Next Article in Journal
Application Prospect of K Used for Catalytic Removal of NOx, COx, and VOCs from Industrial Flue Gas: A Review
Next Article in Special Issue
CO2 Hydrogenation to Methane over Ni-Catalysts: The Effect of Support and Vanadia Promoting
Previous Article in Journal
Isolation Strategy towards Earth-Abundant Single-Site Co-Catalysts for Photocatalytic Hydrogen Evolution Reaction
Previous Article in Special Issue
Visible-Light Photocatalyst to Remove Indoor Ozone under Ambient Condition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Low-Dimensional Nanostructured Photocatalysts for Efficient CO2 Conversion into Solar Fuels

1
Department of Photonics, National Sun Yat-sen University, No. 70, Lien-Hai Rd, Kaohsiung 80424, Taiwan
2
Department of Chemistry, Faculty of Science, Assiut University, Assiut 71516, Egypt
3
Department of Engineering Science and Mechanics, Pennsylvania State University, University Park, PA 16802, USA
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(4), 418; https://doi.org/10.3390/catal11040418
Submission received: 24 February 2021 / Revised: 19 March 2021 / Accepted: 24 March 2021 / Published: 25 March 2021

Abstract

:
The ongoing energy crisis and global warming caused by the massive usage of fossil fuels and emission of CO2 into atmosphere continue to motivate researchers to investigate possible solutions. The conversion of CO2 into value-added solar fuels by photocatalysts has been suggested as an intriguing solution to simultaneously mitigate global warming and provide a source of energy in an environmentally friendly manner. There has been considerable effort for nearly four decades investigating the performance of CO2 conversion by photocatalysts, much of which has focused on structure or materials modification. In particular, the application of low-dimensional structures for photocatalysts is a promising pathway. Depending on the materials and fabrication methods, low-dimensional nanomaterials can be formed in zero dimensional structures such as quantum dots, one-dimensional structures such as nanowires, nanotubes, nanobelts, and nanorods, and two-dimensional structures such as nanosheets and thin films. These nanostructures increase the effective surface area and possess unique electrical and optical properties, including the quantum confinement effect in semiconductors or the localized surface plasmon resonance effect in noble metals at the nanoscale. These unique properties can play a vital role in enhancing the performance of photocatalytic CO2 conversion into solar fuels by engineering the nanostructures. In this review, we provide an overview of photocatalytic CO2 conversion and especially focus on nanostructured photocatalysts. The fundamental mechanism of photocatalytic CO2 conversion is discussed and recent progresses of low-dimensional photocatalysts for efficient conversion of CO2 into solar fuels are presented.

1. Introduction

The fast-developing modern technology and explosive world population growth have resulted in a huge demand for and consumption of energy. According to an investigation from the U.S. Energy Information Administration, more than 600 quadrillion Btus of energy were spent in 2020 and it is expected that the demand will continue to skyrocket annually. To meet this huge energy demand every year, energy production has been predominately dependent on fossil fuels such as coal, oil, or natural gas. The energy production by fossil fuels is inextricably linked to the gigantic CO2 emission of more than 30 billion metric tons every year and, in turn, the accumulated CO2 in our atmosphere is deemed to be the main cause of many environmental problems such as global warming and erratic weather patterns. In this context, there is great motivation to find a way of reducing atmospheric CO2 and producing energy at the same time. For these problems, CO2 conversion by photocatalyst materials under light illumination could be an expedient solution. This is because natural sunlight provides clean, renewable, and abundant energy, and photocatalysts can be activated by light energy from the Sun, while simultaneously consuming CO2 for energy production.
In 1978, by Halmann [1], the first demonstration of the photocatalytic conversion of CO2 in aqueous solution into liquid fuels such as methanol, formic acid, and formaldehyde was achieved over p-type gallium phosphide semiconductor. The same year, the photoartificial synthesis by SrTiO3 photocatalysts for CH4 production through the gas-solid phase reaction of CO2 and H2O was reported by Hemminger [2]. In 1979, another pioneering work by Fujishima and his coworkers introduced the artificial synthesis of solar fuels from a CO2-saturated electrolyte under light illumination. In this study, liquid CO2 was converted with various semiconductor photocatalysts such as TiO2, ZnO, CdS, GaP, SiC, and WO3 to produce methane, methanol, formaldehyde and formic acid [3]. Since these historical works mentioned above, several semiconductor materials have been investigated for the conversion of CO2 into useful fuels, including graphitic carbon nitride (g-C3N4), graphene, conjugated polymers, covalent organic framework, metal organic frameworks, metal chalcogenides, metal oxides, black phosphorus, bismuth-based materials, and perovskites [4,5,6,7,8,9,10,11,12,13,14]. Moreover, a variety of strategies and approaches have been applied to improve the photocatalyst performance through elemental doping, solid solution, heterostructure, nanostrucutralization, surface engineering and modification, crystal facet engineering, cocatalysts utilization, or dimensionality tailoring [15,16,17,18,19].
Of the strategies researched to boost the efficiency of CO2 conversion, decreasing the dimensionality and constructing nanostructures of the photocatalyst have attracted a lot of attention owing to their favorable advantages in photocatalysis: first, nanostructured photocatalysts suppress the carrier recombination due to their higher crystallinity compared to non-nanostructured materials [20,21,22]; second, the implementation of low dimensionality modifies the electronic structure of bulk materials due to the quantum confinement effect in semiconductors or localized surface plasmon resonance effect in noble metals at the nanoscale; third, low-dimensional materials possess larger surface-area-to-volume ratio in comparison with bulk materials, providing more reaction sites. All three features can contribute to improved solar-driven catalytic reactions [19,23].
In this review, we summarize the fundamental knowledge about CO2 conversion into useful fuels by photocatalysts and offer insights on the pivotal role of low dimensional structure in photocatalysts for efficient conversion of CO2 into chemical fuels such as CO, CH4, CH3OH, and C2H5OH that are useful sources of energy. We highlight the recent developments of low-dimensional nanostructures for photocatalysts toward chemical fuels production via the photoreduction of CO2. Finally, we provide the strategies to enhance the performance of CO2 conversion over low-dimensionally nanostructured photocatalysts.

2. The Main Fundamentals of CO2 Photoconversion into Solar Fuels and Hydrocarbon Species

2.1. Nature of CO2

Carbon dioxide (CO2) is one of the primary greenhouse gases but, at the same time, it is the main resource for solar fuel production when coupled with proton donors such as H2O for photocatalytic CO2 conversion. Hence, understanding the nature of the gaseous CO2 molecule itself is necessary for efficient utilization of photocatalysts. CO2 is a stable linear molecule among carbon compounds because of it is in the highest oxidation state of carbon, C+4 [19]. The CO2 molecule has two C=O bonds with a dissociation energy of ~750 kJ/mol, which is quite larger than those in other chemical bonds such as C-H (~430 kJ/mol) and C-C (~336 kJ/mol). For this reason, the reduction of CO2 to produce solar fuels requires additional energy to break the C=O bond and form, for example, a C-H bond [24]. Owing to its stability and strong bonding, photocatalytic reduction of CO2 into solar fuels can be achieved primarily with the support of proton donors such as H2O or H2 [25].

2.2. CO2 Adsorption on the Surface of Photocatalysts

The adsorption and activation of CO2 on a solid surface is one of the essential steps in achieving CO2 reduction and photocatalytic performance. The adsorption mechanism of CO2 on several semiconductor photocatalysts has been investigated [26,27]. For instance, the adsorption of CO2 on TiO2 surface has been investigated by Minot and coworkers [26]. Various adsorption modes of CO2 over a rutile TiO2 surface have been studied using first-principles calculations. The oxygen atom of CO2 molecules favors the interaction with the acidic titanium cation over the surface forming a Ti–OCO bond [26]. The adsorption of CO2 on a photocatalyst surface includes the interaction between the CO2 molecule and the surface atoms of the photocatalyst. This absorption may occur with a charge transfer from the photocatalyst to the linear and stable CO2 molecule which can induce the formation of partially charged and bent adsorbate, CO 2 δ · , and this adsorbate can form three different molecular structures, as shown in Figure 1: oxygen coordination, carbon coordination, and mixed coordination [24,28,29]. The beneficial feature of CO 2 δ · , is that it has decrease in the lowest unoccupied molecular orbital (LUMO) of the CO2 energy level as linear CO2 molecule transformed into the bent structure. This would facilitate the charge transfer between a photocatalyst and CO 2 δ · , [24,28,29].
In the photocatalytic reaction, the photocatalyst donates electron to the adsorbed species on the surface to initiate the reduction process of CO2 in the presence of protons. As shown in Table 1, the solar fuel production is determined by the number of electrons and protons included in the reaction [20,24]. For example, two electrons are required for CO evolution, while methane production is an eight-electron reaction. The adsorption of CO2 on the photocatalyst surface can be improved by a variety of strategies. First, decreasing the structural dimensionality of photocatalyst can improve the surface area of the photocatalyst to allow more adsorption. Second, enhanced density of active sites by incorporation of surface defects, such as oxygen and sulfur vacancies, can improve the CO2 adsorption. Third, utilization of noble metal nanoparticles can help improve the adsorption due to lowered activation energy of the CO2 reduction [24,30].

2.3. The Mechanism of Efficient CO2 Photoconversion

Upon absorption of light over the photocatalyst, charge carrier pairs are generated to achieve the photosynthesis of solar fuels accompanying the water splitting as shown in Table 1. To execute the conversion of CO2 into useful fuels from the thermodynamic point of view, the conduction band minimum (CBM), and the valence band maximum (VBM) of a photocatalyst should bracket the redox potential of CO2 and the oxidation potential of water, respectively, as shown in Table 1 and Figure 2. As results of the reactions, various solar fuels can be formed dependent on the number of electrons and protons in the presence of CO2 and water under the illumination.
The artificial photosynthesis of solar fuels using semiconductor photocatalysts consists of three essential steps, as described in Figure 3. Firstly, incident photons of light with energy higher than that of semiconductor band gap (Eg) induce the generation of the electron-hole pairs (Process (i)). Secondly, the photogenerated charge carriers are transferred to the photocatalyst surface (Process (ii)). Thirdly, the electrons and holes react on the surface of photocatalyst with CO2 and H2O for evolution of solar fuels (Process (iii)). For efficient photocatalytic conversion of CO2 into fuels, the ideal semiconductor photocatalyst should have optimized band gap for efficient light harvesting and photocarrier generation, facile charge separation and transportation, vigorously activated sites and high surface area for maximum adsorption of CO2 and water [30,31].

3. State-of-the-Art Progress in Low-Dimensional Nanostructured Photocatalyst

Low-dimensional materials are defined as nanostructured materials characterized by sizes smaller than 100 nm in at least one dimension. Low-dimensional materials include insulators, semiconductors, and metals [33]. In terms of their morphologies, the low-dimensional materials can be classified into three categories as can be seen in Figure 4. First, zero-dimensional materials such as nanoparticles, nanosphere, and quantum dots (QDs). Second, one-dimensional materials including nanowires (NWs), nanotubes (NTs), nanobelts (NBs), nanorods (NRs), and nanoribbons. Third, two-dimensional materials which includes nanosheets and thin film [23,34]. Low-dimensional semiconductor materials are widely used for efficient artificial synthesis of solar fuels and constructed with various materials such as titanium dioxide [21], carbon [35], phosphorus [36], transition metal dichalcogenides [37], group IV monochalcogenides (MX, M=Sn, Si, Ge, and X=S, Se) [38], conjugated polymers [39], and perovskite [40]. In this section, we introduce the important features of low-dimensional photocatalytic materials and recent advances of low-dimensional nanostructured photocatalysts related to CO2 conversion into solar fuels.

3.1. Low-Dimensional Semiconductor Materials in Photocatalytic Application

Low-dimensional semiconductor materials possess unique optical and electronic properties which are distinct from those of their bulk counterparts. The motion of electrons in low-dimensional materials is confined by their nanostructures, which is known as quantum confinement effect [41]. The quantum confinement results in the unique properties of the low-dimensional nanomaterials that would benefit the performance of photocatalysts.
In general, the optical properties of low-dimensional semiconductor materials are affected by this quantum confinement. Osterloh et al. [42] reported that the optical and photocatalytic properties of CdSe nanocrystals can be tuned by the quantum confinement effect. CdSe QDs of diameter ranged between 1.8 and 6 nm were fabricated by a facile aqueous method using 2-mercaptoethanol as a surface capping agent and 2-propanol as precipitating agent. The decrease in the size of CdSe QDs is accompanied with widening of optical band gap, which is directly related to the efficient utilization of the photocatalyst [42]. In addition, the quantum confinement effect has influence on the electrical properties of the semiconductors such as modulation of the potential energies of conduction and valence bands. In general, the electronic band structure of low-dimensional materials varies with dimensionality as a result of the different distribution of densities of state. The dimensional reduction of semiconductors is accompanied with energy discontinuity compared to 3D materials. Namely, the density of states (DOS) of 3D materials are distributed in E1/2 in terms of energy, E while there is a step-like DOS distribution in 2D materials. The DOS for quantum wires is proportional to E−1/2. Once the electron is confined in all directions as in quantum dots, the density of states become series of δ-peak functions and exists only at discrete energies [34]. In addition, reducing the dimensionality of the semiconductor photocatalyst could contribute to increase in DOS. Xie and coworkers fabricated ultrathin TiO2 flakes via sonication of lamellar TiO2-octylamine. It is reported that the transformation of 3D TiO2 into ultrathin TiO2 flakes facilitate the transfer of the photoexcited electron into the conduction band of TiO2. They observed an increased DOS at the edge of band, suggesting enhanced electrical properties of TiO2 materials [43].
These unique electronic properties of low-dimensional materials also benefit the CO2 conversion performance of photocatalysts. TiO2 nanosheets of 1.66-nm thickness reported by Xie and coworkers exhibited formate production rate of 1.9 µmol∙g−1∙h−1 and this conversion rate is 450 times higher than that by bulk TiO2 [43]. Chen and coworkers have functionalized graphene QDs by addition of 1,1’-bi(2-naphthylamine) to coal graphene QDs followed by solvothermal reaction at 180 °C to form C=N bonds [11]. The as-synthesized graphene QD displayed outstanding activity of 0.695 µmol∙g−1∙h−1 toward CO2 conversion into methanol under visible light radiation, which is three times higher than that of coal graphene QDs. In addition, the band gap of functionalized graphene QDs has been engineered for modulation of valence band maximum and conduction band minimum which can control the photocatalyst selectivity toward the desired product [11]. Recently, Gong and coauthors reported the confinement effect in low-dimensional C3N4/CdSe photocatalyst through the photocatalytic conversion of CO2 into methanol. Herein, CdSe QDs with 1.2–4.1 nm of diameters were synthesized using 2-mercaptopropionic acid and 2-propanol as surface-passivating and size-selective precipitating agents, respectively. The CdSe QD was loaded on polymeric C3N4 using an impregnation method. It was found that the CdSe QDs with 2.2 nm particle size achieved 73% of selectivity and 186.4 µmol∙g−1∙h−1 of methanol production. The methanol activity and selectivity were enhanced by virtue of tuning the conduction band potential [44].
We have discussed some general features of low-dimensional photocatalyst above. Hereafter, we focus more specifically on research in terms of dimensions and introduce more literature reviews on low-dimensional photocatalytic application. Table 2 summaries various low-dimensional photocatalysts that have been introduced including the examples following.

3.1.1. Zero-Dimensional Structure-Based Photocatalysts for CO2 Conversion

Zero-dimensional (0D) structures, namely quantum dots (QDs), are nanostructures in which the motion of electrons is spatially confined in three dimensions because of their ultrasmall size (typically, 1.2–10 nm), showing unique optical and electrical properties. With these beneficial properties, various QD photocatalysts have been demonstrated with various materials, including CdS [75], PbS [76], ZnSe [77], CuInS2 [78], BiVO4 [79], NiS2 [80], carbon, graphene [11,81], black phosphorus [13,82], Bi2S3 [83], and perovskite [45,46,48]. Other than these QD materials mentioned here, several 0D materials have been suggested as promising candidate for efficient CO2 conversion.
One of those promising QD materials is carbon QDs (CQD). CQDs are small amorphous carbon nanoparticles with sp3 hybridization and possess less than 10 nm in size [84]. Unlike conventional semiconductor QDs, CQDs absorb broad range of the solar spectrum extending to near infrared region due to π-plasmon absorption in the core carbon nanocrystals [85]. The light absorption in the UV-visible region could be ascribed to the π→π* transition of conjugated carbon atoms. Besides, the photoluminescent characteristics can be controlled by variation of the size of nanoparticles [86]. The excellent electronic conductivity of carbon dots causes promoted charge carrier separation and transfer. In addition, CQDs have ability to decrease the activation energy of the CO2 reaction, consequently the adsorption of CO2 molecules on the photocatalyst surface has been enhanced [87].
Another great candidate is graphene QDs (GQDs). GQDs are crystalline material with sp2 hybridized carbon. The size of GQD is ranged between 2 and 10 nm. In the past few years, GQDs have attracted tremendous attention for energy conversion ascribed not only to the quantum confinement and edge effects but also to its electronic and physiochemical properties. GQD is noted for its oxygenated functional groups on the surface which can act as active sites for some molecules such as CO2 [88].
Perovskite QDs (PQDs) are also attractive materials for 0D structure-based photocatalysts owing to their innate properties such as long diffusion length and high absorption coefficient [89,90] and their photophysical properties such as tunable band gap and high carriers mobility when they are fabricated as nanocrystals. Zhou and coworkers synthesized stable lead-free Cs2AgBiBr6 perovskite QDs via hot injection method as shown in Figure 5a [47]. The synthesized QDs demonstrate the uniform distribution of cubic Cs2AgBiBr6 nanocrystals with 9.5 nm of the average size when prepared at the optimum injection temperature of 200 °C. The Cs2AgBi Br6 nanocrystals (NCs) were utilized to convert CO2 into solar fuels. The prepared Cs2AgBiBr6 NCs possess a matched conduction band with the redox potentials of CO2 to produce CH4 and CO. As shown in Figure 5b, the enlarged band gap of the NCs (2.52 eV) compared to their bulk counterpart (1.95 eV) is attributed to the quantum confinement effect. The nanocrystals of Cs2AgBiBr6 showed improved photocatalytic performance to reduce CO2 into solar fuels compared to bulk Cs2AgBiBr6, which was 2.22 and 0.14 µmol∙g−1 of CO and CH4, respectively, after 6 h illumination. Zhou washed the prepared photocatalyst with ethanol to avoid the poor conductivity of the ligand layer. As plotted in Figure 5c, the washed nanocrystals boosted the evolution yield of CO2 and CH4 to 14.1 and 9.6 µmol∙g−1 after 6 h illumination. The nanostructured Cs2AgBiBr6 are favorable for outstanding CO2 conversion due to its efficient charge carrier transfer compared to bulk Cs2AgBiBr6 [47].
Black phosphorus quantum dots (BPQDs) have been regarded as rising stars in the design of photocatalysts for CO2 conversion into value-added fuels. Fan and coworkers have decorated WO3 nanowires by BPQD of 3–4 nm size via a hydrothermal method. The designed BPQDs/WO3 photocatalyst showed a significant enhancement of CO2 conversion into CO and C2H4 [13]. Zhang et al. also reported that the loading of BPQDs on g-C3N4 can enhance the charge carrier separation, and consequently the conversion of CO2 into CO can be improved [82].
Bi2S3 QDs have been suggested to be excellent candidates to improve the photocatalyst performance [83]. Guo and his coworkers have introduced Bi2S3 QDs into 2D g-C3N4 via a hydrothermal method. Bi2S3 QDs could promote the visible light absorption, specific surface area, and the separation of photogenerated charge carriers. Consequently, the obtained nanocomposite showed higher CO2 conversion into CO with production yield of 54.74 µmol∙g−1.

3.1.2. One-Dimensional Structure-Based Photocatalyst for CO2 Conversion

One-dimensional (1D) structure refers to materials where the motion of the electron is confined in two dimensions. Typically, nanowires, nanotubes, and nanorods are common morphologies for 1D nanomaterials. Owing to this unique design, the 1D structure endows materials with excellent charge transport and extended carrier lifetimes due to the unique distribution of state density coupled with intrinsically enlarged reactivity due to a large surface area in comparison with bulk materials [21,51,91,92]. To utilize these interesting features, various semiconductor materials have been fabricated with a 1D structure and these nanostructures have attracted huge attention for improving the photocatalytic conversion of CO2 into solar fuels [93]. Among the materials, we focus on several promising and interesting materials for 1D structure-based photocatalysts.
TiO2 is one of the most researched photocatalytic materials due to its low cost, non-toxicity, and chemical stability. For simplicity, we say TiO2 but, in reality this name describes a number of different structures and stoichiometries. Compared to TiO2 nanoparticles, when they fabricated in a 1D structure, 1D-structured TiO2 provides higher photoreduction performance of CO2 toward solar fuels as a result of its distinct structure and properties. For instance, 1D-structured TiO2 such as nanorods and nanotubes possess low recombination rates of the photogenerated charge carriers [21]. Zou et al. synthesized one dimensional TiO2 nanotubes (TNT) and nanorods (TNR) by a one-step hydrothermal method [21]. TNTs have uniform open-ended tubular morphology with outer and inner diameters of 8.24 nm and 1.75 nm, respectively, as seen in Figure 6a, while TNRs were grown in the lengths of 200 nm–2 µm and diameter of 20–90 nm (Figure 6b). As shown in Figure 6c, the conversion performance of CO2 over TNTs and TNRs were about 2.33-fold and 1.48-fold higher than that of TiO2 nanoparticles with an average particle size of 16 nm. The enhanced performance of CO2 conversion by TNT and TNR resulted in an improved charge carrier migration. It is also found that the increase in the specific surface area provides more active sites over the TiO2 nanotubes [21].
However, TiO2 is a n-type semiconductor with a band gap ranging between 3 and 3.2 eV, causing light absorption of less than 7% from AM1.5G solar spectrum and low photocatalytic activity [94,95,96,97,98]. To overcome these innate disadvantages of TiO2, several strategies have been demonstrated. Shankar et al. prepared TiO2 nanotubes using a flame annealing method to achieve the higher CH4 production yield among all TiO2 stand-alone photocatalysts [99]. Liu et al. fabricated phosphate-modified TiO2 nanotubes with an inner diameter of 120 nm via an impregnation method for CO2 conversion into solar fuels [100]. The fabricated tubular TiO2 was formed in a composite with graphitic carbon nitride (g-C3N4) through P–O bonds for photoreduction of CO2 into acetic acid, formic acid, and methanol. Roy et al. fabricated multi-leg TiO2 nanotubes of an average diameter of 365 nm via an electrochemical anodization method [101]. Graphene oxide was wrapped around the multi-leg nanotubes by the electrostatic interaction for further performance enhancement. Li et al. prepared reduced TiO2 nanorods using a two-step hydrothermal method followed by reduction with sodium borohydride [102]. The prepared reduced TiO2 nanorods exhibited higher CO2 adsorption capacity due to the presence of oxygen vacancies. Holmium- and nitrogen-doped TiO2 nanorods were prepared using a hydrothermal method by Hangzhou and his coworkers. The as-prepared samples exhibited higher absorption ability in the visible region [103]. Ohno et al. reported the crystal facet effect of brookite TiO2 nanorods for efficient conversion of CO2 into methanol [24,104]. The engineering of the facets was achieved by a hydrothermal method using polyvinyl alcohol or polyvinyl pyrrolidone. It was found that the TiO2 nanorods with mainly 210 crystal facets can act as reduction sites for CO2 molecules. Amin et al. prepared TiO2 nanowires with different lengths by controlling the contact time with NaOH during the reaction [105]. Gold nanoparticles were loaded on the fabricated nanowires to convert CO2 into CO with evolution rate of 1237 µmol∙g−1∙h−1 as a result of the surface plasmonic effect of gold nanoparticles [105].
Graphitic carbon nitride (g-C3N4) is another promising material for photocatalytic CO2 conversion. g-C3N4 is a metal-free polymeric semiconductor photocatalyst. g-C3N4 has received great attention in photocatalysis research due to its promising photocatalytic performance. A tremendous effort has been devoted to improving the photocatalytic activity of g-C3N4 including elemental doping, nanostructuralization, hybridization, and surface modification [106,107]. Bulk g-C3N4 exhibited inferior photocatalytic CO2 reduction performance due to the low surface area and the higher recombination rate. Dimensionality control of g-C3N4 has been a promising tactic to improve the CO2 conversion performance to generate solar fuels. In particular, 1D-structured g-C3N4 are favorable in obtaining large surface area and excellent electron-hole pair separation. Xu and coworkers have demonstrated tubular g-C3N4 and explained the high reduction activity of nanotubes g-C3N4 by three factors [51]: (1) the efficient charge separation because of the characteristic tubular structure (2) the higher surface area of nanotubes g-C3N4 causing sufficient reactive sites and efficient CO2 adsorption and (3) the promoted optical absorption. Chen and coworkers focused on the transformation of two dimensional g-C3N4 nanosheets into one dimensional g-C3N4 nanotubes using incorporation of phosphine gas in the reaction to produce phosphorus-doped g-C3N4 nanotubes (P-g-C3N4) [50]. The morphological structure of g-C3N4 nanotubes was confirmed by TEM as shown in Figure 7a. It is found that the tubes have a diameter of 200 nm and a wall thickness of 30–50 nm. As shown in Figure 7b, the as-prepared P-g-C3N4 nanotubes exhibited enhancement in solar-to-fuels efficiency in comparison with that of 2D g-C3N4. The higher CO2 reduction to CH4 and CO ascribed not only to the enlarged surface area and extended photon absorption, but also to the band structure tailoring. Moreover, the fabricated g-C3N4 nanotubes exhibited improved charge transfer, photocurrent and reduced charge transfer resistance from PL, TPC, and EIS measurements, respectively.

3.1.3. Two-Dimensional Structure-Based Photocatalysts for CO2 Conversion

Two-dimensional (2D) materials with thicknesses of a few atoms to <100 nm, and flake-shaped morphology are known as nanosheets (NS) [108]. Owing to the quantum confinement effect, 2D materials are characterized by distinct electronic and optical properties compared to bulk materials. However, the optical and electronic characteristics of nanosheets are improved further when these materials are thinned to ultrathin nanosheets (mono layers) [33]. Graphene is the most common 2D material which has been studied in many fields, including photocatalysis, due to its high charge carrier mobility [109]. MoS2 [110,111,112], WS2 [113], boron nitride [114], g-C3N4 [115], black phosphorus [116,117,118], bismuth-based materials [14,57,64,119,120,121], perovskite [122], and Mxene [123] have been used as 2D photocatalysts to produce solar fuels. Two dimensional materials have sparked interest for the transformation of CO2 into solar fuels because the traditional semiconductor materials possess low photoreduction performance. On the other hand, 2D materials exhibited maximum surface area, minimum migration distance of photogenerated charge carriers, ability for formation of vacancy-type active sites, and dangling bonds at edges [33]. Ultrathin nanosheets forming grain boundaries-free cause facile charge carrier transfer and suppress recombination rates [124,125]. Current efforts are being focused on the development and utilization of 2D semiconductor materials to reduce CO2 effectively. Here, we highlight the state-of-the-art progress in utilization of 2D photocatalysts for CO2 conversion into fuels.
As reported by Ye et. al., 2D g-C3N4 nanosheets can be obtained from the exfoliation of bulk g-C3N4, the g-C3N4 nanosheets disperse uniformly with diameters ranging from 60 to 110 nm and a thickness of 1 nm [54]. The as-prepared 2D g-C3N4 exhibited efficient solar-to-fuel conversion, CO and CH4 evolution rate of 5.407 µmol∙g−1∙h−1 and 1.549 µmol∙g−1∙h−1, respectively, which are much higher than the values of bulk-C3N4 due to the higher CO2 adsorption capacity of g-C3N4 nanosheet compared to its bulk counterpart.
An ultrathin g-C3N4 nanosheet was prepared by exfoliation of bulk g-C3N4 in NH3 atmosphere as reported by Jaroniec et al. [55]. An FESEM image demonstrated the ultrathin morphology of the exfoliated g-C3N4 as shown in Figure 8b. Compared to the aggregated particles of bulk g-C3N4 in Figure 8a, NH3-exfoliated g-C3N4 is characterized with hierarchical structure assembled by ultrathin nanosheets. As shown in Figure 8c, the nanosheets consists of 9–10 layers of g-C3N4 that correspond to about 3 nm of thickness. The as-prepared g-C3N4 nanosheets show a higher CO2 adsorption affinity (by a factor of 4) over bulk g-C3N4 due to their high surface area. The quantum confinement effect shifted the band gap of 2D g-C3N4 to 2.98 eV. The remarkable improvement of CO2 reduction into CH4 and methanol over 2D g-C3N4 photocatalyst is supported by the enhanced charge separation and transfer supports as measured in TPC, EIS, and TRPL measurements (Figure 8d to Figure 8f, respectively).
Ultrathin silicon carbide (SiC) nanosheets were prepared using 2D reduced graphene oxide for artificial synthesis of CH4 by Han et al. [56]. It is found that the dangling bonds from the surface of SiC NSs facilitate more effective adsorption of CO2 molecules. This improves the photocatalytic performance of CO2 reduction and the measured production rate of CH4 by ultrathin SiC was higher than that both by bulk SiC and TiO2 [56].
Intrinsic defects on the surface of photocatalysts such as oxygen vacancies [126,127], zinc vacancies [128], nitrogen vacancies [129] and bismuth vacancies [130] have a direct influence on the charge separation and transfer processes. Consequently, the photocatalytic performance of CO2 reduction can be boosted. For instance, Liu and coworkers have been designed an aurivillius phase oxide, defective Bi2MoO6 ultrathin NSs to improve the photocatalytic reduction of CO2. The surface defects were formed by controlling the thickness of NSs to the atomic scale to facilitate the escape of surface atoms. The as prepared NSs exhibited 2.55 times higher CO production rate than that of bulk counterpart [57].
2D nanostructured bismuth-based materials such as bismuth oxyhalide (BiOX, X = Cl, Br and I) [120], Bi2MoO6 [57], Bi2WO6 [64], and BiVO4 [119] have been reported as promising materials for the photoconversion of CO2 into solar fuels. In spite of the low-efficiency of CO2 conversion over bismuth oxyhalide, there is a great interest in the development of BiOX-based photocatalysts. For instance, Zhou and coworkers have prepared oxygen vacancy BiOCl nanosheets of 4 nm thickness via a solvothermal method. The as-synthesized photocatalyst exhibits higher CO production under UV-vis irradiation compared to BiOCl. The superior CO2 photoconversion over BiOCl with oxygen vacancies is attributed to the improved exciton dissociation and enhanced photon absorption [120]. Ye et al. reported use of an ultrathin BiOBr nanosheet as a photocatalyst to improve the CO2 conversion into solar fuels [121].
The effect of halogen type in bismuth oxyhalide have been reported by Wang and his coworkers [14]. BiOX nanosheets were fabricated by hydrothermal and chemical precipitation methods. It is found that the BiOBr exhibited higher CO2 conversion performance under sunlight. The formation rates of CO and CH4 were 21.6 and 1.2 µmol∙g−1∙h−1, respectively. It is demonstrated that the oxygen vacancies on the surface, the modest band structure, and the ionic radius of Br were responsible for the efficient production of solar fuels.
Xu et al. have prepared highly stable black phosphorus monolayers via a green exfoliation method assisted by liquid nitrogen. The photocatalyst exhibited superior CO evolution rate of 112.6 µmol∙g−1∙h−1 under visible light radiation. The high stability of the as-prepared photocatalyst can be attributed to the introduction of hydroxy groups during the exfoliation process. The higher performance of the black phosphorus can be ascribed to band gap engineering and enhanced charge carriers dynamics [118].
Owing to the low specific area of Dion–Jacobson perovskites, Do and coworkers fabricated a 2D-reduced perovskite, HCa2Ta3O10, decorated with Cu and Pt to improve the surface of the photocatalyst. The as-prepared photocatalyst can convert CO2 into methanol and ethanol with formation rates of 7.4 and 113.2 µmol∙g−1∙h−1. The photoconversion performance of CO2 over the reduced perovskite was much higher than that of nonreduced perovskite due to the enhanced light absorption and the improved charge separation [122].

3.2. Low-Dimensional Metal Materials for CO2 Conversion

Metal materials also can have unique properties when they are prepared in a low-dimensional structure. Other than the fact that improved surface area results from a morphological change on the nanoscale, low-dimensionally structured metals can show resonance of surface electrons with the presence of incident light, which is the so-called plasmonic effect. Specifically, the oscillating electric field of the incident light makes surface plasmon resonance with oscillating surface electrons on metals and the resonance help electrons to become more energetic [131]. The resonance frequency varies depending on the materials and the wavelength of the plasmon resonances in nanocrystals depends on the shape and size of nanomaterials [95,132,133]. In general, the plasmon resonance wavelength can be blue shifted with a decrease in the material size [134,135].
Nanostructured metals, in particular those fabricated with noble metals such as Ag, Au, Cu, or Pt, are of special interest because they have shown superior catalytic activity and excellent selectivity in photocatalytic applications [136,137]. The incorporation of these plasmonic metal nanomaterials can act as efficient electron sinks in the photocatalytic reactions. The semiconductor photocatalysts forming a junction with these plasmonic metal nanoparticles exhibit remarkable improvements in the charge transfer, charge separation, and recombination rate and thereby they showed highly improved photocatalytic activity [138]. In addition, the plasmonic resonance of Ag, Au, and Cu occur with visible light [139] and excitation of plasmons on the surface of conducting materials under light illumination at their plasmon wavelength have led to enhancement in photon absorption and photocurrent response [134,135].
These striking features of low-dimensional metal materials have been applied to photocatalysts and have contributed to the general enhancement of photocatalytic performance. For instance, gold NPs influence the photocatalytic reduction of CO2 over TiO2 due to the hot electron injection and near field enhancement that can produce more efficient electron-hole pairs at the surface of the photocatalyst. The smaller size of Au nanoparticles was significantly enhanced the charge separation over Au-modified TiO2 while the larger size has direct outcome of the increased field enhancement [140]. The effect of surface plasmon resonance on TiO2 nanotubes was studied by Cheng and coworkers, Silver (Ag) nanoparticles were loaded on TiO2 nanotubes via an electrochemical deposition method to inject hot electrons to the nanotubes and accelerate the migration of electron-hole pairs to the photocatalyst surface [141]. The Ag nanoparticles have been shown to enhance the utilization of visible light. Consequently, the photocatalytic conversion of CO2 into CH4 and other hydrocarbon species is improved compared to the as-prepared nanotubes without Ag deposition and TiO2 nanoparticles.

4. Strategies for Enhancement in the Light-Driven CO2 Conversion over Low-Dimensional Photocatalysts

Applying low-dimensional structure to the photocatalytic system itself is a proven way for obtaining high CO2 conversion performance. However, further improvement is available by designing the structures or modifying the photocatalytic materials so that characteristics of photocatalysts or photocatalytic system are engineered. Here, we focus on these two major strategies for further enhancement of CO2 conversion by low-dimensional photocatalysts. Table 2 summarizes the examples mentioned in this section.

4.1. Construction of Junction Formed by Low-Dimensional Structures

Of the diverse strategies to promote CO2 conversion performance of photocatalysts, designing or formation of junctions in photocatalytic systems has seen some success by modifying optical and electrical properties through materials and interfaces [63]. Junctions are constructed by coupling of two or more semiconductor materials or metal materials. The formation of a junction allows valuable properties from various materials to be more available in a single system. That helps more efficient light absorption, charge separation and transfer, or more stable performance. Since the photocatalytic systems with junctions have single or multiple interfaces, the engineering of the interfacial characteristics between the materials is essential for efficient photocatalyst performance. Especially, the interface characteristics can influence on carrier behaviors through bulk or at interfaces (ex. Shockley-Read-Hall recombination).
The formation of junctions implies the presence of an internal electric field in the nanomaterial. This internal electric field can contribute to enhanced carrier behaviors such as carrier separation and transfer for the photo-induced charge carriers and to, in the end, the performance of light-driven CO2 conversion. The internal electric field can be induced by growth of a low-dimensional semiconductor on a low-dimensional semiconductor [23,107]. The combination of two or multiple low-dimensional materials could integrate the advantages of both single units and mitigate the shortcomings of single unit by the synergistic effect [142].
One of the advantages of semiconductor QDs is the quantum size effect which is responsible for the optical properties of the photocatalyst. Apart from the acceleration of charge separation and transfer process, the contact between 0D semiconductor and 1D semiconductor provides the nanocomposite with an additional properties such as excessive electroactive sites, high surface area, and homogenous dispersion [143]. 1D Bi2S3 nanotubes have outstanding ability to absorb visible and near infrared light. The tubular structure of Bi2S3 provides the photocatalytic reaction with more active sites than other morphologies [58]. The remarkable optical and electronic properties of tungsten disulfide (WS2) QDs can be realized due to the quantum confinement effect. WS2 QDs can be also dispersed uniformly on the surface of Bi2S3 forming Bi–S channels to facilitate the charge carrier separation transfer process. The designed 0D/1D nanocomposite exhibited outstanding photocatalytic reduction of CO2 into CH3OH and C2H5OH of 38.2 µmol∙g−1 and 27.8 µmol∙g−1 after 4 h radiation, respectively. The improved photoreduction performance is related to the following features. Firstly, the 0D/1D nanocomposite provided combined optical and electrical properties of both WS2 QDs and Bi2S3 nanotubes causing high visible and near infrared light absorption. Secondly, the enlarged surface area of the nanocomposite provided more active adsorptions site for CO2. Thirdly, the low resistive QDs–NTs interface due to the Bi–S bonds plays a critical role for accelerated charge carrier separation [58]. CsPbBr3 is widely used in the photocatalytic reactions but it suffers from the high rate of recombination during the interface transfer due to the strong reductive ability of electrons [64]. Hence, the suppression of undesired electron loss throughout the transfer process at the interface is critical factor for efficient utilization of CsPbBr3. Li et al. fabricated 0D/2D nanocomposite of CsPbBr3/Bi2WO6 via ultrasonic method with intimate contact at the interface to improve the charge separation and transfer. The Bi–Br bonds which is formed at the QDs–NSs interface is responsible for the strong interfacial interaction. The decoration of Bi2WO6 with CsPbBr3 QDs could enhance the CO and CH4 yield by factor of 9.5 over that of pristine CsPbBr3 [64].
The building of 1D semiconductor materials on 2D semiconductors is an efficient strategy for efficient CO2 conversion. Coupling of TiO2 nanofibers with light harvesting semiconductors such as MoS2 nanosheets is an efficient way to overcome the fast recombination of charge carriers and enhance the light absorption efficiency [65]. The electronic properties of MoS2 nanosheets can be tuned by control of the thickness. The superior conversion activity of CO2 into hydrocarbon species, that is, CH4 and CH3OH, resulted from the improved light harvesting, sufficient reactive sites for CO2 adsorption, and the intimate 1D–2D chemical contact between MoS2 and TiO2 which could be favorable for facile and efficient charge separation upon photoexcitation [65]. The increase of the contact area between the two semiconductor nanomaterials is much more favorable to enhance the photocatalytic performance over the photocatalyst. In other words, constructing the 2D/2D interface is favorable for highly separated charge carriers at the interface. Wang et al. prepared 2D/2D heterojunctions by growth of ultrathin tin disulfide (SnS2) onto TiO2 nanosheets via a hydrothermal method [71]. The production yield of CH4 over SnS2/TiO2 was much higher than that of pristine SnS2 and TiO2 nanosheets. The reason for such outstanding performance originates from the increment of the contract area between SnS2 and TiO2 nanosheets [71].
The photocatalytic systems with multiple junctions, that is, with multiple interfaces, displays excellent photocatalytic activity toward solar fuels generation compared to one with/without single junction [67,144,145]. Recently, Macyk and coworkers designed two heterointerface-based photocatalyst, TiO2/C3N4/Ti3C2, via the interfacial assembly of Ti3C2 QDs on the TiO2/C3N4 binary nanocomposite to boost the charge separation and transfer and providing strong redox ability in CO2 photoreduction reaction [145]. The as-synthesized composite exhibited enhanced light absorption, suppressed electron-hole recombination, and demonstrated stable photocurrent sensitivity. The fabricated composite could overcome the disadvantage of TiO2/C3N4 nanocomposites with a single junction by providing more efficient transport channels of electrons-hole pairs due to the strong interaction between Ti3C2 QDs and TiO2/C3N4 NS. Theoretical studies demonstrated that construction of two interfacial electric fields between TiO2/C3N4 and Ti3C2/C3N4 is due to electron transfer processes at the two interfaces. The interfacial built-in electric fields can promote the charge carrier separation and the photocatalytic reduction of CO2 into CO and CH4. Tonda et al. fabricated another multijunction system with a Bi2WO3/RGO/g-C3N4 2D/2D/2D architecture using two-step hydrothermal method for utilization in CO2 and water reduction into useful fuels [144]. This ternary heterojunction exhibited highly improved characteristics in light harvesting ability, CO2 adsorption capacity, photocurrent responses, and interfacial contact area. The photoconversion performance of CO2 over Bi2WO3/RGO/g-C3N4 was dramatically enhanced toward CH4 and CO evolution. The performance of Bi2WO3/RGO/g-C3N4 was 2.5 times higher and 3.8 times higher than those of Bi2WO3/RGO and RGO/g-C3N4, respectively [144].

4.2. Modification of Low-Dimensional Nanomaterials

Modification of nanostructured low-dimensional photocatalysts themselves are another beneficial strategy for the enhancement of photocatalytic CO2 conversion because it helps the properties of photocatalysts to be engineered. The modification can be achieved using several approaches such as introduction of surface oxygen vacancies or the formation of a porous structure.
The efficient utilization of the solar spectrum can be controlled by tuning the band structure of the semiconductor using the elemental doping [19]. Yu et al. synthesized oxygen-doped g-C3N4 nanotubes via exfoliation and a 3D g-C3N4 curling condensation method [146]. It was found that the synthesized photocatalyst consisted of curled nanosheets that had a uniform tubular structure with 20–30 nm of diameter. The oxygen atoms can substitute for the C or N atoms in g-C3N4 under high temperature oxidation conditions. The oxygen doping of 1D g-C3N4 helped the conduction band to be at a more positive potential causing a narrower band gap and efficient light harvesting. This structure exhibited a significant methanol evolution rate of 0.88 µmol∙g−1∙h−1 under visible light radiation. Wu et al. synthesized self-doped black TiO2 nanotubes arrays using a one-step aluminothermic reduction for solar-driven conversion of CO2 into CO [147]. It is found that the average diameter of the nanotubes was 75–85 nm with 5–7 nm of wall thickness. The oxygen vacancies can act as active sites for CO2 molecules for efficient photogenerated charge carrier separation. The visible light absorption of black TiO2 was largely enhanced by virtue of the oxygen vacancies. The resulted photocatalytic conversion was 185.39 µmol∙g−1∙h−1 of CO evolution rate under visible light.
The introduction of defects into semiconductors can improve the photocatalytic activity of CO2 into solar fuels ascribed to the promotion of photogenerated charge carrier separation and the extended light absorption [148]. Liu and coworkers prepared Bi12O17Cl2 nanotubes with surface oxygen defects via solvothermal method [52]. The tubular structure plays crucial role for accelerating the photogenerated charge carrier separation, while the oxygen defects on the surface act as active centers for CO2 activation. It is found that the absorption of Bi12O17Cl2 nanotubes is improved in the visible region compared to its bulk counterpart. The defective ultrathin tubular structure of Bi12O17Cl2 provides effective CO2 conversion into CO with production yield of 16.8 times higher than bulk Bi12O17Cl2. The higher photocatalytic conversion rate can be attributed to faster charge separation on the surface of Bi12O17Cl2 nanotubes.
The porosity of nanostructured semiconductors provides an additional feature to increase the surface area of photocatalysts, and subsequently is favorable for the solar-driven reduction of CO2 into valuable fuels [149]. Huang et al. used a template-free method to prepare porous g-C3N4 with increased surface area [150]. It is reported that the porous g-C3N4 nanotubes had excellent photocatalytic conversion of CO2 into CO of 40 µmol∙g−1 within 4 h illumination. The CO yield was higher than that of bulk g-C3N4 by a factor of 5.6 originated from the higher surface area of the porous tubular structure, and the improved charge carrier separation and transfer process.

5. Conclusions

The solar-driven conversion of CO2 into valuable fuels such as methane, carbon dioxide, methanol, ethanol, acetaldehyde, and formaldehyde, can be an effective way to simultaneously mitigate the effects of global warming and solve energy crisis all over the world. In this work, following a brief introduction of the essential need to decrease the greenhouse gas effect caused by the excessive emission of CO2, the photocatalytic reduction process of CO2 was discussed including the three basic steps: first, light absorption via the photocatalyst to generate electron-hole pairs; second, separation of the photogenerated charge carriers and transfer to the photocatalyst surface; and third, reaction of the photogenerated electrons and holes with CO2 and water, respectively, to produce solar fuels. The high-efficiency conversion of CO2 into chemical fuels requires a balance between the following consideration of the photocatalyst: firstly, the low band gap of the semiconductor is required for enhanced light absorption provided that the CBM and VBM matching with the redox potentials of the reactants. Secondly, the conversion of CO2 into solar fuels is also dependent on the surface area of the photocatalyst which can facilitate the adsorption of the CO2 molecules on the semiconductor surface. Thirdly, the charge dynamics process, namely, e-h separation/transfer, must be controlled to minimize the fast recombination of the photogenerated charge carriers. In this regard, several strategies have been reported to improve the CO2 conversion including elemental doping, heterojunction construction, surface modification, nanostrucutralization, and dimensionality reduction. A review was presented of the advantages of nanostructured low-dimensional photocatalysts, including the higher surface area, the suppressed recombination rate due to its crystallinity, the quantum confinement effect in semiconductors, and plasmonic effect on the surface of metals nanoparticles. Further, the state-of-the-art progress of 0D, 1D, and 2D nanomaterials for CO2 reduction into solar fuels was presented. Finally, we report on two strategies to boost the photocatalytic activity of CO2 into valuable fuels: first, designing junction structures to enhance the photogenerated charge carrier separation and transfer to the surface and second the modification of low-dimensional photocatalysts using doping or porosity to enhance the light absorption and surface area. By tuning the dimensionality of nanostructured semiconductors, we are able to improve the optical and electronic properties, allowing significant enhancement of the CO2 conversion into solar fuels. Ultimately, for solar-driven conversion of CO2 into usable fuel to make an impact on society, efficient processes will also have to be inexpensive to fabricate and easily scalable. Engineered low-dimensional materials photocatalytic materials show great promise toward this end.

Author Contributions

Conceptualization, supervision, project administration, funding acquisition, H.L.; writing-original draft preparation, H.L., H.A.E.O., M.W.H.; writing-review and editing, H.L., M.W.H.; visualization, H.A.E.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology of Taiwan, grant no. 109-2218-E-110-005-MY2.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The authors acknowledge the financial support of the Ministry of Science and Technology of Taiwan, grant no. 109-2218-E-110-005-MY2.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Halmann, M. Photoelectrochemical reduction of aqueous carbon dioxide on p-type gallium phosphide in liquid junction solar cells. Nature 1978, 275, 115–116. [Google Scholar] [CrossRef]
  2. Hemminger, J.; Carr, R.; Somorjai, G. The photoassisted reaction of gaseous water and carbon dioxide adsorbed on the SrTiO3 (111) crystal face to form methane. Chem. Phys. Lett. 1978, 57, 100–104. [Google Scholar] [CrossRef]
  3. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photoelectrocatalytic reduction of carbon dioxide in aqueous suspensions of semiconductor powders. Nature 1979, 277, 637–638. [Google Scholar] [CrossRef]
  4. Zeng, S.; Kar, P.; Thakur, U.K.; Shankar, K. A review on photocatalytic CO2 reduction using perovskite oxide nanomaterials. Nanotechnology 2018, 29, 052001. [Google Scholar] [CrossRef]
  5. Zhou, B.; Song, J.; Xie, C.; Chen, C.; Qian, Q.; Han, B. Mo–Bi–Cd Ternary Metal Chalcogenides: Highly Efficient Photocatalyst for CO2 Reduction to Formic Acid Under Visible Light. ACS Sustain. Chem. Eng. 2018, 6, 5754–5759. [Google Scholar] [CrossRef]
  6. Bie, C.; Zhu, B.; Xu, F.; Zhang, L.; Yu, J. In situ grown monolayer N-doped graphene on CdS hollow spheres with seamless contact for photocatalytic CO2 reduction. Adv. Mater. 2019, 31, 1902868. [Google Scholar] [CrossRef] [PubMed]
  7. Billo, T.; Shown, I.; Kumar Anbalagan, A.; Effendi, T.A.; Sabbah, A.; Fu, F.-Y.; Chu, C.-M.; Woon, W.-Y.; Chen, R.-S.; Lee, C.-H. A mechanistic study of molecular CO2 interaction and adsorption on carbon implanted SnS2 thin film for photocatalytic CO2 reduction activity. Nano Energy 2020, 72, 104717. [Google Scholar] [CrossRef]
  8. 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]
  9. Liu, W.; Li, X.; Wang, C.; Pan, H.; Liu, W.; Wang, K.; Zeng, Q.; Wang, R.; Jiang, J. A scalable general synthetic approach toward ultrathin imine-linked two-dimensional covalent organic framework nanosheets for photocatalytic CO2 reduction. J. Am. Chem. Soc. 2019, 141, 17431–17440. [Google Scholar] [CrossRef]
  10. Yang, C.; Huang, W.; da Silva, L.C.; Zhang, K.A.; Wang, X. Functional conjugated polymers for CO2 reduction using visible light. Chem. Eur. J. 2018, 24, 17454–17458. [Google Scholar] [CrossRef] [PubMed]
  11. Yan, Y.; Chen, J.; Li, N.; Tian, J.; Li, K.; Jiang, J.; Liu, J.; Tian, Q.; Chen, P. Systematic bandgap engineering of graphene quantum dots and applications for photocatalytic water splitting and CO2 reduction. ACS Nano 2018, 12, 3523–3532. [Google Scholar] [CrossRef] [PubMed]
  12. Sun, Z.; Wang, H.; Wu, Z.; Wang, L. g-C3N4 based composite photocatalysts for photocatalytic CO2 reduction. Catal. Today 2018, 300, 160–172. [Google Scholar] [CrossRef]
  13. Gao, W.; Bai, X.; Gao, Y.; Liu, J.; He, H.; Yang, Y.; Han, Q.; Wang, X.; Wu, X.; Wang, J. Anchoring of black phosphorus quantum dots onto WO3 nanowires to boost photocatalytic CO2 conversion into solar fuels. Chem. Comm. 2020, 56, 7777–7780. [Google Scholar] [CrossRef]
  14. Ren, X.; Gao, M.; Zhang, Y.; Zhang, Z.; Cao, X.; Wang, B.; Wang, X. Photocatalytic reduction of CO2 on BiOX: Effect of halogen element type and surface oxygen vacancy mediated mechanism. Appl. Catal. B 2020, 274, 119063. [Google Scholar] [CrossRef]
  15. Wang, H.; Zhang, L.; Wang, K.; Sun, X.; Wang, W. Enhanced photocatalytic CO2 reduction to methane over WO3· 0.33 H2O via Mo doping. Appl. Catal. B 2019, 243, 771–779. [Google Scholar] [CrossRef]
  16. Gao, M.; Yang, J.; Sun, T.; Zhang, Z.; Zhang, D.; Huang, H.; Lin, H.; Fang, Y.; Wang, X. Persian buttercup-like BiOBrxCl1-x solid solution for photocatalytic overall CO2 reduction to CO and O2. Appl. Catal. B 2019, 243, 734–740. [Google Scholar] [CrossRef]
  17. Samanta, S.; Yadav, R.; Kumar, A.; Sinha, A.K.; Srivastava, R. Surface modified C, O co-doped polymeric g-C3N4 as an efficient photocatalyst for visible light assisted CO2 reduction and H2O2 production. Appl. Catal. B 2019, 259, 118054. [Google Scholar] [CrossRef]
  18. Shi, R.; Chen, Y. Controlled formation of defective shell on TiO2 (001) facets for enhanced photocatalytic CO2 reduction. ChemCatChem 2019, 11, 2270–2276. [Google Scholar] [CrossRef]
  19. Tu, W.; Zhou, Y.; Zou, Z. Photocatalytic conversion of CO2 into renewable hydrocarbon fuels: State-of-the-art accomplishment, challenges, and prospects. Adv. Mater. 2014, 26, 4607–4626. [Google Scholar] [CrossRef]
  20. Razzaq, A.; In, S.-I. TiO2 based nanostructures for photocatalytic CO2 conversion to valuable chemicals. Micromachines 2019, 10, 326. [Google Scholar] [CrossRef] [Green Version]
  21. Huang, C.-Y.; Guo, R.-T.; Pan, W.-G.; Tang, J.-Y.; Zhou, W.-G.; Liu, X.-Y.; Qin, H.; Jia, P.-Y. One-dimension TiO2 nanostructures with enhanced activity for CO2 photocatalytic reduction. Appl. Surf. Sci. 2019, 464, 534–543. [Google Scholar] [CrossRef]
  22. Patil, S.B.; Basavarajappa, P.S.; Ganganagappa, N.; Jyothi, M.; Raghu, A.; Reddy, K.R. Recent advances in non-metals-doped TiO2 nanostructured photocatalysts for visible-light driven hydrogen production, CO2 reduction and air purification. Int. J. Hydrog. Energy 2019, 44, 13022–13039. [Google Scholar] [CrossRef]
  23. Xu, H.-M.; Wang, H.-C.; Shen, Y.; Lin, Y.-H.; Nan, C.-W. Low-dimensional nanostructured photocatalysts. J. Adv. Ceram. 2015, 4, 159–182. [Google Scholar] [CrossRef] [Green Version]
  24. 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]
  25. Mao, J.; Li, K.; Peng, T. Recent advances in the photocatalytic CO2 reduction over semiconductors. Catal. Sci. Technol. 2013, 3, 2481–2498. [Google Scholar] [CrossRef]
  26. Markovits, A.; Fahmi, A.; Minot, C. A theoretical study of CO2 adsorption on TiO2. J. Mol. Struct. THEOCHEM 1996, 371, 219–235. [Google Scholar] [CrossRef]
  27. Sharma, N.; Das, T.; Kumar, S.; Bhosale, R.; Kabir, M.; Ogale, S. Photocatalytic activation and reduction of CO2 to CH4 over single phase nano Cu3SnS4: A combined experimental and theoretical study. ACS Appl. Energy Mater. 2019, 2, 5677–5685. [Google Scholar] [CrossRef]
  28. Indrakanti, V.P.; Kubicki, J.D.; Schobert, H.H. Photoinduced activation of CO2 on Ti-based heterogeneous catalysts: Current state, chemical physics-based insights and outlook. Energy Environ. Sci. 2009, 2, 745–758. [Google Scholar] [CrossRef]
  29. Álvarez, A.; Borges, M.; Corral-Pérez, J.J.; Olcina, J.G.; Hu, L.; Cornu, D.; Huang, R.; Stoian, D.; Urakawa, A. CO2 activation over catalytic surfaces. ChemPhysChem 2017, 18, 3135–3141. [Google Scholar] [CrossRef] [Green Version]
  30. Fu, J.; Jiang, K.; Qiu, X.; Yu, J.; Liu, M. Product selectivity of photocatalytic CO2 reduction reactions. Mater. Today 2020, 32, 222–243. [Google Scholar] [CrossRef]
  31. Linsebigler, A.L.; Lu, G.; Yates, J.T., Jr. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  32. Khan, A.A.; Tahir, M. Recent advancements in engineering approach towards design of photo-reactors for selective photocatalytic CO2 reduction to renewable fuels. J. CO2 Util. 2019, 29, 205–239. [Google Scholar] [CrossRef]
  33. Sun, Z.; Talreja, N.; Tao, H.; Texter, J.; Muhler, M.; Strunk, J.; Chen, J. Catalysis of carbon dioxide photoreduction on nanosheets: Fundamentals and challenges. Angew. Chem. Int. Ed. 2018, 57, 7610–7627. [Google Scholar] [CrossRef] [PubMed]
  34. Yoffe, A.D. Low-dimensional systems: Quantum size effects and electronic properties of semiconductor microcrystallites (zero-dimensional systems) and some quasi-two-dimensional systems. Adv. Phys. 1993, 42, 173–262. [Google Scholar] [CrossRef]
  35. Lim, Y.; Lee, D.-K.; Kim, S.M.; Park, W.; Cho, S.Y.; Sim, U. Low dimensional carbon-based catalysts for efficient photocatalytic and photo/electrochemical water splitting reactions. Materials 2020, 13, 114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ren, X.; Philo, D.; Li, Y.; Shi, L.; Chang, K.; Ye, J. Recent advances of low-dimensional phosphorus-based nanomaterials for solar-driven photocatalytic reactions. Coord. Chem. Rev. 2020, 424, 213516. [Google Scholar] [CrossRef]
  37. Peng, W.; Li, Y.; Zhang, F.; Zhang, G.; Fan, X. Roles of two-dimensional transition metal dichalcogenides as cocatalysts in photocatalytic hydrogen evolution and environmental remediation. Ind. Eng. Chem. Res. 2017, 56, 4611–4626. [Google Scholar] [CrossRef]
  38. Chowdhury, C.; Karmakar, S.; Datta, A. Monolayer group IV–VI monochalcogenides: Low-dimensional materials for photocatalytic water splitting. J. Phys. Chem. C 2017, 121, 7615–7624. [Google Scholar] [CrossRef]
  39. Dai, C.; Liu, B. Conjugated polymers for visible-light-driven photocatalysis. Energy Environ. Sci. 2020, 13, 24–52. [Google Scholar] [CrossRef]
  40. Huynh, K.A.; Nguyen, D.L.T.; Nguyen, V.H.; Vo, D.V.N.; Trinh, Q.T.; Nguyen, T.P.; Kim, S.Y.; Le, Q.V. Halide perovskite photocatalysis: Progress and perspectives. J. Chem. Technol. Biotechnol. 2020, 95, 2579–2596. [Google Scholar] [CrossRef]
  41. Khan, M.S.; Zhang, F.; Osada, M.; Mao, S.S.; Shen, S. Graphitic Carbon Nitride-Based Low-Dimensional Heterostructures for Photocatalytic Applications. Sol. RRL 2020, 4, 1900435. [Google Scholar] [CrossRef]
  42. Zhao, J.; Holmes, M.A.; Osterloh, F.E. Quantum confinement controls photocatalysis: A free energy analysis for photocatalytic proton reduction at CdSe nanocrystals. ACS Nano 2013, 7, 4316–4325. [Google Scholar] [CrossRef]
  43. Qamar, S.; Lei, F.; Liang, L.; Gao, S.; Liu, K.; Sun, Y.; Ni, W.; Xie, Y. Ultrathin TiO2 flakes optimizing solar light driven CO2 reduction. Nano Energy 2016, 26, 692–698. [Google Scholar] [CrossRef]
  44. Li, A.; Wang, T.; Li, C.; Huang, Z.; Luo, Z.; Gong, J. Adjusting the reduction potential of electrons by quantum confinement for selective photoreduction of CO2 to methanol. Angew. Chem. Int. Ed. 2019, 58, 3804–3808. [Google Scholar] [CrossRef] [PubMed]
  45. Hou, J.; Cao, S.; Wu, Y.; Gao, Z.; Liang, F.; Sun, Y.; Lin, Z.; Sun, L. Inorganic colloidal perovskite quantum dots for robust solar CO2 reduction. Chem. Eur. J. 2017, 23, 9481–9485. [Google Scholar] [CrossRef] [PubMed]
  46. Sheng, J.; He, Y.; Li, J.; Yuan, C.; Huang, H.; Wang, S.; Sun, Y.; Wang, Z.; Dong, F. Identification of halogen-associated active sites on bismuth-based perovskite quantum dots for efficient and selective CO2-to-CO photoreduction. ACS Nano 2020, 14, 13103–13114. [Google Scholar] [CrossRef]
  47. Zhou, L.; Xu, Y.F.; Chen, B.X.; Kuang, D.B.; Su, C.Y. Synthesis and Photocatalytic Application of Stable Lead-Free Cs2AgBiBr6 Perovskite Nanocrystals. Small 2018, 14, 1703762. [Google Scholar] [CrossRef] [PubMed]
  48. Que, M.; Zhao, Y.; Pan, L.; Yang, Y.; He, Z.; Yuan, H.; Chen, J.; Zhu, G. Colloidal formamidinium lead bromide quantum dots for photocatalytic CO2 reduction. Mater. Lett. 2021, 282, 128695. [Google Scholar] [CrossRef]
  49. Zhu, C.; Wei, X.; Li, W.; Pu, Y.; Sun, J.; Tang, K.; Wan, H.; Ge, C.; Zou, W.; Dong, L. Crystal-Plane Effects of CeO2 {110} and CeO2 {100} on Photocatalytic CO2 Reduction: Synergistic Interactions of Oxygen Defects and Hydroxyl Groups. ACS Sustain. Chem. Eng. 2020, 8, 14397–14406. [Google Scholar] [CrossRef]
  50. Liu, B.; Ye, L.; Wang, R.; Yang, J.; Zhang, Y.; Guan, R.; Tian, L.; Chen, X. Phosphorus-doped graphitic carbon nitride nanotubes with amino-rich surface for efficient CO2 capture, enhanced photocatalytic activity, and product selectivity. ACS Appl. Mater. Interfaces 2018, 10, 4001–4009. [Google Scholar] [CrossRef]
  51. Mo, Z.; Zhu, X.; Jiang, Z.; Song, Y.; Liu, D.; Li, H.; Yang, X.; She, Y.; Lei, Y.; Yuan, S. Porous nitrogen-rich g-C3N4 nanotubes for efficient photocatalytic CO2 reduction. Appl. Catal. B 2019, 256, 117854. [Google Scholar] [CrossRef]
  52. Di, J.; Zhu, C.; Ji, M.; Duan, M.; Long, R.; Yan, C.; Gu, K.; Xiong, J.; She, Y.; Xia, J. Defect-Rich Bi12O17Cl2 Nanotubes Self-Accelerating Charge Separation for Boosting Photocatalytic CO2 Reduction. Angew. Chem. Int. Ed. 2018, 57, 14847–14851. [Google Scholar] [CrossRef]
  53. Di, J.; Song, P.; Zhu, C.; Chen, C.; Xiong, J.; Duan, M.; Long, R.; Zhou, W.; Xu, M.; Kang, L. Strain-Engineering of Bi12O17Br2 Nanotubes for Boosting Photocatalytic CO2 Reduction. ACS Mater. Lett. 2020, 2, 1025–1032. [Google Scholar] [CrossRef]
  54. Xu, G.; Zhang, H.; Wei, J.; Zhang, H.-X.; Wu, X.; Li, Y.; Li, C.; Zhang, J.; Ye, J. Integrating the g-C3N4 Nanosheet with B–H bonding decorated metal–organic framework for CO2 activation and photoreduction. ACS Nano 2018, 12, 5333–5340. [Google Scholar] [CrossRef] [PubMed]
  55. Xia, P.; Zhu, B.; Yu, J.; Cao, S.; Jaroniec, M. Ultra-thin nanosheet assemblies of graphitic carbon nitride for enhanced photocatalytic CO2 reduction. J. Mater. Chem. A 2017, 5, 3230–3238. [Google Scholar] [CrossRef]
  56. Han, C.; Wang, B.; Wu, C.; Shen, S.; Zhang, X.; Sun, L.; Tian, Q.; Lei, Y.; Wang, Y. Ultrathin SiC nanosheets with high reduction potential for improved CH4 generation from photocatalytic reduction of CO2. ChemistrySelect 2019, 4, 2211–2217. [Google Scholar] [CrossRef]
  57. Di, J.; Zhao, X.; Lian, C.; Ji, M.; Xia, J.; Xiong, J.; Zhou, W.; Cao, X.; She, Y.; Liu, H. Atomically-thin Bi2MoO6 nanosheets with vacancy pairs for improved photocatalytic CO2 reduction. Nano Energy 2019, 61, 54–59. [Google Scholar] [CrossRef]
  58. 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]
  59. Zeng, Z.; Yan, Y.; Chen, J.; Zan, P.; Tian, Q.; Chen, P. Boosting the photocatalytic ability of Cu2O nanowires for CO2 conversion by MXene quantum dots. Adv. Funct. Mater. 2019, 29, 1806500. [Google Scholar] [CrossRef]
  60. Xie, Z.; Xu, Y.; Li, D.; Chen, L.; Meng, S.; Jiang, D.; Chen, M. Construction of CuO quantum Dots/WO3 nanosheets 0D/2D Z-scheme heterojunction with enhanced photocatalytic CO2 reduction activity under visible-light. J. Alloys Compd. 2021, 858, 157668. [Google Scholar] [CrossRef]
  61. Ong, W.-J.; Putri, L.K.; Tan, Y.-C.; Tan, L.-L.; Li, N.; Ng, Y.H.; Wen, X.; Chai, S.-P. Unravelling charge carrier dynamics in protonated g-C3N4 interfaced with carbon nanodots as co-catalysts toward enhanced photocatalytic CO2 reduction: A combined experimental and first-principles DFT study. Nano Res. 2017, 10, 1673–1696. [Google Scholar] [CrossRef]
  62. Shi, H.; Long, S.; Hu, S.; Hou, J.; Ni, W.; Song, C.; Li, K.; Gurzadyan, G.G.; Guo, X. Interfacial charge transfer in 0D/2D defect-rich heterostructures for efficient solar-driven CO2 reduction. Appl. Catal. B 2019, 245, 760–769. [Google Scholar] [CrossRef]
  63. Wang, R.; Shen, J.; Sun, K.; Tang, H.; Liu, Q. Enhancement in photocatalytic activity of CO2 reduction to CH4 by 0D/2D Au/TiO2 plasmon heterojunction. Appl. Surf. Sci. 2019, 493, 1142–1149. [Google Scholar] [CrossRef]
  64. Wang, J.; Wang, J.; Li, N.; Du, X.; Ma, J.; He, C.; Li, Z. Direct Z-Scheme 0D/2D Heterojunction of CsPbBr3 Quantum Dots/Bi2WO6 Nanosheets for Efficient Photocatalytic CO2 Reduction. ACS Appl. Mater. Interfaces 2020, 12, 31477–31485. [Google Scholar] [CrossRef]
  65. Xu, F.; Zhu, B.; Cheng, B.; Yu, J.; Xu, J. 1D/2D TiO2/MoS2 hybrid nanostructures for enhanced photocatalytic CO2 reduction. Adv. Opt. Mater. 2018, 6, 1800911. [Google Scholar] [CrossRef] [Green Version]
  66. Tahir, M.; Tahir, B.; Nawawi, M.; Hussain, M.; Muhammad, A. Cu-NPs embedded 1D/2D CNTs/pCN heterojunction composite towards enhanced and continuous photocatalytic CO2 reduction to fuels. Appl. Surf. Sci. 2019, 485, 450–461. [Google Scholar] [CrossRef]
  67. Bian, J.; Qu, Y.; Zhang, X.; Sun, N.; Tang, D.; Jing, L. Dimension-matched plasmonic Au/TiO2/BiVO4 nanocomposites as efficient wide-visible-light photocatalysts to convert CO2 and mechanistic insights. J. Mater. Chem. A 2018, 6, 11838–11845. [Google Scholar] [CrossRef]
  68. Cao, S.; Shen, B.; Tong, T.; Fu, J.; Yu, J. 2D/2D heterojunction of ultrathin MXene/Bi2WO6 nanosheets for improved photocatalytic CO2 reduction. Adv. Funct. Mater. 2018, 28, 1800136. [Google Scholar] [CrossRef]
  69. Kong, X.Y.; Lee, W.Q.; Mohamed, A.R.; Chai, S.-P. Effective steering of charge flow through synergistic inducing oxygen vacancy defects and pn heterojunctions in 2D/2D surface-engineered Bi2WO6/BiOI cascade: Towards superior photocatalytic CO2 reduction activity. Chem. Eng. J. 2019, 372, 1183–1193. [Google Scholar] [CrossRef]
  70. Tahir, B.; Tahir, M.; Yunus, M.A.C.; Mohamed, A.R.; Siraj, M.; Fatehmulla, A. 2D/2D Mt/m-CN composite with enriched interface charge transfer for boosting photocatalytic CO2 hydrogenation by H2 to CH4 under visible light. Appl. Surf. Sci. 2020, 520, 146296. [Google Scholar] [CrossRef]
  71. She, H.; Zhou, H.; Li, L.; Zhao, Z.; Jiang, M.; Huang, J.; Wang, L.; Wang, Q. Construction of a two-dimensional composite derived from TiO2 and SnS2 for enhanced photocatalytic reduction of CO2 into CH4. ACS Sustain. Chem. Eng. 2018, 7, 650–659. [Google Scholar] [CrossRef]
  72. Lu, M.; Li, Q.; Zhang, C.; Fan, X.; Li, L.; Dong, Y.; Chen, G.; Shi, H. Remarkable photocatalytic activity enhancement of CO2 conversion over 2D/2D g-C3N4/BiVO4 Z-scheme heterojunction promoted by efficient interfacial charge transfer. Carbon 2020, 160, 342–352. [Google Scholar] [CrossRef]
  73. Huo, Y.; Zhang, J.; Wang, Z.; Dai, K.; Pan, C.; Liang, C. Efficient interfacial charge transfer of 2D/2D porous carbon nitride/bismuth oxychloride step-scheme heterojunction for boosted solar-driven CO2 reduction. J. Colloid Interface Sci. 2021, 585, 684–693. [Google Scholar] [CrossRef]
  74. Zhang, R.; Huang, Z.; Li, C.; Zuo, Y.; Zhou, Y. Monolithic g-C3N4/reduced graphene oxide aerogel with in situ embedding of Pd nanoparticles for hydrogenation of CO2 to CH4. Appl. Surf. Sci. 2019, 475, 953–960. [Google Scholar] [CrossRef]
  75. Wang, J.; Xia, T.; Wang, L.; Zheng, X.; Qi, Z.; Gao, C.; Zhu, J.; Li, Z.; Xu, H.; Xiong, Y. Enabling Visible-Light-Driven Selective CO2 Reduction by Doping Quantum Dots: Trapping Electrons and Suppressing H2 Evolution. Angew. Chem. Int. Ed. 2018, 57, 16447–16451. [Google Scholar] [CrossRef]
  76. Wang, C.; Thompson, R.L.; Ohodnicki, P.; Baltrus, J.; Matranga, C. Size-dependent photocatalytic reduction of CO2 with PbS quantum dot sensitized TiO2 heterostructured photocatalysts. J. Mater. Chem. 2011, 21, 13452–13457. [Google Scholar] [CrossRef]
  77. Kuehnel, M.F.; Sahm, C.D.; Neri, G.; Lee, J.R.; Orchard, K.L.; Cowan, A.J.; Reisner, E. ZnSe quantum dots modified with a Ni (cyclam) catalyst for efficient visible-light driven CO2 reduction in water. Chem. Sci. 2018, 9, 2501–2509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Huang, J.; Gatty, M.G.; Xu, B.; Pati, P.B.; Etman, A.S.; Tian, L.; Sun, J.; Hammarström, L.; Tian, H. Covalently linking CuInS2 quantum dots with a Re catalyst by click reaction for photocatalytic CO2 reduction. Dalton Trans. 2018, 47, 10775–10783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Chen, L.; Zhang, M.; Yang, J.; Li, Y.; Sivalingam, Y.; Shi, Q.; Xie, M.; Han, W. Synthesis of BiVO4 quantum dots/reduced graphene oxide composites for CO2 reduction. Mater. Sci. Semicond. Process. 2019, 102, 104578. [Google Scholar] [CrossRef]
  80. Qin, H.; Guo, R.-T.; Liu, X.-Y.; Shi, X.; Wang, Z.-Y.; Tang, J.-Y.; Pan, W.-G. 0D NiS2 quantum dots modified 2D g-C3N4 for efficient photocatalytic CO2 reduction. Colloids Surf. A Physicochem. Eng. Asp. 2020, 600, 124912. [Google Scholar] [CrossRef]
  81. Yadav, D.; Yadav, R.K.; Kumar, A.; Park, N.J.; Baeg, J.O. Functionalized Graphene Quantum Dots as Efficient Visible-Light Photocatalysts for Selective Solar Fuel Production from CO2. ChemCatChem 2016, 8, 3389–3393. [Google Scholar] [CrossRef]
  82. Han, C.; Li, J.; Ma, Z.; Xie, H.; Waterhouse, G.I.; Ye, L.; Zhang, T. Black phosphorus quantum dot/g-C3N4 composites for enhanced CO2 photoreduction to CO. Sci. China Mater. 2018, 61, 1159–1166. [Google Scholar] [CrossRef] [Green Version]
  83. Guo, R.-T.; Liu, X.-Y.; Qin, H.; Wang, Z.-Y.; Shi, X.; Pan, W.-G.; Fu, Z.-G.; Tang, J.-Y.; Jia, P.-Y.; Miao, Y.-F. Photocatalytic reduction of CO2 into CO over nanostructure Bi2S3 quantum dots/g-C3N4 composites with Z-scheme mechanism. Appl. Surf. Sci. 2020, 500, 144059. [Google Scholar] [CrossRef]
  84. Sharma, S.; Dutta, V.; Singh, P.; Raizada, P.; Rahmani-Sani, A.; Hosseini-Bandegharaei, A.; Thakur, V.K. Carbon quantum dot supported semiconductor photocatalysts for efficient degradation of organic pollutants in water: A review. J. Clean. Prod. 2019, 228, 755–769. [Google Scholar] [CrossRef]
  85. Fernando, K.S.; Sahu, S.; Liu, Y.; Lewis, W.K.; Guliants, E.A.; Jafariyan, A.; Wang, P.; Bunker, C.E.; Sun, Y.-P. Carbon quantum dots and applications in photocatalytic energy conversion. ACS Appl. Mater. Interfaces 2015, 7, 8363–8376. [Google Scholar] [CrossRef] [PubMed]
  86. Wang, X.; Feng, Y.; Dong, P.; Huang, J. A mini review on carbon quantum dots: Preparation, properties, and electrocatalytic application. Front. Chem. 2019, 7, 671. [Google Scholar] [CrossRef]
  87. Li, Q.; Wang, S.; Sun, Z.; Tang, Q.; Liu, Y.; Wang, L.; Wang, H.; Wu, Z. Enhanced CH4 selectivity in CO2 photocatalytic reduction over carbon quantum dots decorated and oxygen doping g-C3N4. Nano Res. 2019, 12, 2749–2759. [Google Scholar] [CrossRef]
  88. Zeng, Z.; Chen, S.; Tan, T.T.Y.; Xiao, F.-X. Graphene quantum dots (GQDs) and its derivatives for multifarious photocatalysis and photoelectrocatalysis. Catal. Today 2018, 315, 171–183. [Google Scholar] [CrossRef]
  89. Kim, S.Y.; Jang, H.W. Lead-free all-inorganic halide perovskite quantum dots: Review and outlook. J. Korean Ceram. Soc. 2020, 57, 455–479. [Google Scholar]
  90. Moon, H.; Lee, C.; Lee, W.; Kim, J.; Chae, H. Stability of quantum dots, quantum dot films, and quantum dot light-emitting diodes for display applications. Adv. Mater. 2019, 31, 1804294. [Google Scholar] [CrossRef]
  91. Lee, H.; Horn, M.W. Sculptured platinum nanowire counter electrodes for dye-sensitized solar cells. Thin Solid Film. 2013, 540, 208–211. [Google Scholar] [CrossRef]
  92. Lee, H.; In, S.; Horn, M.W. Plasmonic enhancement of CO2 conversion to methane using sculptured copper thin films grown directly on TiO2. Thin Solid Films 2014, 565, 105–110. [Google Scholar] [CrossRef]
  93. Xiao, F.X.; Miao, J.; Tao, H.B.; Hung, S.F.; Wang, H.Y.; Yang, H.B.; Chen, J.; Chen, R.; Liu, B. One-dimensional hybrid nanostructures for heterogeneous photocatalysis and photoelectrocatalysis. Small 2015, 11, 2115–2131. [Google Scholar] [CrossRef] [PubMed]
  94. Sui, Y.; Su, C.; Yang, X.; Hu, J.; Lin, X. Ag-AgBr nanoparticles loaded on TiO2 nanofibers as an efficient heterostructured photocatalyst driven by visible light. J. Mol. Catal. A Chem. 2015, 410, 226–234. [Google Scholar] [CrossRef]
  95. Wang, M.; Pang, X.; Zheng, D.; He, Y.; Sun, L.; Lin, C.; Lin, Z. Nonepitaxial growth of uniform and precisely size-tunable core/shell nanoparticles and their enhanced plasmon-driven photocatalysis. J. Mater. Chem. A 2016, 4, 7190–7199. [Google Scholar] [CrossRef]
  96. Nasr, M.; Eid, C.; Habchi, R.; Miele, P.; Bechelany, M. Recent progress on titanium dioxide nanomaterials for photocatalytic applications. ChemSusChem 2018, 11, 3023–3047. [Google Scholar] [CrossRef] [PubMed]
  97. Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  98. Fujishima, A.; Zhang, X.; Tryk, D.A. TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008, 63, 515–582. [Google Scholar] [CrossRef]
  99. Kar, P.; Zeng, S.; Zhang, Y.; Vahidzadeh, E.; Manuel, A.; Kisslinger, R.; Alam, K.M.; Thakur, U.K.; Mahdi, N.; Kumar, P. High rate CO2 photoreduction using flame annealed TiO2 nanotubes. Appl. Catal. B 2019, 243, 522–536. [Google Scholar] [CrossRef]
  100. Wu, J.; Feng, Y.; Li, D.; Han, X.; Liu, J. Efficient photocatalytic CO2 reduction by P–O linked g-C3N4/TiO2-nanotubes Z-scheme composites. Energy 2019, 178, 168–175. [Google Scholar] [CrossRef]
  101. Rambabu, Y.; Kumar, U.; Singhal, N.; Kaushal, M.; Jaiswal, M.; Jain, S.L.; Roy, S.C. Photocatalytic reduction of carbon dioxide using graphene oxide wrapped TiO2 nanotubes. Appl. Surf. Sci. 2019, 485, 48–55. [Google Scholar] [CrossRef]
  102. Zhao, H.; Chen, J.; Rao, G.; Deng, W.; Li, Y. Enhancing photocatalytic CO2 reduction by coating an ultrathin Al2O3 layer on oxygen deficient TiO2 nanorods through atomic layer deposition. Appl. Surf. Sci. 2017, 404, 49–56. [Google Scholar] [CrossRef] [Green Version]
  103. Wang, X.; Zhang, Z.; Huang, Z.; Dong, P.; Nie, X.; Jin, Z.; Zhang, X. Synergistic effect of N-Ho on photocatalytic CO2 reduction for N/Ho co-doped TiO2 nanorods. Mater. Res. Bull. 2019, 118, 110502. [Google Scholar] [CrossRef]
  104. Ohno, T.; Higo, T.; Murakami, N.; Saito, H.; Zhang, Q.; Yang, Y.; Tsubota, T. Photocatalytic reduction of CO2 over exposed-crystal-face-controlled TiO2 nanorod having a brookite phase with co-catalyst loading. Appl. Catal. B 2014, 152, 309–316. [Google Scholar] [CrossRef] [Green Version]
  105. Tahir, M.; Tahir, B.; Amin, N.A.S. Gold-nanoparticle-modified TiO2 nanowires for plasmon-enhanced photocatalytic CO2 reduction with H2 under visible light irradiation. Appl. Surf. Sci. 2015, 356, 1289–1299. [Google Scholar] [CrossRef]
  106. Shen, M.; Zhang, L.; Shi, J. Converting CO2 into fuels by graphitic carbon nitride-based photocatalysts. Nanotechnology 2018, 29, 412001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Akhundi, A.; Habibi-Yangjeh, A.; Abitorabi, M.; Rahim Pouran, S. Review on photocatalytic conversion of carbon dioxide to value-added compounds and renewable fuels by graphitic carbon nitride-based photocatalysts. Catal. Rev. 2019, 61, 595–628. [Google Scholar] [CrossRef]
  108. Bakar, S.A.; Ribeiro, C. Nitrogen-doped titanium dioxide: An overview of material design and dimensionality effect over modern applications. J. Photochem. Photobiol. C 2016, 27, 1–29. [Google Scholar] [CrossRef]
  109. Narita, A.; Wang, X.-Y.; Feng, X.; Müllen, K. New advances in nanographene chemistry. Chem. Soc. Rev. 2015, 44, 6616–6643. [Google Scholar] [CrossRef] [Green Version]
  110. Khan, B.; Raziq, F.; Faheem, M.B.; Farooq, M.U.; Hussain, S.; Ali, F.; Ullah, A.; Mavlonov, A.; Zhao, Y.; Liu, Z. Electronic and nanostructure engineering of bifunctional MoS2 towards exceptional visible-light photocatalytic CO2 reduction and pollutant degradation. J. Hazard. Mater. 2020, 381, 120972. [Google Scholar] [CrossRef]
  111. Wang, X.; He, J.; Mao, L.; Cai, X.; Sun, C.; Zhu, M. CsPbBr3 perovskite nanocrystals anchoring on monolayer MoS2 nanosheets for efficient photocatalytic CO2 reduction. Chem. Eng. J. 2020, 128077. [Google Scholar] [CrossRef]
  112. Wang, J.; Li, L.; Wang, L.; Liu, Y.; Sun, W.; Li, W.; Li, G. In situ growth of MoS2 nanosheet arrays and TS2 (T = Fe, Co, and Ni) nanocubes onto molybdate for efficient oxygen evolution reaction and improved hydrogen evolution reaction. ACS Omega 2018, 3, 464–471. [Google Scholar] [CrossRef]
  113. He, J.; Chen, L.; Yi, Z.-Q.; Au, C.-T.; Yin, S.-F. CdS nanorods coupled with WS2 nanosheets for enhanced photocatalytic hydrogen evolution activity. Ind. Eng. Chem. Res. 2016, 55, 8327–8333. [Google Scholar] [CrossRef]
  114. Cao, Y.; Zhang, R.; Zhou, T.; Jin, S.; Huang, J.; Ye, L.; Huang, Z.; Wang, F.; Zhou, Y. B–O Bonds in Ultrathin Boron Nitride Nanosheets to Promote Photocatalytic Carbon Dioxide Conversion. ACS Appl. Mater. Interfaces 2020, 12, 9935–9943. [Google Scholar] [CrossRef] [PubMed]
  115. Ali, W.; Zhang, X.; Zhang, X.; Ali, S.; Zhao, L.; Shaheen, S.; Jing, L. Improved visible-light activities of g-C3N4 nanosheets by co-modifying nano-sized SnO2 and Ag for CO2 reduction and 2, 4-dichlorophenol degradation. Mater. Res. Bull. 2020, 122, 110676. [Google Scholar] [CrossRef]
  116. Wang, X.; He, J.; Li, J.; Lu, G.; Dong, F.; Majima, T.; Zhu, M. Immobilizing perovskite CsPbBr3 nanocrystals on Black phosphorus nanosheets for boosting charge separation and photocatalytic CO2 reduction. Appl. Catal. B 2020, 277, 119230. [Google Scholar] [CrossRef]
  117. Li, J.; Liu, P.; Huang, H.; Li, Y.; Tang, Y.; Mei, D.; Zhong, C. Metal-free 2D/2D black phosphorus and covalent triazine framework heterostructure for CO2 photoreduction. ACS Sustain. Chem. Eng. 2020, 8, 5175–5183. [Google Scholar] [CrossRef]
  118. Zhu, X.; Huang, S.; Yu, Q.; She, Y.; Yang, J.; Zhou, G.; Li, Q.; She, X.; Deng, J.; Li, H. In-situ hydroxyl modification of monolayer black phosphorus for stable photocatalytic carbon dioxide conversion. Appl. Catal. B 2020, 269, 118760. [Google Scholar] [CrossRef]
  119. 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]
  120. Ma, Z.; Li, P.; Ye, L.; Zhou, Y.; Su, F.; Ding, C.; Xie, H.; Bai, Y.; Wong, P.K. Oxygen vacancies induced exciton dissociation of flexible BiOCl nanosheets for effective photocatalytic CO2 conversion. J. Mater. Chem. A 2017, 5, 24995–25004. [Google Scholar] [CrossRef]
  121. Ye, L.; Jin, X.; Liu, C.; Ding, C.; Xie, H.; Chu, K.H.; Wong, P.K. Thickness-ultrathin and bismuth-rich strategies for BiOBr to enhance photoreduction of CO2 into solar fuels. Appl. Catal. B 2016, 187, 281–290. [Google Scholar] [CrossRef]
  122. Vu, N.N.; Nguyen, C.C.; Kaliaguine, S.; Do, T.O. Reduced Cu/Pt–HCa2Ta3O10 Perovskite Nanosheets for Sunlight-Driven Conversion of CO2 into Valuable Fuels. Adv. Sustain. Syst. 2017, 1, 1700048. [Google Scholar] [CrossRef]
  123. Pan, A.; Ma, X.; Huang, S.; Wu, Y.; Jia, M.; Shi, Y.; Liu, Y.; Wangyang, P.; He, L.; Liu, Y. CsPbBr3 perovskite nanocrystal grown on MXene nanosheets for enhanced photoelectric detection and photocatalytic CO2 reduction. J. Phys. Chem. Lett. 2019, 10, 6590–6597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Xie, J.; Li, S.; Wang, R.; Zhang, H.; Xie, Y. Grain boundary engineering in atomically-thin nanosheets achieving bright white light emission. Chem. Sci. 2014, 5, 1328–1335. [Google Scholar] [CrossRef]
  125. Liu, Q.; Wu, D.; Zhou, Y.; Su, H.; Wang, R.; Zhang, C.; Yan, S.; Xiao, M.; Zou, Z. Single-crystalline, ultrathin ZnGa2O4 nanosheet scaffolds to promote photocatalytic activity in CO2 reduction into methane. ACS Appl. Mater. Interfaces 2014, 6, 2356–2361. [Google Scholar] [CrossRef] [PubMed]
  126. Li, J.; Zhang, M.; Guan, Z.; Li, Q.; He, C.; Yang, J. Synergistic effect of surface and bulk single-electron-trapped oxygen vacancy of TiO2 in the photocatalytic reduction of CO2. Appl. Catal. B 2017, 206, 300–307. [Google Scholar] [CrossRef]
  127. Liang, M.; Borjigin, T.; Zhang, Y.; Liu, B.; Liu, H.; Guo, H. Controlled assemble of hollow heterostructured g-C3N4@ CeO2 with rich oxygen vacancies for enhanced photocatalytic CO2 reduction. Appl. Catal. B 2019, 243, 566–575. [Google Scholar] [CrossRef]
  128. He, Y.; Rao, H.; Song, K.; Li, J.; Yu, Y.; Lou, Y.; Li, C.; Han, Y.; Shi, Z.; Feng, S. 3D hierarchical ZnIn2S4 nanosheets with rich Zn vacancies boosting photocatalytic CO2 reduction. Adv. Funct. Mater. 2019, 29, 1905153. [Google Scholar] [CrossRef]
  129. Fu, J.; Bao, H.; Liu, Y.; Mi, Y.; Qiu, Y.; Zhuo, L.; Liu, X.; Luo, J. Oxygen Doping Induced by Nitrogen Vacancies in Nb4N5 Enables Highly Selective CO2 Reduction. Small 2020, 16, 1905825. [Google Scholar]
  130. Di, J.; Chen, C.; Zhu, C.; Song, P.; Xiong, J.; Ji, M.; Zhou, J.; Fu, Q.; Xu, M.; Hao, W. Bismuth vacancy-tuned bismuth oxybromide ultrathin nanosheets toward photocatalytic CO2 reduction. ACS Appl. Mater. Interfaces 2019, 11, 30786–30792. [Google Scholar] [CrossRef]
  131. Li, H.; Zhang, L. Photocatalytic performance of different exposed crystal facets of BiOCl. Curr. Opin. Green Sustain. Chem. 2017, 6, 48–56. [Google Scholar] [CrossRef]
  132. Yao, Y.-C.; Dai, X.-R.; Hu, X.-Y.; Huang, S.-Z.; Jin, Z. Synthesis of Ag-decorated porous TiO2 nanowires through a sunlight induced reduction method and its enhanced photocatalytic activity. Appl. Surf. Sci. 2016, 387, 469–476. [Google Scholar] [CrossRef]
  133. Wei, Z.; Rosa, L.; Wang, K.; Endo, M.; Juodkazis, S.; Ohtani, B.; Kowalska, E. Size-controlled gold nanoparticles on octahedral anatase particles as efficient plasmonic photocatalyst. Appl. Catal. B 2017, 206, 393–405. [Google Scholar] [CrossRef] [PubMed]
  134. Schaadt, D.; Feng, B.; Yu, E. Enhanced semiconductor optical absorption via surface plasmon excitation in metal nanoparticles. Appl. Phys. Lett. 2005, 86, 063106. [Google Scholar] [CrossRef]
  135. Kim, K.-H.; Husakou, A.; Herrmann, J. Linear and nonlinear optical characteristics of composites containing metal nanoparticles with different sizes and shapes. Opt. Express 2010, 18, 7488–7496. [Google Scholar] [CrossRef]
  136. Xu, F.; Meng, K.; Cheng, B.; Yu, J.; Ho, W. Enhanced Photocatalytic Activity and Selectivity for CO2 Reduction over a TiO2 Nanofibre Mat Using Ag and MgO as Bi-Cocatalyst. ChemCatChem 2019, 11, 465–472. [Google Scholar] [CrossRef]
  137. Lu, C.; Li, J.; Yan, J.; Li, B.; Huang, B.; Lou, Z. Surface plasmon resonance and defects on tungsten oxides synergistically boost high-selective CO2 reduction for ethylene. Appl. Mater. Today 2020, 20, 100744. [Google Scholar] [CrossRef]
  138. Kamat, P.V. Photophysical, photochemical and photocatalytic aspects of metal nanoparticles. J. Phys. Chem. B 2002, 106, 7729–7744. [Google Scholar] [CrossRef]
  139. Rheinberger, T.; Ohm, D.; Zhumaev, U.E.; Domke, K.F. Extending surface plasmon resonance spectroscopy to platinum surfaces. Electrochim. Acta 2019, 314, 96–101. [Google Scholar] [CrossRef]
  140. Zhao, H.; Zheng, X.; Feng, X.; Li, Y. CO2 reduction by plasmonic Au nanoparticle-decorated TiO2 photocatalyst with an ultrathin Al2O3 Interlayer. J. Phys. Chem. C 2018, 122, 18949–18956. [Google Scholar] [CrossRef]
  141. Low, J.; Qiu, S.; Xu, D.; Jiang, C.; Cheng, B. Direct evidence and enhancement of surface plasmon resonance effect on Ag-loaded TiO2 nanotube arrays for photocatalytic CO2 reduction. Appl. Surf. Sci. 2018, 434, 423–432. [Google Scholar] [CrossRef]
  142. He, Z.; Zhang, J.; Li, X.; Guan, S.; Dai, M.; Wang, S. 1D/2D Heterostructured Photocatalysts: From Design and Unique Properties to Their Environmental Applications. Small 2020, 16, 2005051. [Google Scholar] [CrossRef]
  143. Li, B.; Cao, Z.; Wang, S.; Wei, Q.; Shen, Z. BiVO4 quantum dot-decorated BiPO4 nanorods 0D/1D heterojunction for enhanced visible-light-driven photocatalysis. Dalton Trans. 2018, 47, 10288–10298. [Google Scholar] [CrossRef]
  144. Jo, W.-K.; Kumar, S.; Eslava, S.; Tonda, S. Construction of Bi2WO6/RGO/g-C3N4 2D/2D/2D hybrid Z-scheme heterojunctions with large interfacial contact area for efficient charge separation and high-performance photoreduction of CO2 and H2O into solar fuels. Appl. Catal. B 2018, 239, 586–598. [Google Scholar] [CrossRef]
  145. He, F.; Zhu, B.; Cheng, B.; Yu, J.; Ho, W.; Macyk, W. 2D/2D/0D TiO2/C3N4/Ti3C2 MXene composite S-scheme photocatalyst with enhanced CO2 reduction activity. Appl. Catal. B 2020, 272, 119006. [Google Scholar] [CrossRef]
  146. Fu, J.; Zhu, B.; Jiang, C.; Cheng, B.; You, W.; Yu, J. Hierarchical porous O-doped g-C3N4 with enhanced photocatalytic CO2 reduction activity. Small 2017, 13, 1603938. [Google Scholar] [CrossRef] [PubMed]
  147. Gao, J.; Shen, Q.; Guan, R.; Xue, J.; Liu, X.; Jia, H.; Li, Q.; Wu, Y. Oxygen vacancy self-doped black TiO2 nanotube arrays by aluminothermic reduction for photocatalytic CO2 reduction under visible light illumination. J. CO2 Util. 2020, 35, 205–215. [Google Scholar] [CrossRef]
  148. Xiu, Z.; Guo, M.; Zhao, T.; Pan, K.; Xing, Z.; Li, Z.; Zhou, W. Recent advances in Ti3+ self-doped nanostructured TiO2 visible light photocatalysts for environmental and energy applications. Chem. Eng. J. 2020, 382, 123011. [Google Scholar] [CrossRef]
  149. Wang, H.; Liu, X.; Niu, P.; Wang, S.; Shi, J.; Li, L. Porous two-dimensional materials for photocatalytic and electrocatalytic applications. Matter 2020, 2, 1377–1413. [Google Scholar] [CrossRef]
  150. Tian, N.; Xiao, K.; Zhang, Y.; Lu, X.; Ye, L.; Gao, P.; Ma, T.; Huang, H. Reactive sites rich porous tubular yolk-shell g-C3N4 via precursor recrystallization mediated microstructure engineering for photoreduction. Appl. Catal. B 2019, 253, 196–205. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the different types of CO2 adsorption modes (Adapted from [24]).
Figure 1. Schematic illustration of the different types of CO2 adsorption modes (Adapted from [24]).
Catalysts 11 00418 g001
Figure 2. The energy band structures of semiconductor photocatalysts and the corresponding redox potentials of CO2 reduction into solar fuels (Adapted from [32]).
Figure 2. The energy band structures of semiconductor photocatalysts and the corresponding redox potentials of CO2 reduction into solar fuels (Adapted from [32]).
Catalysts 11 00418 g002
Figure 3. A schematic diagram of CO2 photocatalytic conversion process over photocatalyst. Process (i): light absorption and generation of photocarriers via a semiconductor photocatalyst. Process (ii): charge carrier separation and transfer to the surface of photocatalyst. Process (iii): reactions of CO2 and H2O with electrons and holes, respectively to produce solar fuels.
Figure 3. A schematic diagram of CO2 photocatalytic conversion process over photocatalyst. Process (i): light absorption and generation of photocarriers via a semiconductor photocatalyst. Process (ii): charge carrier separation and transfer to the surface of photocatalyst. Process (iii): reactions of CO2 and H2O with electrons and holes, respectively to produce solar fuels.
Catalysts 11 00418 g003
Figure 4. Classification of low-dimensional materials based on the morphology into two dimensional (2D), one dimensional (1D), and zero dimensional (0D) nanomaterials.
Figure 4. Classification of low-dimensional materials based on the morphology into two dimensional (2D), one dimensional (1D), and zero dimensional (0D) nanomaterials.
Catalysts 11 00418 g004
Figure 5. (a) TEM images of Cs2AgBiBr6 NCs synthesized at different temperature at 200 °C (b) UV-Visible spectra of colloidal Cs2AgBiBr6 NCs, and the inset shows the Tauc plots of colloidal Cs2AgBiBr6 NCs. (c) The photocatalytic CO2 reduction performance of the as-prepared Cs2AgBiBr6 NCs and washed NCs (Adapted from [47]).
Figure 5. (a) TEM images of Cs2AgBiBr6 NCs synthesized at different temperature at 200 °C (b) UV-Visible spectra of colloidal Cs2AgBiBr6 NCs, and the inset shows the Tauc plots of colloidal Cs2AgBiBr6 NCs. (c) The photocatalytic CO2 reduction performance of the as-prepared Cs2AgBiBr6 NCs and washed NCs (Adapted from [47]).
Catalysts 11 00418 g005
Figure 6. TEM images of (a) TNT and (b) TNR (c) the yield of H2 and CH4 after 9 h irradiation over all catalysts (Adapted from [21]).
Figure 6. TEM images of (a) TNT and (b) TNR (c) the yield of H2 and CH4 after 9 h irradiation over all catalysts (Adapted from [21]).
Catalysts 11 00418 g006
Figure 7. (a) TEM image of P-g-C3N4 nanotubes (b) photocatalytic reduction activity for CO2 conversion to produce CO and CH4 of g-C3N4 and P-g-C3N4, (Adapted from [50]).
Figure 7. (a) TEM image of P-g-C3N4 nanotubes (b) photocatalytic reduction activity for CO2 conversion to produce CO and CH4 of g-C3N4 and P-g-C3N4, (Adapted from [50]).
Catalysts 11 00418 g007
Figure 8. Surface morphologies of (a) bulk g-C3N4 and (b) g-C3N4 nanosheets, (c) atomic force microscopic image of g-C3N4 nanosheets, (d) transient photocurrent responses (λ = 420 nm), (e) electrochemical impedance spectra, and (f) time-resolved PL spectra of the bulk g-C3N4 (bulk-CN) and ultrathin g-C3N4 nanosheets (NS-CN) (Adapted from [55]).
Figure 8. Surface morphologies of (a) bulk g-C3N4 and (b) g-C3N4 nanosheets, (c) atomic force microscopic image of g-C3N4 nanosheets, (d) transient photocurrent responses (λ = 420 nm), (e) electrochemical impedance spectra, and (f) time-resolved PL spectra of the bulk g-C3N4 (bulk-CN) and ultrathin g-C3N4 nanosheets (NS-CN) (Adapted from [55]).
Catalysts 11 00418 g008
Table 1. Standard electrochemical potentials of CO2 and H2O at 25 °C and atmospheric pressure [20,24].
Table 1. Standard electrochemical potentials of CO2 and H2O at 25 °C and atmospheric pressure [20,24].
ReactionEo at pH 7
(V vs. NHE)
Solar Fuel
CO2 reductionCO2 + 2H+ + 2e → CO + H2O−0.51CO
CO2 + 8H+ + 8e → CH4 + 2H2O−0.24CH4
CO2 + 6H+ + 6e → CH3OH + H2O−0.39CH3OH
2CO2 + 12H+ + 12e → C2H5OH + 3H2O−0.33C2H5OH
CO2 + 2H+ + 2e → HCOOH−0.58HCOOH
CO2 + 4H+ + 4e → HCHO + H2O−0.48HCHO
H2O oxidation2H2O → O2 +4H++0.81O2
Table 2. Summary of low-dimensional photocatalysts used in photocatalytic reduction of CO2 into solar fuels. NC: Nanocrystal, GQD: Graphene Quantum Dot, NF: Nanofiber, NS: Nanosheet, CNT: carbon nanotube, Mt: montmorillonite, m-CN: modified g-C3N4, NR: nanorod, NW: nanowire, CND: carbon nano dot, p-CN: protonated g-C3N4, PGCN: porous g-C3N4, TEOA: triethanolamine, bpy: bipyridine, C3N4: Melon-based polymeric carbon nitride, UTNS: ultrathin nanosheet, P-g-C3N4: Phosphorus doped g-C3N4.
Table 2. Summary of low-dimensional photocatalysts used in photocatalytic reduction of CO2 into solar fuels. NC: Nanocrystal, GQD: Graphene Quantum Dot, NF: Nanofiber, NS: Nanosheet, CNT: carbon nanotube, Mt: montmorillonite, m-CN: modified g-C3N4, NR: nanorod, NW: nanowire, CND: carbon nano dot, p-CN: protonated g-C3N4, PGCN: porous g-C3N4, TEOA: triethanolamine, bpy: bipyridine, C3N4: Melon-based polymeric carbon nitride, UTNS: ultrathin nanosheet, P-g-C3N4: Phosphorus doped g-C3N4.
DimensionalityMorphologyPhotocatalystLight SourceReducing AgentMain ProductActivity
[µmol∙g−1∙h−1]
Ref.
0DQD
3–12 nm
CsPbBr3300 W Xe lampH2OCO
CH4
4.26
1.53
[45]
QD
2.3 nm
Cs3Bi2I9300 W Xe lampH2OCO1.15[46]
QD
2.9 nm
Cs3Bi2Br9300 W Xe lampH2OCO26.95
QD
2.4 nm
Cs3Bi2Cl9300 W Xe lampH2OCO21.01
NC
9.5 nm
Cs2AgBiBr6AM 1.5GEthyl acetateCO
CH4
2.35
1.6
[47]
QD
9.45 nm
FAPbBr3300 W Xe arc lampH2OCO
CH4
181.25
16.9
[48]
QD
5.86 nm
GQD300 W Xe lamp
420 nm cutoff filter
H2OCH3OH0.695[11]
1DNRCeO2300 W Xe lampH2OCO0.020[49]
NTTiO2300 W Xe arc lamp
320 nm < λ < 780 nm
H2OCH42.128[21]
NRTiO2300 W Xe arc lamp
320 nm < λ < 780 nm
H2OCH41.41[21]
NTP-g-C3N4300 W Xe lampH2O/TEOACO
CH4
2.37
1.81
[50]
NTPGCN300 W Xe lampH2O/MeCN /TEOACO103.6[51]
NTBi12O17Cl2300 W Xe lampH2OCO48.6[52]
NTBi12O17Br2300 W Xe lampH2OCO34.5[53]
2DNSg-C3N4300 W Xe arc lampMeCN/TEOA
(4:1)
CO
CH4
5.407
1.549
[54]
UTNSg-C3N4300 W Xe lampH2OCH4
CH3OH
1.39
1.87
[55]
UTNSSiC300 W Xe lampH2OCO
CH4
1.29
3.11
[56]
UTNSBi2MoO6300 W Xe lampH2OCO3.62[57]
0D/1DQD/NWBlack P/WO3300 W Xenon arc lampH2OCO
C2H4
~ 135
~ 11
[13]
QD (10 nm)/NTWS2/Bi2S3300 W Xe arc lampH2OCH3OH
C2H5OH
9.55
6.95
[58]
QD (3.5 nm)/NWTi3C2/Cu2O300 W Xe lampH2OCH3OH78.50[59]
0D/2DQD (1.6 nm)/NSCuO/WO3300 W Xe lamp
λ > 400 nm
H2OCO1.58[60]
ND (4.4 nm)/NSCND/p-CNXe arc lampH2OCO
CH4
5.88
2.92
[61]
QD/NSTiO2/g-C3N4300 W Xe lamp
λ > 400 nm
MeCN/TEOACO77.8[62]
QD (5nm)/NSAu/TiO2300 W Xe arc lampH2OCO
CH4
19.75
70.34
[63]
QD (7nm) /NSCsPbBr3/Bi2WO6300 W Xe lamp
λ > 400 nm
H2OCO/CH4503 µmol∙ g−1[64]
1D/2DNF/NSTiO2/MoS2350 W Xe lampH2OCH4
CH3OH
2.86
2.55
[65]
NT/NSCNT/g-C3N4200 W Hg and solar simulatorH2OCO
CH4
410
74
[66]
NR/NS/NFAu/TiO2/BiVO4300 W Xe lampH2OCO
CH4
2.5
7.5
[67]
2D/2DNS/NSTi3C2/Bi2WO6300 W Xe lampH2OCH4
CH3OH
1.78
0.44
[68]
NS/NSBi2WO6/BiOI500 W Xe arc lamp
λ < 400 nm
H2OCH42.92[69]
NS/NSMt/m-CN35 W Xenon lampH2O/H2CO
CH4
505
330
[70]
NS/NSSnS2/TiO2300 W Xe lampH2OCH423[71]
NS/NSg-C3N4/BiVO4300 W Xe lamp
λ ≥420 nm
H2OCO
CH4
5.19
4.57
[72]
NS/NSPGCN/Bi12O17Cl2300 W xenon lampH2OCH424.4[73]
NP-NS/NSPd-g-C3N4 /RGOA300 W Xe lampH2OCH46.4[74]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Omr, H.A.E.; Horn, M.W.; Lee, H. Low-Dimensional Nanostructured Photocatalysts for Efficient CO2 Conversion into Solar Fuels. Catalysts 2021, 11, 418. https://doi.org/10.3390/catal11040418

AMA Style

Omr HAE, Horn MW, Lee H. Low-Dimensional Nanostructured Photocatalysts for Efficient CO2 Conversion into Solar Fuels. Catalysts. 2021; 11(4):418. https://doi.org/10.3390/catal11040418

Chicago/Turabian Style

Omr, Hossam A. E., Mark W. Horn, and Hyeonseok Lee. 2021. "Low-Dimensional Nanostructured Photocatalysts for Efficient CO2 Conversion into Solar Fuels" Catalysts 11, no. 4: 418. https://doi.org/10.3390/catal11040418

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