lau, K.C.; Liu, Q.; Paulikas, A.P.; Sun, C.-J.; Cai, Z.; Guest, J.R.; Ren, Y. Facet-dependent active sites of a single Cu2O particle photocatalyst for CO2 reduction to methanol. Nat. Energy 2019, 4, 957–968. [Google Scholar] [CrossRef]
  16. Chen, G.; Waterhouse, G.I.; Shi, R.; Zhao, J.; Li, Z.; Wu, L.Z.; Tung, C.H.; Zhang, T. From solar energy to fuels: Recent advances in light-driven C1 chemistry. Angew. Chem. Int. Ed. 2019, 58, 17528–17551. [Google Scholar] [CrossRef]
  17. Singh, M.R.; Bell, A.T. Design of an artificial photosynthetic system for production of alcohols in high concentration from CO2. Energy Environ. Sci. 2016, 9, 193–199. [Google Scholar] [CrossRef] [Green Version]
  18. Sans, J.; Sanz, V.; Turon, P.; Alemán, C. Enhanced CO2 conversion into ethanol by permanently polarized hydroxyapatite through C-C Coupling. ChemCatChem 2021, 13, 5025–5033. [Google Scholar] [CrossRef]
  19. Usubharatana, P.; McMartin, D.; Veawab, A.; Tontiwachwuthikul, P. Photocatalytic process for CO2 emission reduction from industrial flue gas streams. Ind. Eng. Chem. Res. 2006, 45, 2558–2568. [Google Scholar] [CrossRef]
  20. Jeyalakshmi, V.; Mahalakshmy, R.; Krishnamurthy, K.R.; Viswanathan, B. Photocatalytic reduction of carbon dioxide in alkaline medium on La modified sodium tantalate with different CO-Catalysts. under UV–Visible radiation. Catal. Today 2016, 266, 160–167. [Google Scholar] [CrossRef]
  21. Li, D.; Kassymova, M.; Cai, X.; Zang, S.-Q.; Jiang, H.-L. Photocatalytic CO2 reduction over metal-organic framework-based materials. Coord. Chem. Rev. 2020, 412, 213262. [Google Scholar] [CrossRef]
  22. Kumaravel, V.; Bartlett, J.; Pillai, S.C. Photoelectrochemical conversion of carbon dioxide (CO2) into fuels and value-added products. ACS Energy Lett. 2020, 5, 486–519. [Google Scholar] [CrossRef] [Green Version]
  23. Ping, G.; Wang, C.; Chen, D.; Liu, S.; Huang, X.; Qin, L.; Huang, Y.; Shu, K. Fabrication of self-organized TiO2 nanotube arrays for photocatalytic reduction of CO2. J. Solid State Electr. 2013, 17, 2503–2510. [Google Scholar] [CrossRef]
  24. Ribeiro, C.S.; Lansarin, M.A. Enhanced photocatalytic activity of Bi2WO6 with PVP addition for CO2 reduction into ethanol under visible light. Environ. Sci. Pollut. R 2021, 28, 23667–23674. [Google Scholar] [CrossRef]
  25. Li, F.; Zhang, D.; Xiang, Q. Nanosheet-assembled hierarchical flower-like g-C3N4 for enhanced photocatalytic CO2 reduction activity. Chem. Commun. 2020, 56, 2443–2446. [Google Scholar] [CrossRef] [PubMed]
  26. Cheng, M.; Yang, S.; Chen, R.; Zhu, X.; Liao, Q.; Huang, Y. Copper-decorated TiO2 nanorod thin films in optofluidic planar reactors for efficient photocatalytic reduction of CO2. Int. J. Hydrogen Energy 2017, 42, 9722–9732. [Google Scholar] [CrossRef]
  27. Li, P.; Liu, L.; An, W.; Wang, H.; Guo, H.; Liang, Y.; Cui, W. Ultrathin porous g-C3N4 nanosheets modified with AuCu alloy nanoparticles and C-C coupling photothermal catalytic reduction of CO2 to ethanol. Appl. Catal. B Environ. 2020, 266, 118618. [Google Scholar] [CrossRef]
  28. Habisreutinger, S.N.; Schmidt Mende, L.; Stolarczyk, J.K. Photocatalytic reduction of CO2 on TiO2 and other semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef]
  29. 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]
  30. Amrillah, T.; Supandi, A.R.; Puspasari, V.; Hermawan, A.; Seh, Z.W. MXene-based photocatalysts and electrocatalysts for CO2 conversion to chemicals. Trans. Tianjin Univ. 2022, 4, 307–322. [Google Scholar] [CrossRef]
  31. Cheng, S.; Sun, Z.; Lim, K.H.; Gani, T.Z.H.; Zhang, T.; Wang, Y.; Yin, H.; Liu, K.; Guo, H.; Du, T. Emerging strategies for CO2 photoreduction to CH4: From experimental to data-driven design. Adv. Energy Mater. 2022, 12, 2200389. [Google Scholar] [CrossRef]
  32. Chang, X.; Wang, T.; Gong, J. CO2 photo-reduction: Insights into CO2 activation and reaction on surfaces of photocatalysts. Energy Environ. Sci. 2016, 9, 2177–2196. [Google Scholar] [CrossRef]
  33. Song, Y.; Chen, W.; Wei, W.; Sun, Y. Advances in clean fuel ethanol production from electro-, photo- and photoelectro-catalytic CO2 Reduction. Catalysts 2020, 10, 1287. [Google Scholar] [CrossRef]
  34. Dai, W.; Yu, J.; Xu, H.; Hu, X.; Luo, X.; Yang, L.; Tu, X. Synthesis of hierarchical flower-like Bi2MoO6 microspheres as efficient photocatalyst for photoreduction of CO2 into solar fuels under visible light. Crystengcomm 2016, 18, 3472–3480. [Google Scholar] [CrossRef]
  35. Ribeiro, C.S.; Lansarin, M.A. Facile solvo-hydrothermal synthesis of Bi2MoO6 for the photocatalytic reduction of CO2 into ethanol in water under visible light. React. Kinet. Mech. Cat. 2019, 127, 1059–1071. [Google Scholar] [CrossRef]
  36. Liu, Y.; Huang, B.; Dai, Y.; Zhang, X.; Qin, X.; Jiang, M.; Whangbo, M.-H. Selective ethanol formation from photocatalytic reduction of carbon dioxide in water with BiVO4 photocatalyst. Catal. Commun. 2009, 11, 210–213. [Google Scholar] [CrossRef]
  37. Sanchez-Rodriguez, D.; Berenice Jasso-Salcedo, A.; Hedin, N.; Church, T.L.; Aizpuru, A.; Alonso Escobar-Barrios, V. Semiconducting nanocrystalline Bismuth Oxychloride (BiOCl) for Photocatalytic Reduction of CO2. Catalysts 2020, 10, 998. [Google Scholar] [CrossRef]
  38. Han, Q.; Zhou, Y.; Tang, L.; Li, P.; Tu, W.; Li, L.; Lia, H.; Zou, Z. Synthesis of single-crystalline, porous TaON microspheres toward visible-light photocatalytic conversion of CO2 into liquid hydrocarbon fuels. RSC Adv. 2016, 6, 90792–90796. [Google Scholar] [CrossRef]
  39. Ashiq, M.N.; Wang, Y.; Ehsan, M.F.; He, T. Photoreduction of carbon dioxide using strontium zirconate nanoparticles. Sci. China Mater. 2015, 58, 634–639. [Google Scholar]
  40. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. Dual cocatalysts in TiO2 photocatalysis. Adv. Mater. 2019, 31, 1807660. [Google Scholar] [CrossRef]
  41. Shayegan, Z.; Lee, C.-S.; Haghighat, F. TiO2 photocatalyst for removal of volatile organic compounds in gas phase-A review. Chem. Eng. J. 2018, 334, 2408–2439. [Google Scholar] [CrossRef]
  42. Dai, W.; Long, J.; Yang, L.; Zhang, S.; Xu, Y.; Luo, X.; Zou, J.; Luo, S. Oxygen migration triggering molybdenum exposure in oxygen vacancy-rich ultra-thin Bi2MoO6 nanoflakes: Dual binding sites governing selective CO2 reduction into liquid hydrocarbons. J. Energy Chem. 2021, 61, 281–289. [Google Scholar] [CrossRef]
  43. Nhu-Nang, V.; Chinh-Chien, N.; Kaliaguine, S.; Trong-On, D. Reduced Cu/Pt-HCa2Ta3O10 perovskite nanosheets for sunlight-driven conversion of CO2 into valuable fuels. Adv. Sustain. Syst. 2017, 1, 1700048. [Google Scholar]
  44. Wang, Q.; Wang, X.; Yu, Z.; Jiang, X.; Chen, J.; Tao, L.; Wang, M.; Shen, Y. Artificial photosynthesis of ethanol using type-II g-C3N4/ZnTe heterojunction in photoelectrochemical CO2 reduction system. Nano Energy 2019, 60, 827–835. [Google Scholar] [CrossRef]
  45. Li, P.; Liu, L.; An, W.; Wang, H.; Cui, W. Efficient photothermal catalytic CO2 reduction to CH3CH2OH over Cu2O/g-C3N4 assisted by ionic liquids. Appl. Surf. Sci. 2021, 565, 150448. [Google Scholar] [CrossRef]
  46. Huang, Y.; Yan, C.-F.; Guo, C.-Q.; Huang, S.-L. Enhanced photoreduction activity of carbon dioxide over Co3O4/CeO2 catalysts under visible light irradiation. Int. J. Photoenergy 2015, 2015, 230808. [Google Scholar] [CrossRef] [Green Version]
  47. Dai, W.; Yu, J.; Deng, Y.; Hu, X.; Wang, T.; Luo, X. Facile synthesis of MoS2/Bi2WO6 nanocomposites for enhanced CO2 photoreduction activity under visible light irradiation. Appl. Surf. Sci. 2017, 403, 230–239. [Google Scholar] [CrossRef]
  48. Meng, A.; Wu, S.; Cheng, B.; Yu, J.; Xu, J. Hierarchical TiO2/Ni(OH)2 composite fibers with enhanced photocatalytic CO2 reduction performance. J. Mater. Chem. A 2018, 6, 4729–4736. [Google Scholar] [CrossRef]
  49. Zhao, D.; Xuan, Y.; Zhang, K.; Liu, X. Highly selective production of ethanol over hierarchical Bi@Bi2MoO6 composite via bicarbonate-assisted photocatalytic CO2 reduction. Chemsuschem 2021, 14, 3293–3302. [Google Scholar] [CrossRef]
  50. Xu, Y.; Wang, S.; Yang, J.; Han, B.; Nie, R.; Wang, J.; Wang, J.; Jing, H. In-situ grown nanocrystal TiO2 on 2D Ti3C2 nanosheets for artificial photosynthesis of chemical fuels. Nano Energy 2018, 51, 442–450. [Google Scholar] [CrossRef]
  51. Li, H.; Li, C.; Han, L.; Li, C.; Zhang, S. Photocatalytic reduction of CO2 with H2O on CuO/TiO2 catalysts. Energy Sources Part A 2016, 38, 420–426. [Google Scholar] [CrossRef]
  52. Abou Asi, M.; He, C.; Su, M.; Xia, D.; Lin, L.; Deng, H.; Xiong, Y.; Qiu, R.; Li, X.-z. Photocatalytic reduction of CO2 to hydrocarbons using AgBr/TiO2 nanocomposites under visible light. Catal. Today 2011, 175, 256–263. [Google Scholar] [CrossRef]
  53. Korovin, E.; Selishchev, D.; Kozlov, D. Photocatalytic CO2 reduction on the TiO2 P25 under the high power UV-LED irradiation. Top. Catal. 2016, 59, 1292–1296. [Google Scholar] [CrossRef]
  54. Seeharaj, P.; Kongmun, P.; Paiplod, P.; Prakobmit, S.; Sriwong, C.; Kim-Lohsoontorn, P.; Vittayakorn, N. Ultrasonically-assisted surface modified TiO2/rGO/CeO2 heterojunction photocatalysts for conversion of CO2 to methanol and ethanol. Ultrason. Sonochem. 2019, 58, 104657. [Google Scholar] [CrossRef] [PubMed]
  55. Wei, Y.; Cheng, Z.; Lin, J. An overview on enhancing the stability of lead halide perovskite quantum dots and their applications in phosphor-converted LEDs. Chem. Soc. Rev. 2019, 48, 310–350. [Google Scholar] [CrossRef] [PubMed]
  56. Yan, Y.; Gong, J.; Chen, J.; Zeng, Z.; Huang, W.; Pu, K.; Liu, J.; Chen, P. Recent advances on graphene quantum dots: From chemistry and physics to applications. Adv. Mater. 2019, 31, 1808283. [Google Scholar] [CrossRef]
  57. Dai, W.; Yu, J.; Luo, S.; Hu, X.; Yang, L.; Zhang, S.; Li, B.; Luo, X.; Zou, J. WS2 quantum dots seeding in Bi2S3 nanotubes: A novel Vis-NIR light sensitive photocatalyst with low-resistance junction interface for CO2 reduction. Chem. Eng. J. 2020, 389, 123430. [Google Scholar] [CrossRef]
  58. Dai, W.; Xiong, W.; Yu, J.; Zhang, S.; Li, B.; Yang, L.; Wang, T.; Luo, X.; Zou, J.; Luo, S. Bi2MoO6 quantum dots in situ grown on reduced graphene oxide layers: A novel electron-rich interface for efficient CO2 reduction. ACS Appl. Mater. Interfaces 2020, 12, 25861–25874. [Google Scholar] [CrossRef]
  59. Cheng, M.; Yang, S.; Chen, R.; Zhu, X.; Liao, Q.; Huang, Y. Visible light responsive CdS sensitized TiO2 nanorod array films for efficient photocatalytic reduction of gas phase CO2. Mol. Catal. 2018, 448, 185–194. [Google Scholar] [CrossRef]
  60. Xiao, J.-D.; Jiang, H.-L. Metal-organic frameworks for photocatalysis and photothermal catalysis. Acc. Chem. Res. 2018, 52, 356–366. [Google Scholar] [CrossRef]
  61. Xiao, X.; Zou, L.; Pang, H.; Xu, Q. Synthesis of micro/nanoscaled metal-organic frameworks and their direct electrochemical applications. Chem. Soc. Rev. 2020, 49, 301–331. [Google Scholar] [CrossRef] [PubMed]
  62. Meng, Y.; Zhang, L.; Jiu, H.; Zhang, Q.; Zhang, H.; Ren, W.; Sun, Y.; Li, D. Construction of g-C3N4/ZIF-67 photocatalyst with enhanced photocatalytic CO2 reduction activity. Mat. Sci. Semicon. Proc. 2019, 95, 35–41. [Google Scholar] [CrossRef]
  63. 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]
  64. Cardoso, J.; Stulp, S.; De Brito, J.; Flor, J.; Frem, R.; Zanoni, M. 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]
  65. Dai, W.; Xu, H.; Yu, J.; Hu, X.; Luo, X.; Tu, X.; Yang, L. Photocatalytic reduction of CO2 into methanol and ethanol over conducting polymers modified Bi2WO6 microspheres under visible light. Appl. Surf. Sci. 2015, 356, 173–180. [Google Scholar] [CrossRef]
  66. Pastrana-Martínez, L.; Silva, A.; Fonseca, N.; Vaz, J.; Figueiredo, J.; Faria, J. Photocatalytic reduction of CO2 with water into methanol and ethanol using graphene derivative-TiO2 composites: Effect of pH and copper (I) oxide. Top. Catal. 2016, 59, 1279–1291. [Google Scholar] [CrossRef]
  67. Wang, L.; Wei, Y.; Fang, R.; Wang, J.; Yu, X.; Chen, J.; Jing, H. Photoelectrocatalytic CO2 reduction to ethanol via graphite-supported and functionalized TiO2 nanowires photocathode. J. Photoch. Photobio. A 2020, 391, 112368. [Google Scholar] [CrossRef]
  68. Wang, Y.; Jia, H.; Gong, H.; Zhou, L.; Qiu, Z.; Fang, X.; Du, T. Fabrication of trimodal porous silica/g-C3N4 nanotubes for efficient visible light photocatalytic reduction of CO2 to ethanol. Chem. Eng. J. 2021, 426, 130877. [Google Scholar] [CrossRef]
  69. Maimaitizi, H.; Abulizi, A.; Kadeer, K.; Talifu, D.; Tursun, Y. In situ synthesis of Pt and N co-doped hollow hierarchical BiOCl microsphere as an efficient photocatalyst for organic pollutant degradation and photocatalytic CO2 reduction. Appl. Surf. Sci. 2020, 502, 144083. [Google Scholar] [CrossRef]
  70. Li, N.; Li, Y.; Jiang, R.; Zhou, J.; Liu, M. Photocatalytic coupling of methane and CO2 into C2-hydrocarbons over Zn doped g-C3N4 catalysts. Appl. Surf. Sci. 2019, 498, 143861. [Google Scholar] [CrossRef]
  71. Aslam, U.; Rao, V.G.; Chavez, S.; Linic, S. Catalytic conversion of solar to chemical energy on plasmonic metal nanostructures. Nat. Catal. 2018, 1, 656–665. [Google Scholar] [CrossRef]
  72. Agrawal, A.; Cho, S.H.; Zandi, O.; Ghosh, S.; Johns, R.W.; Milliron, D.J. Localized surface plasmon resonance in semiconductor nanocrystals. Chem. Rev. 2018, 118, 3121–3207. [Google Scholar] [CrossRef]
  73. Vu, N.N.; Kaliaguine, S.; Do, T.O. Plasmonic photo catalysts for sunlight-driven reduction of CO2: Details, developments, and perspectives. Chemsuschem 2020, 13, 3967–3991. [Google Scholar] [CrossRef]
  74. Li, S.; Miao, P.; Zhang, Y.; Wu, J.; Zhang, B.; Du, Y.; Han, X.; Sun, J.; Xu, P. Recent advances in plasmonic nanostructures for enhanced photocatalysis and electrocatalysis. Adv. Mater. 2021, 33, 2000086. [Google Scholar] [CrossRef]
  75. Yu, H.; Sun, C.; Xuan, Y.; Zhang, K.; Chang, K. Full solar spectrum driven plasmonic-assisted efficient photocatalytic CO2 reduction to ethanol. Chem. Eng. J. 2022, 430, 132940. [Google Scholar] [CrossRef]
  76. Lertthanaphol, N.; Pienutsa, N.; Chusri, K.; Sornsuchat, T.; Chanthara, P.; Seeharaj, P.; Kim-Lohsoontorn, P.; Srinives, S. One-step hydrothermal synthesis of precious metal-doped Titanium dioxide-graphene oxide composites for photocatalytic conversion of CO2 to ethanol. ACS Omega 2021, 6, 35769–35779. [Google Scholar] [CrossRef] [PubMed]
  77. Zeng, L.Z.; Wang, Z.Y.; Wang, Y.K.; Wang, J.; Guo, Y.; Hu, H.H.; He, X.F.; Wang, C.; Lin, W.B. Photoactivation of Cu centers in metal-organic frameworks for selective CO2 conversion to ethanol. J.Am.Chem.Soc. 2020, 142, 75–79. [Google Scholar] [CrossRef]
  78. Wang, G.; He, C.T.; Huang, R.; Mao, J.J.; Wang, D.S.; Li, Y.D. 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]
  79. Almomani, F.; Bhosale, R.; Khraisheh, M.; Kumar, A.; Tawalbeh, M. Photocatalytic conversion of CO2 and H2O to useful fuels by nanostructured composite catalysis. Appl. Surf. Sci. 2019, 483, 363–372. [Google Scholar] [CrossRef]
  80. Lu, Y.; Cao, H.; Xu, S.; Feng, W.; Hou, G.; Tang, Y.; Zhang, H.; Zheng, G. CO2 photoelectroreduction with enhanced ethanol selectivity by high valence rhenium-doped copper oxide composite catalysts. J. Colloid Interface Sci. 2021, 599, 497–506. [Google Scholar] [CrossRef] [PubMed]
  81. Abou Asi, M.; Zhu, L.; He, C.; Sharma, V.K.; Shu, D.; Li, S.; Yang, J.; Xiong, Y. Visible-light-harvesting reduction of CO2 to chemical fuels with plasmonic Ag@AgBr/CNT nanocomposites. Catal. Today 2013, 216, 268–275. [Google Scholar] [CrossRef]
  82. An, C.; Wang, J.; Qin, C.; Jiang, W.; Wang, S.; Li, Y.; Zhang, Q. Synthesis of Ag@AgBr/AgCl heterostructured nanocashews with enhanced photocatalytic performance via anion exchange. J. Mater. Chem. 2012, 22, 13153–13158. [Google Scholar] [CrossRef]
  83. Becerra, J.; Duc-Trung, N.; Gopalakrishnan, V.-N.; Trong-On, D. 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]
  84. Compagnoni, M.; Villa, A.; Bandori, E.; Morgan, D.J.; Prati, L.; Dimitratos, N.; Rossetti, I.; Ramis, G. Surface probing by spectroscopy on titania-supported gold nanoparticles for a photoreductive application. Catalysts 2018, 8, 623. [Google Scholar] [CrossRef] [Green Version]
  85. Elavarasan, M.; Yang, W.; Velmurugan, S.; Chen, J.-N.; Yang, T.C.K.; Yokoi, T. Highly efficient photothermal reduction of CO2 on Pd2Cu dispersed TiO2 photocatalyst and operando DRIFT spectroscopic analysis of reactive intermediates. Nanomaterials 2022, 12, 332. [Google Scholar] [CrossRef]
  86. Guérinot, A.; Cossy, J. Cobalt-catalyzed cross-couplings between alkyl halides and grignard reagents. Acc. Chem. Res. 2020, 53, 1351–1363. [Google Scholar] [CrossRef]
  87. Shu, T.; Cossy, J. Enantioselective cross-couplings between halide derivatives and organometallics by using iron and cobalt catalysts.: Formation of C-C bonds. Chem. Eur. J. 2021, 27, 11021–11029. [Google Scholar] [CrossRef]
  88. Liu, L.; Puga, A.V.; Cored, J.; Concepcion, P.; Pérez-Dieste, V.; García, H.; Corma, A. Sunlight-assisted hydrogenation of CO2 into ethanol and C2+ hydrocarbons by sodium-promoted Co@ C nanocomposites. Appl. Catal. B Environ. 2018, 235, 186–196. [Google Scholar] [CrossRef]
  89. Lebarbier, V.M.; Karim, A.M.; Engelhard, M.H.; Wu, Y.; Xu, B.Q.; Petersen, E.J.; Datye, A.K.; Wang, Y. The effect of zinc addition on the oxidation state of cobalt in Co/ZrO2 catalysts. Chemsuschem 2011, 4, 1679–1684. [Google Scholar] [CrossRef]
  90. Huang, Y.; Du, P.; Shi, W.-X.; Wang, Y.; Yao, S.; Zhang, Z.-M.; Lu, T.-B.; Lu, X. Filling COFs with bimetallic nanoclusters for CO2-to-alcohols conversion with H2O oxidation. Appl. Catal. B Environ. 2021, 288, 120001. [Google Scholar] [CrossRef]
  91. Peng, T.; Wang, K.; He, S.; Chen, X.; Dai, W.; Fu, X. Photo-driven selective CO2 reduction by H2O into ethanol over Pd/Mn-TiO2: Suitable synergistic effect between Pd and Mn sites. Catal. Sci. Technol. 2021, 11, 2261–2272. [Google Scholar] [CrossRef]
  92. Liu, L.; Wang, S.; Huang, H.; Zhang, Y.; Ma, T. Surface sites engineering on semiconductors to boost photocatalytic CO2 reduction. Nano Energy 2020, 75, 104959. [Google Scholar] [CrossRef]
  93. Fan, Y.; Zhang, C.; Mamatkulov, S.; Ruzimuradov, O.; Low, J. Semiconductor facet junctions for photocatalytic CO2 reduction. Pure Appl. Chem. 2022. [Google Scholar] [CrossRef]
  94. Xiang, Y.; Cheng, B.-R.; Li, D.-F.; Zhou, B.-X.; Yang, T.-F.; Ding, S.-S.; Huang, G.-F.; Pan, A.; Huang, W.-Q. Facile one-step in-situ synthesis of type-II CeO2/CeF3 composite with tunable morphology and photocatalytic activity. Ceram. Int. 2016, 42, 16374–16381. [Google Scholar] [CrossRef]
  95. Boltersdorf, J.; Forcherio, G.T.; McClure, J.P.; Baker, D.R.; Leff, A.C.; Lundgren, C. Visible light-promoted plasmon resonance to induce “hot” hole transfer and photothermal conversion for catalytic oxidation. J. Phys. Chem. C 2018, 122, 28934–28948. [Google Scholar] [CrossRef]
Scheme 1. Schematic illustration of the catalysts used for photocatalytic CO2 conversion to ethanol.
Scheme 1. Schematic illustration of the catalysts used for photocatalytic CO2 conversion to ethanol.
Catalysts 12 01549 sch001
Figure 1. Band structure of several typical semiconductors with respect to CO2 reduction potentials towards different products at pH = 7. Reproduced with permission from reference [33].
Figure 1. Band structure of several typical semiconductors with respect to CO2 reduction potentials towards different products at pH = 7. Reproduced with permission from reference [33].
Catalysts 12 01549 g001
Figure 2. (a) Schematic illustration of the photocatalytic CO2 reduction to alcohols over TiO2 nanotubes, and (b) proposed mechanism of photocatalytic CO2 reduction to methanol and ethanol over TiO2 nanotubes. Reproduced with permission from reference [23].
Figure 2. (a) Schematic illustration of the photocatalytic CO2 reduction to alcohols over TiO2 nanotubes, and (b) proposed mechanism of photocatalytic CO2 reduction to methanol and ethanol over TiO2 nanotubes. Reproduced with permission from reference [23].
Catalysts 12 01549 g002
Figure 3. Proposed reaction pathway of photocatalytic CO2 reduction to methanol and ethanol. Reproduced with permission from reference [42].
Figure 3. Proposed reaction pathway of photocatalytic CO2 reduction to methanol and ethanol. Reproduced with permission from reference [42].
Catalysts 12 01549 g003
Figure 4. Proposed mechanism for photocatalytic CO2 reduction to methanol and ethanol over a TiO2/rGO/CeO2 catalyst. Reproduced with permission from reference [54].
Figure 4. Proposed mechanism for photocatalytic CO2 reduction to methanol and ethanol over a TiO2/rGO/CeO2 catalyst. Reproduced with permission from reference [54].
Catalysts 12 01549 g004
Figure 5. Schematic illustration of the structure of WS2 QD/Bi2S3. Reproduced with permission from reference [57].
Figure 5. Schematic illustration of the structure of WS2 QD/Bi2S3. Reproduced with permission from reference [57].
Catalysts 12 01549 g005
Figure 6. Schematic illustration of photoexcited electron–hole separation process over g-C3N4/CuO@MIL-125(Ti). Reproduced with permission from reference [63].
Figure 6. Schematic illustration of photoexcited electron–hole separation process over g-C3N4/CuO@MIL-125(Ti). Reproduced with permission from reference [63].
Catalysts 12 01549 g006
Figure 7. (a) Free energy diagram for CO2 reduction to ethanol and ethylene on the Cu@Ni interface, (b) free energy diagram for CO2 reduction to ethanol and ethylene on Cu (100) surface. The red line represents the lowest energy path. The Cu, Ni, C, O, and H atoms are shown in brown, blue, gray, red, and white, respectively. Reproduced with permission from reference [75].
Figure 7. (a) Free energy diagram for CO2 reduction to ethanol and ethylene on the Cu@Ni interface, (b) free energy diagram for CO2 reduction to ethanol and ethylene on Cu (100) surface. The red line represents the lowest energy path. The Cu, Ni, C, O, and H atoms are shown in brown, blue, gray, red, and white, respectively. Reproduced with permission from reference [75].
Catalysts 12 01549 g007
Figure 8. Generation of CuI via light irradiation over a CuII(HxPO4)y@Ru-Uio catalyst (* represents the active site). Reproduced with permission from reference [77].
Figure 8. Generation of CuI via light irradiation over a CuII(HxPO4)y@Ru-Uio catalyst (* represents the active site). Reproduced with permission from reference [77].
Catalysts 12 01549 g008
Figure 9. Yields of products over different de-doped CuO/TiO2-NTs. Reproduced with permission from reference [80].
Figure 9. Yields of products over different de-doped CuO/TiO2-NTs. Reproduced with permission from reference [80].
Catalysts 12 01549 g009
Figure 10. (a,b) Effects of loading of Au on the photocatalytic activity of Au/ZIF-67 in a photocatalytic CO2 reduction. (c) Proposed mechanism for photocatalytic CO2 reduction to methanol and ethanol over Au/ZIF-67. Reproduced with permission from reference [83].
Figure 10. (a,b) Effects of loading of Au on the photocatalytic activity of Au/ZIF-67 in a photocatalytic CO2 reduction. (c) Proposed mechanism for photocatalytic CO2 reduction to methanol and ethanol over Au/ZIF-67. Reproduced with permission from reference [83].
Catalysts 12 01549 g010
Table 1. Brief summary of some typical catalysts for photocatalytic CO2 reduction to ethanol.
Table 1. Brief summary of some typical catalysts for photocatalytic CO2 reduction to ethanol.
Catalyst Category CatalystReaction ConditionPerformanceRef.
Pristine semiconductorsTiO2Reactor—home-made glass reactor, 50 mm in diameter and 100 mm in height;
Reactant—10 mL deionized water and water saturated CO2;
Light source—100 W Xe lamp, 35 mW cm−2.
Ethanol formation rate of ~9.0 nmol cm−2 h−1. [23]
Pristine semiconductorsBi2MoO6Reactor—closed vessel; reactant—50 mL deionized water and saturated CO2;
Light source—300 W Xe arc lamp (PLS-SXE300) with an ultraviolet cutoff filter (λ ≥ 420 nm).
Ethanol yield of 4.7 μmol g−1 h−1.[34]
Semiconductors with vacancy sitesReduced HCa2Ta3O10Reactor—an in situ closed circulation system;
Reactant—CO2 saturated with H2O vapor;
Light source—150 W Xe arc lamp, 100 mW cm−2.
Ethanol yield of 113.0 μmol g−1 h−1.[43]
HeterojunctionsTiO2/Ti3C2Reactor—two electrode system;
Reactant—0.1 M KHCO3 aqueous solution (pH = 6.8, 50 mL) saturated by CO2;
Light source—300 W xenon lamp (PLS-SXE300/300UV) with 200 mW cm−2 light intensity;
External bias potential—-0.6 V.
Ethanol formation rate of ~10.0 μmol cm−2 h−1.[50]
HeterojunctionsTiO2/rGO/CeO2Reactor—sealed photocatalytic reactor;
Reactant—150 mL distilled water with saturated CO2;
Light source—UV light (a 15 W UV-C mercury lamp, peak light intensity 254 nm);
Catalyst—0.15 g.
Ethanol yield of 271.0 μmol g−1 h−1.[54]
Hybrid catalysts constructed between a semiconductor and a non-semiconductor materialWS2 QD/Bi2S3Reactor—closed 200 mL quartz glass reactor;
Reactant—50 mL of ultrapure water with saturated CO2;
Light source—300 W Xe arc lamp (PLS-SXE300).
Ethanol yield of 7.0 μmol g−1 h−1.[57]
Hybrid catalysts constructed between a semiconductor and a non-semiconductor materialg-C3N4/CuO@MIL-125(Ti)Reactor—visual micro autoclave lined with 100 mL polytetrafluoroethylene;
Reactant—1.0 mL water and 0.3% CO2;
Pressure—1.0 MPa;
Light source—300 W Xe lamp, 326.1W m−2.
Ethanol yield of 501.9 μmol g−1 h−1.[63]
Cu-based catalystsSrTiO3/Cu@Ni/TiNReactor—Labsolar 6 A system (Beijing Perfectlight Technology Co., Ltd.);
Reactant—10 mL ultrapure water and saturated CO2;
Light source—300 W Xe lamp, 600 mW cm−2.
Ethanol yield of 21.3 μmol g−1 h−1 and an ethanol selectivity of 79%.[75]
Ag-based catalystsAg@AgBr/CNT Reactor—stainless steel vessel;
Reactant—100 mL 0.2 M KHCO3 solution, pure CO2 (99.99%) with a pressure of 7.5 MPa;
Light source—A 150 W Xe lamp (Shanghai Aojia Lighting Appliance Co. Ltd.) with UV cutoff filter (λ > 420 nm).
Ethanol yield of 5.0 μmol g−1 h−1.[81]
Au-based catalystsAu/ZIF-67 Reactor—horizontal-glass-type photoreactor;
Reactant—10 mL of aqueous solution with 10 wt % triethanolamine, 67 mg NaHCO3, purged with CO2;
Light source—Abet 103 with light intensity fixed at 150 mW cm−2.
Ethanol yield of 0.5 mmol g−1 h−1.[83]
Plasmonic alloy-based catalystsAuCu/g-C3N4Reactor—high-temperature and-high pressure CEL-HPR reactor with a volume of 250 mL (Beijing Zhongjiao Jinyuan Technology Co., Ltd.);
Reactant—100 mL ultrapure water, high-purity CO2 (99.999%), 0.8 MPa;
Light source—300 W Xe lamp (λ > 420 nm),
Temperature—80–160 °C.
An ethanol yield and selectivity of 0.9 mmol g−1 h−1 and 93.1%, respectively. [27]
Co-based catalystsNa-Co@CReactor—quartz cell reactor;
Reactant—CO2, N2, and H2 of 20, 20, and 100 cm3 at standard conditions, with a final pressure of ~2.8 bar;
Light source—Xe lamp (1000 W) coupled with an AM1.5 filter;
Temperature—235 °C.
~6% selectivity towards ethanol.[88]
Pd-based catalystsPdIn@N3-COF Reactor—double-walled 80 mL quartz photoreactor;
Reactant—10 mL ultrapure water with saturated CO2;
Light source—300 W Xe lamp (CEL-HXF300, CEAULICHT) with a 400 nm filter.
A total yield toward alcohols of 33.3 μmol g−1 h−1 and a selectivity to ethanol of 26%.[90]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, D.; Hao, C.; Liu, H.; Zhang, R.; Li, Y.; Guo, J.; Vilancuo, C.C.; Guo, J. Photocatalytic CO2 Conversion to Ethanol: A Concise Review. Catalysts 2022, 12, 1549. https://doi.org/10.3390/catal12121549

AMA Style

Li D, Hao C, Liu H, Zhang R, Li Y, Guo J, Vilancuo CC, Guo J. Photocatalytic CO2 Conversion to Ethanol: A Concise Review. Catalysts. 2022; 12(12):1549. https://doi.org/10.3390/catal12121549

Chicago/Turabian Style

Li, Dezheng, Chunnan Hao, Huimin Liu, Ruiqi Zhang, Yuqiao Li, Jiawen Guo, Clesio Calebe Vilancuo, and Jiapeng Guo. 2022. "Photocatalytic CO2 Conversion to Ethanol: A Concise Review" Catalysts 12, no. 12: 1549. https://doi.org/10.3390/catal12121549

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