Next Article in Journal
MOS Capacitance Measurements for PEALD TiO2 Dielectric Films Grown under Different Conditions and the Impact of Al2O3 Partial-Monolayer Insertion
Previous Article in Journal
Comparative Adsorptive Removal of Phosphate and Nitrate from Wastewater Using Biochar-MgAl LDH Nanocomposites: Coexisting Anions Effect and Mechanistic Studies
Previous Article in Special Issue
Efficient Charge Carrier Separation in l-Alanine Acids Derived N-TiO2 Nanospheres: The Role of Oxygen Vacancies in Tetrahedral Ti4+ Sites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in TiO2-Based Photocatalysts for Reduction of CO2 to Fuels

1
Laboratory of Advanced Materials Chemistry, Advanced Institute of Materials Science, Ton Duc Thang University, Ho Chi Minh City 700000, Vietnam
2
Faculty of Applied Sciences, Ton Duc Thang University, Ho Chi Minh City 700000, Vietnam
3
Institute of Research and Development, Duy Tan University, Da Nang 550000, Vietnam
4
Key Laboratory of Advanced Materials for Energy and Environmental Applications, Lac Hong University, Bien Hoa 810000, Vietnam
5
Faculty of Materials Technology, Ho Chi Minh City University of Technology (HCMUT), Vietnam National University–Ho Chi Minh City (VNU–HCM), 268 Ly Thuong Kiet, District 10, Ho Chi Minh City 700000, Vietnam
6
Center of Excellence for Green Energy and Environmental Nanomaterials (CE@GrEEN), Nguyen Tat Thanh University, 300A Nguyen Tat Thanh, District 4, Ho Chi Minh City 755414, Vietnam
7
Cambridge Centre for Advanced Research and Education in Singapore (CARES), Campus for Research Excellence and Technological Enterprise (CREATE), 1 Create Way, Singapore 138602, Singapore
8
Department of Materials Science and Engineering, Korea University, 145 Anam-ro, Seongbuk-gu, Seoul 02841, Korea
9
Department of Materials Science, Institute for Surface Science and Corrosion, University of Erlangen-Nuremberg, Martensstrasse 7, 91058 Erlangen, Germany
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2020, 10(2), 337; https://doi.org/10.3390/nano10020337
Submission received: 30 November 2019 / Revised: 2 January 2020 / Accepted: 8 January 2020 / Published: 17 February 2020

Abstract

:
Titanium dioxide (TiO2) has attracted increasing attention as a candidate for the photocatalytic reduction of carbon dioxide (CO2) to convert anthropogenic CO2 gas into fuels combined with storage of intermittent and renewable solar energy in forms of chemical bonds for closing the carbon cycle. However, pristine TiO2 possesses a large band gap (3.2 eV), fast recombination of electrons and holes, and low selectivity for the photoreduction of CO2. Recently, considerable progress has been made in the improvement of the performance of TiO2 photocatalysts for CO2 reduction. In this review, we first discuss the fundamentals of and challenges in CO2 photoreduction on TiO2-based catalysts. Next, the recently emerging progress and advances in TiO2 nanostructured and hybrid materials for overcoming the mentioned obstacles to achieve high light-harvesting capability, improved adsorption and activation of CO2, excellent photocatalytic activity, the ability to impede the recombination of electrons-holes pairs, and efficient suppression of hydrogen evolution are discussed. In addition, approaches and strategies for improvements in TiO2-based photocatalysts and their working mechanisms are thoroughly summarized and analyzed. Lastly, the current challenges and prospects of CO2 photocatalytic reactions on TiO2-based catalysts are also presented.

Graphical Abstract

1. Introduction

Heavy dependence on fossil fuels since the past few centuries for industrialization and economic growth has depleted the sources of carbon-emitting fossil fuels and has led to record-breaking atmospheric concentrations of carbon dioxide (CO2), which is a major component of greenhouse gases, thus contributing to global warming and climatic changes [1,2,3]. To mitigate the problematic impacts of CO2 on human life and the environment, efforts have been made to reduce carbon emissions caused by capturing, storing, and transforming CO2 into other valuable substances [4,5,6]. In addition, CO2 is considered as one of the cheapest and most abundant carbon resources to produce various C1 products such as carbon monoxide (CO), methane (CH4), formic acid (HCOOH), and methanol (CH3OH), which are critical for the synthesis of other value-added chemicals [7,8].
Lately, the electroreduction of CO2 and photoreduction strategies have been paid increasing attention to as promising procedures to utilize CO2 as a carbon building-block for synthesizing hydrocarbon fuels [9,10,11]. Electroreduction of CO2 requires electricity as the driving force for the electrochemical reaction [12,13,14,15]. Therefore, the requirement of the energy input is concerning. This energy can be generated from non-renewable energy resources. Meanwhile, solar energy is among the most critical renewable energy sources along with wind energy, wave energy, and hydrogen energy [16]. Photoreduction systems activated by solar energy are a green but intermittent power source for catalyzing the reduction of CO2. Such systems have been considered one of the most sustainable and cost-effective approaches to exploit solar energy combined with the utilization of anthropogenic CO2 as a starting material of the carbon cycle to store energy in the form of chemical fuels [8,17,18].
Titanium dioxide (TiO2) has been the benchmark for photocatalytic studies over nearly four decades since the seminal work by Fujishima and Honda, who showed the high photochemical stability, low-cost features, facile preparation, abundance, and low toxicity of TiO2 [19]. TiO2 exhibits the most extensive applications as a semiconductor and can be used in a wide range of photocatalysis applications including water splitting, photodegradation of pollutants, dye bleaching, and water desalination, thus contributing to solving the environmental and energy crisis [20,21,22,23,24,25,26,27]. Parallel to water splitting technologies, an extensive line of studies on TiO2-based photocatalysts for CO2 reduction has been conducted to overcome obstacles including limited light-harvesting, charge separation hindering improved process efficiencies, inefficient selectivity of products, the inability to suppress the competing the hydrogen evolution reaction (HER), and catalytic instability [28,29,30,31]. In this study, we therefore review the photocatalytic conversion of CO2 to fuels using the recently advanced TiO2-based materials. The fundamentals of CO2 photoreduction are discussed to understand the factors affecting the performance of photocatalysts, and the updated progress in the enhancement of its efficiency is comprehensively discussed. The effective strategies for overcoming the limitations of TiO2-based photocatalysts are categorized as follows: (i) Design of nanostructured TiO2-based catalysts, (ii) modification of TiO2 facial properties, and (iii) cocatalysts for TiO2 photocatalysts. Lastly, the current challenges and directions for future progress are provided, which could offer guidelines for the future development of CO2 reduction on TiO2-based photocatalysts.

2. Fundamentals of CO2 Photocatalytic Reduction

Thermodynamically, CO2 is one of the most stable linear molecules, showing the highest oxidation state of the carbon atom with a Gibbs free energy of ∆Go = −394.4 kJ mol−1. A high input energy is required to activate the CO2 molecule, following the equation:
∆Go = nFEo.
In reality, however, to break the C=O bond and bend the linear geometry of CO2, a more negative potential than Eo is required to make the ∆Go value negative enough to spontaneously accelerate the conversion of CO2 and to compensate for the overpotential (Equation (2)):
E = η + Eo
where ∆Go is the standard Gibbs free energy, E is the required potential, Eo is the standard potential, η is the overpotential, n is the number of transferred electrons, and F is the Faraday constant.
In addition, the presence of reductants providing protons to react with the adsorbed CO2 on a photocatalyst surface and electrons to form products assists the process of CO2 conversion by lowering the required input energy in comparison to the absence of reducing agents, in which the reduction of CO2 proceeds by a photocatalyst alone and is much more difficult because of the high stability of inert CO2. Therefore, photoreduction of CO2 is commonly conducted with a reductant, such as H2 or/and H2O, making the conversion of CO2 into fuels more feasible and more efficient from the view point of consumed energy [32,33]. In a CO2(g)/H2O(g/L) or CO2(g)/H2(g) system, photoexcited electrons are generated under light irradiation and are transferred, and then they react with the adsorbed CO2 on the photocatalyst surface along with protons provided by the reducing agents to yield the reduced products. H2O (vapor or liquid) is well known as an inexpensive and safer reductant than H2 gas which is dangerous, although HER is also coupled in the H2O phase, which lowers the CO2 reduction efficiency because of their similar potentials for generating intermediates. Various products from the CO2 reduction reaction combined with the competing HER and their standard potentials are presented in Table 1.
Importantly, the formation of unstable radical anion CO2 has been proved to be the most widely recognized mechanism to activate CO2 molecules in the initial stage of the photoreduction process. However, this activation requires a very negative equilibrium potential of −1.49 V vs. reversible hydrogen electrode (RHE) at pH = 7 to form the CO2 intermediate, resulting in a significant increase in the required negative potential applied for electrochemical reduction of CO2 compared with the standard potentials in Table 1.
The overall principles of the photoreduction CO2/H2O(g/L) system are illustrated in Figure 1. Under light irradiation, incident photons with an equal or higher energy than that of the bandgap semiconductor (e.g., 3.2 eV for anatase TiO2), which act as the photocatalyst, are absorbed and the pair-charges, electrons and holes, are generated. This is followed by the separate transportation of the generated electrons and holes to the photocatalyst surface. Subsequently, the photogenerated electrons reduce CO2 to value-added chemicals (step number 1) while water is oxidized by holes to oxygen molecules (step number 2). The recombination of the electron-hole pairs, which may consume a lot amount of photogenerated charges, in the bulk and on the surface can occur in step numbers 3 and 4, respectively. Proton reduction, which is called HER, appears to be spontaneous in step number 1 under aqueous conditions and competes with the reduction of CO2.
For the selection of the photocatalyst material, the conductive band (CB) and valence band (VB) positions of semiconductors are critical factors, which help analyze the possibility for the redox reaction. Alternative semiconductors possess different CB and VB positions, thus varying the band gap, which is the potential energy between the CB and VB (Figure 2a) [34,35,36,37,38,39,40]. Theoretically, for instance, anatase TiO2 with CB and VB potentials of −0.5 eV and 2.7 eV, respectively, enables the reduction of CO2 into several fuels (such as CH4, CH3OH, and CO) and oxidizes H2O because its CB is more negative than the reduction potential to form products, whereas its VB more positive than the redox potential of the reductant. Nonetheless, the high band gap (3.2 eV) requires a high energy of photons with a short wavelength (UV light) to activate photoexcitation.
Insights into the interaction of adsorbed CO2 with the semiconductor surface are important because the structure and composition of the catalytic sites can significantly vary the pathways and selectivity of different products. Thus, Curtiss et al. proposed diverse binding configurations of activated CO2 molecules on the TiO2 surfaces (Figure 2b) [41]. They also found that the cation sites on the Ti surface can lower the reaction barrier to activate CO2 and the stability of CO2 intermediates to enhance CO2 photoreduction.
Briefly, the photoreduction of CO2 on TiO2 can encounter some intrinsic challenges such as the low light-harvesting efficiency, charge recombination, thermodynamic difficulty in activation of CO2, the required high negative overpotential for intermediates, the wide distribution of products from CO2 reduction in kinetics, and the competition of the favorable HER from H2O reduction. These limitations could be overcome to enhance the photocatalytic activity of CO2 on TiO2-based photocatalysts by focusing on some critical factors including surface molecular structures and charge transfer behaviors, which will be discussed in the following sections.

3. Advances in TiO2-Based Photocatalysts for CO2 Reduction

3.1. Nanostructured TiO2 Design

Various nanostructured TiO2-based photocatalysts with different surface molecular structures play a significant role on specifying the surface adsorption properties and surface electronic structures, and thus, the photocatalytic activity of a photocatalyst is determined by its electronic structure resulted from the light absorption efficiency and redox potential of the excited charges. In addition, the surface charge transfer configurations also contribute to the photocatalytic activity by accelerating electron and hole transportation and impeding electron-hole recombination [42,43]. To that end, many studies have reported advances for CO2 photoreduction on engineering of crystal phases [44,45,46,47,48], facets [41,49,50,51,52,53,54,55,56,57], and TiO2-dispersed on porous materials [58,59,60,61].

3.1.1. Crystal Phases

TiO2 possesses three types of crystal phases: Anatase, rutile, and brookite. The first two crystal structures are usually used as photocatalysts. The photocatalytic activity can be improved by creating a mixture of different crystal structures rather than using a single crystal phase due to the improved charge separation and transfer efficiency induced by heterojunction effects [44,45,46]. To this end, Chai et al. controlled the composition of anatase and rutile phases in TiO2 to enhance the photoactivity by simply varying the annealing temperature [47]. As expected, TiO2 nanoparticles (NPs) were composed of both anatase and rutile crystal phases, leading to an increase in the electron-hole separation caused by the created heterojunction (Figure 3a,b). They also found that the optimal fraction of phase content for their TiO2 photocatalyst was 17.5% rutile phase. Moreover, another factor that contributes to the catalytic activity is morphology. A synergistic influence of crystal phase content and morphology can be expected to improve photocatalytic performance. To elucidate this hypothesis, 1-D hierarchical mesoporous TiO2 nanofibers were synthesized using combined electrospinning and sol-gel methods. The obtained photocatalyst exhibited higher activity than TiO2 NPs prepared only by the sol-gel procedure [48]. The fast charge transport and impeded electron-hole recombination caused by the nanofiber morphology led to improved activity. Meanwhile, the 1-D fiber photocatalyst with a higher ratio of rutile to anatase phases (20:80) produced four and 2.5 times higher H2 and CO, respectively, than the counterpart containing a lower ratio of rutile to anatase phase (7:93), revealing the crucial role of crystalline phase composition in photocatalytic activity.

3.1.2. Effect of Facets

Facet engineering of TiO2 photocatalysts, which is another promising strategy to achieve high photocatalytic activity via CO2 adsorption on the catalyst surface, photoinduced activation, and charge separation, has been extensively studied [41,49,50,51,52,53,54,55]. In general, the {101} facet is more thermodynamically stable than {001} facets, suggesting that anatase TiO2 crystals usually possess the dominant exposed {101} facets [49]. Using first-principle calculations, Curtiss et al. investigated the binding configuration of intermediates on anatase {101} surfaces and found that the reduced {101} anatase surfaces favor CO2 binding along with charge transfer to CO2 [41,50]. As a result, the adsorption of CO2 and the charge transfer from the surface of anatase TiO2 photocatalysts to CO2 was theoretically demonstrated to be enhanced on the preferentially exposed {101} facets. Nevertheless, Han et al. unveiled that {001} facets are more reactive in photocatalytic reduction of CO2 than the low-energy {101} surfaces because of the high density of active unsaturated Ti atoms and active surface oxygen atoms [51]. Therefore, the coexposed facet approach is believed to promote spatial charge separation and form a facet-related heterojunction, thus resulting in longer charge carrier lifetimes, which can hinder electron-hole recombination and accelerate electron and hole transportation while maintaining the high adsorption of CO2. In a subsequent study, Yu et al. appropriately manipulated the efficient ratio of coexposed {101} and {001} facets to synergistically leverage the CO2 adsorption and reactivity on both facets [52]. By varying the HF amount used during the synthesis of TiO2 nanosheets, the growth of the {101} and {001} facets can be adjusted. Therein, the optimal HF amount used is 4.5 mL to obtain the {001} and {101} exposed facets in a proportion of 55:45, with a CH4 yield from photoreduction of CO2 of 1.35 μmol g−1 h−1 (Figure 3c,d).
Nonetheless, Kar et al. very recently reported an efficient method to transform anatase phase TiO2 nanotubes into the rutile phase with the {110} facet as the dominant plane using the flame annealing method (Figure 3e) [56]. Flame-annealed nanotubes prepared in a water-based electrolyte (FANT-aq) exhibited superior yield (156.5 μmol g−1 h−1) in the presence of the rutile phase as the only crystalline phase. High activity of the FANT-aq can be attributed to the increased visible-light absorption due to the reduced band gap induced by the pure rutile phase, defects and sub-energy levels, square morphology, and the preferentially oriented rutile {110} facet.
Remarkably, Yang et al. lately prepared hydrogenated nanotube/nanowire photocatalysts assembled from TiO2 nanoflakes with an exposed clean {111} facet, which is believed to be the origin of the enhanced solar absorption and the good charge separation to attain a superior photocatalytic performance up to 1708.1 μmol g−1 h−1 CH4 at 373 K under the irradiation of a 300-W Xe lamp, which is one of the best performances for methane production. Importantly, a spontaneous electric field (Es) is demonstrated to exist between the polar TiO2 {111} and ( 1 ¯ 1 ¯ 1 ¯ ) planes, under which photoinduced electrons and holes migrate to the positive Ti–TiO2 {111} and negative O–TiO2 ( 1 ¯ 1 ¯ 1 ¯ ) polar surfaces, respectively, which is followed by the subsequent redox reactions to reduce CO2 into CH4, as shown in Figure 4 [57].

3.1.3. Well-Dispersed TiO2 on Porous Materials

Porous materials have been utilized as a supporter for catalysts because of their high surface area, well-dispersed active sites, large adsorption capacities, and alternative pore sizes suitable for facile diffusion and transportation of gas and liquid reactants. Therefore, porous materials including zeolite and hierarchical silica have been considered as potential supporting candidates in photocatalytic reduction of CO2. As one of the earliest investigations, in a study by Anpo et al., highly dispersed TiO2 on zeolites (Ti-MCM) were synthesized by a hydrothermal method for photoreduction of CO2 to CH4 and CH3OH [58,59,60]. The well-dispersed Ti zeolites catalyzed the photoreduction of CO2 into CH4 and exhibited approximately 10-fold higher photocatalytic efficiency than TiO2 powder. Interestingly, high activity for CH3OH formation was achieved on Ti-MCM-48, whereas the bulk TiO2 counterpart only produced CH4 as the product. The results indicate that the 3-D channel structure of zeolites in Ti-MCM-48 is not only beneficial for the high dispersion of TiO2, but is also beneficial for the high amount of adsorbed CO2 and production of CH3OH. It is plausible that the formation of charge transfers excited complexes (Ti3+-O)* on the isolated TiO2 species in zeolites. Moreover, OH∙ radicals are formed from the decomposition of H2O, which react with methyl from reduction of CO2 and finally generate CH3OH [59]. In a subsequent study, Ti-SBA-15 synthesized with a higher pore size exhibited 106 μmol CH4 g−1 h−1 production via photoreduction of CO2 into CH4 and CH3OH under UV irradiation. This activity is 20 times higher than that of Ti-MCM-41 and Ti-MCM-48 [61].

3.2. Modification on TiO2 Surface

Surface modifications of the crystal structure can act as active sites on metals and/or metal-oxide catalysts, modifying the electronic structure and binding strength of intermediates, and thus, influencing the catalytic activity. Two popular approaches on TiO2-based photocatalyst surface are related to oxygen vacancies [62,63,64,65,66,67,68,69,70,71,72] and basic-site functionalized modifications [73,74,75,76,77,78].

3.2.1. Oxygen Vacancies

Oxygen defects are one of the most famous and efficient approaches to enhance photocatalytic activities [62,63]. The behavior of CO2 adsorbed on oxygen defects (Vo) on the TiO2 {110} surface was investigated by Lee et al. [64] Their results unveiled that with the existence of Vo, one oxygen atom of CO2 can be situated at the position of the oxygen defect and can the CO2 molecule on TiO2 with a bridge formed by the adsorption of the oxygen atom from CO2, which can enhance CO2 trapping and lower the activation barrier (Figure 5a,b). Recently, Alexandrova et al. employed density functional theory calculations to systematically study the role of oxygen vacancies on anatase TiO2 {101} surface (Figure 5c–e) [65].
The significance of oxygen vacancies on defected TiO2 has been ascertained in the enhancement of CO2 adsorption, activation, dissolution, and stabilization of reaction intermediates. Li et al. found that the photocatalytic activity of the enriched oxygen defect on TiO2 nanocrystals could be increased compared with that on unmodified crystals [66]. By controlling coexposed {001} and {101} facets, the highest performance was observed on TiO2−x{001}-{101} with a CO production rate of 55 and 27 μmol g−1 h−1 (quantum yield of 0.31 and 0.134%) under UV-VIS and visible light, respectively (Figure 6a). This excellent performance, even under visible light, can be attributed to the high CO2 adsorption capacity, combination of {001} and {101} facets, and the new energy state of Ti3+ cations, which improve the activation and conversion kinetics of the adsorbed species. Later, a facile hydrothermal method was reported to synthesize TiO2 with a high concentration of Ti3+ [67]. Because of the oxygen vacancies and coexposed {001} and {101} facets, the photocatalytic activity was improved with wide-spectrum solar light absorption to enhance CO2 photoconversion and CH4 selectivity under sunlight irradiation.
Oxygen vacancies in TiO2 can also be formed from the introduction of other elements to achieve high efficiencies under visible light. This is because the formation of Ti3+ caused by oxygen deficiency creates an intermediate band and thus accelerates electron-hole pair separation. In addition, the oxygen vacancy sites on the surface of the photocatalysts could allow adsorption of more CO2 compared with the bulk surface [62,68,69,70,71,72]. Ye et al. prepared an oxygen-deficient perovskite structure of self-doped SrTiO3−δ for CO2 photoreduction under visible light [62]. Using a one-step combustion method followed by calcination in Ar at temperatures ranging from 1200 to 1400 °C, oxygen vacancies were generated coupled with Ti3+ to activate visible-light absorption. The increase in oxygen vacancies was found to be positively correlated with the CO2 chemical adsorption capacitance. The optimized SrTiO2.83 photocatalysts showed the highest conversion of CO2 in photoreduction to CH4 with a quantum efficiency of 0.21% at 600 nm. Analogously, for the first time, Li et al. reported the spontaneous dissolution of CO2 into CO even in the dark on a defective and partially oxygen-depleted Cu(I)/TiO2−x surface [68]. The photocatalyst was prepared by thermal annealing of Cu(OH)2/TiO2 in an inert environment at moderate temperatures to generate surface oxygen vacancies due to partial oxygen loss and the reduction of Cu2+ to Cu+ (Figure 6b). Moreover, the oxidation state of copper on TiO2 induced by the pretreatment conditions can play an important role in the photoreduction performance of catalysts. The active charge separation of H2-pretreated Cu/TiO2 can be attributed to the synergistic effect of the mixture of Cu+ and Cu and the oxygen vacancy defects [69]. In addition, metallic Cu particles have been reported to promote the formation of oxygen vacancies in black TiO2-coated Cu NPs through metal-oxide interactions, thus improving the visible-light adsorption and the adsorption of CO2 to improve photocatalytic activity [70]. Furthermore, oxygen vacancies are also formed from the introduction of co-dopants [71]. Lee et al. reported high production yields of CH4 and CO under visible light, which are 933 and 588 μmol g−1 h−1, respectively, on 2Cu@4V-TiO2/PU (PU = polyurethane) [72].

3.2.2. Surfaces of Basic Functional Sites

Because CO2 molecules can act as an electron acceptor because of the electrophilic carbon atom, some studies have explored the utilization of basic hydroxides, oxides, and amine functional groups to promote the chemisorption and activation of photocatalysts. Ye et al. modified the TiO2 surface by loading 3 wt% solid NaOH to obtain 52 μmol g−1 CH4 within 6 h; in contrast, bare TiO2 did not show any CH4-production activity [73]. Similarly, Wang et al. loaded MgO onto the TiO2 surface to transform CO2 into CO via photoreduction with H2O vapor [74]. The introduction of MgO into Pt-TiO2 can generate 2.2 μmol of CH4 within 10 h [75]. Lately, Li et al. coated an ultrathin MgO layer on porous TiO2 mixed anatase-rutile phases by atomic layer deposition (ALD) [76]. They increased the number of atomic MgO layers from 1 to 100 and found that five layers of MgO exhibited 4- and 21-fold higher CO production compared with pristine porous-TiO2 and common P25, respectively (Figure 6c–e). The enhanced activity can be ascribed to the increased concentration of surface Ti3+ species and hydroxyl groups caused by the uniform dispersion of MgO layers, which act as active sites for adsorption and photoreduction of CO2. Furthermore, the deposited MgO layers promoted electron-hole separation because of the passivation of TiO2 states on the surface. Likewise, amine-functionalized TiO2 was synthesized using monoethanolamine (MEA) to improve CO2 adsorption, thus enhancing the photocatalytic activity by approximately three times than the pristine TiO2 [77]. Liu et al. also reported the high visible-light harvesting ability and improved CO2 adsorption capacity of hierarchical amine-functionalized titanate nanosheet-based yolk@shell microspheres using one-pot diethylenetriamine mediated anhydrous alcoholysis method [78]. As a result, the conversion of CO2 exhibited a CH3OH yield of 8 μmol g−1 h−1 under visible-light irradiation.

3.3. Cocatalysts

The introduction of cocatalysts into TiO2 photocatalysts can be beneficial for CO2 reduction not only by enhancing the charge transfer via formation of a heterojunction, thus improving the charge carrier separation, but also by forming supplementary active sites for the reaction to increase the adsorption of CO2 and stabilization of intermediates in the photoreaction [79,80,81,82].

3.3.1. Metal and Metal-Oxide Cocatalysts

Metals and metal oxides have been frequently employed as cocatalysts with TiO2 to improve the performance of TiO2 [35,75,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96].
When a cocatalyst system is illuminated, a transfer of photoinduced electrons from TiO2 to the metallic surface possessing a larger work function at the interface can occur, and thus, metallic sites can act as an electron-sink to activate CO2 and generate the bent CO2 intermediate [6,80]. Wang et al. reported the effect of noble metal cocatalysts on the photocatalytic activity of TiO2 and found that the rate of CH4 formation increased as follows: Ag < Rh < Au < Pd < Pt. This result agrees well with the increase in the charged-pair separation efficiency [75]. Pd NPs on the surface of TiO2 have also been reported to act as sites for the adsorption and activation of CO2 [82].
In addition, the size of metallic NPs is a crucial factor determining the activity and CH4-formation rate. The optimum size of Pt NPs on 1-D nanostructured TiO2 single crystals to produce CH4 at an excellent rate of 1361 μmol g−1 h−1 was found to be 1.04 ± 0.08 nm; the result was associated with a quantum yield of 2.41%, which is advanced compared with the production by pristine TiO2 and P25 catalysts [35]. The energy level of the small Pt coated on 1-D TiO2 was rationally positioned at a higher level than the CB edge of TiO2, impeding photoinduced electron transfer (Figure 7a), which is beneficial for CO2RR. Recently, Liu et al. synthesized ultrafine 1.1-nm Pt nanoparticles photoreduced and highly dispersed on ultrathin TiO2 fabricated by deprotonated ethylene glycol as a support (P-Pt/TiO2-U) [83]. The synergistic effects from both ultrathin TiO2 and ultrafine highly dispersed Pt are as follows: Increased electron rate transfer on ultrathin TiO2 nanosheets with abundant defects and ultra-large surface area, the facilitated separation of photogenerated electron-hole pairs induced by ultrafine Pt NPs, which was proved to improve the light-harvesting capacity and quantum efficiency, and finally, the improved adsorption ability of CO2 caused by the synergy of metal and support (Figure 7b,c).
As mentioned above, Cu is a popularly used transition metal for cocatalysts of TiO2. Numerous studies have assessed its contribution to cocatalyst systems. Cu-moieties decorated on the surface of N-doped TiO2 nanotubes and Cu on TiO2-SiO2 catalysts have demonstrated improved CH4 formation from CO2RR. Interestingly, Corma et al. reported the conversion of CO2 to formic acid with a yield of 25.7 μmol g−1 h−1 on Cu-doped TiO2 under UV-rich illumination with Na2S [84]. CuO-TiO2 hollow microspheres with CuO extensively dispersed on the shell have been reported to improve the light-harvesting efficiency to produce CO and CH4 from photoreduction of CO2 under UV conditions [85]. This result can be attributed to the improved electron trapping ability induced by CuO. Meanwhile, hydrogenated hollow microspheres, in which Cu2+ was reduced to Cu0, could further enhance CH4 production because of the hole capturing caused by Cu0 and H2 formation from water dissociation for CH4 evolution. Recently, Jiang and co-workers employed atomically dispersed Cu supported on ultrathin TiO2 nanosheets to photocatalytically reduce CO2 to CO [86]. Most importantly, Cu is an abundant, non-toxic, and low-cost metal, which is a potential alternative to noble metals. In this study, a method is also presented to recycle the catalyst for long-term utilization. Similarly, Cu was recently ascertained to suppress the hydrogen evolution in CO2 photoreduction on Cu-TiO2 to selectively produce CO at a high rate [81].
Furthermore, the employment of binary cocatalysts was investigated on Cu2O/Pt/TiO2 and MgO-Pt/TiO2 systems [75,87]. The introduction of Pt on TiO2 was demonstrated to promote the capture of photogenerated electrons and to hinder pair-charge recombination; however, H2 evolution was spontaneously increased. To suppress H2 generation and enhance the chemisorption and activation of CO2, Cu2O was deposited as a shell respected with Pt core, and analogously, MgO was deposited as an amorphous layer onto Pt/TiO2. As a result, high CH4 selectivities of 85% and 83% were obtained on Cu2O/Pt/TiO2 and MgO-Pt/TiO2 catalysts, respectively. In another study, a binary cocatalyst system of Ag/Pd was reported to enhance the photocatalytic activity for CO2 to CH4 on Ag/Pd supported on N-doped TiO2 with a production rate of 79 μmol g−1 h−1 [88]. Recently, a multi-heterojunction was created on TiO2–MnOx–Pt films (Figure 8). Two heterojunctions were created: Aa p–n junction between the MnOx and TiO2 {001} facet and a metal–semiconductor junction between Pt and TiO2 {101} facet. Thus, efficient separation of charged pairs was obtained to produce three times higher CH4 and CH3OH compared with the pristine TiO2 nanosheets [89].
Another study combined some of the effective strategies mentioned above. Ye et al. fabricated a photocatalyst using hydrous hydrazine Au–Cu bimetal as the cocatalyst supported on SrTiO3/TiO2 coaxial nanoarchitecture arrays to create a heterojunction with fast electron-transfer [97]. As a result, excellent performance was reported on the catalyst with CO as the main product with a production rate of 3770 μmol g−1 h−1 along with a CH4 production rate of 421.1 μmol g−1 h−1. Remarkably, the production of C2+ products (C2H6: 190.1, C2H4: 73.3, C3H6: 40.8 μmol g−1 h−1, respectively) is very noticeable due to the formation of C2+ products with higher values in industry from photoreduction of the problematic CO2 gas. This research can be a representative study that manipulated various approaches to achieve a high production rate for photoreduction of CO2, shedding light on the hybrid nanostructured design for further enhancement of the rate in the future.

3.3.2. Plasmonic Effect

One of the most useful effects induced by the incorporation of noble metals on the TiO2 surface to enhance the visible-light-harvesting efficiency and charge separation is the surface plasmon resonance (SPR). Many applications from SPR have recently emerged as an efficient approach to advance the conversion of water and CO2 [90,91,92]. The enhanced activity of plasmonic effect-related photocatalysts is ascribed to the direct charge transfer mechanism or the enhancement of the local electric field induced on the noble metallic atoms. Ag was loaded on TiO2 nanotube arrays to improve CH4 formation from photoreduction of CO2 under visible light. The improved light absorption from the nanotube arrays of TiO2 combined with the deposition of Ag NPs inside TiO2 nanotubes was believed to cause the plasmonic effect with hot electron generation [93].
In addition, Au is also known to exhibit the SPR effect on TiO2. The synergistic combination of the SPR effect from Au NPs and the role of Pt as electron-sink on nanohybrid Au/Pt co-decorated on TiO2 nanofibers was reported for visible-light harvesting and was found to hinder charge recombination of photoexcited TiO2 [94]. Tahir synthesized montmorillonite (MMT) dispersed Au/TiO2 nanocatalysts through a simple sol-gel method [95]. The photocatalytic performance of MMT-dispersed Au/TiO2 under simulated solar light was enhanced because of the SPR effect of Au, catalyzing photoreduction of CO2 to CO with a high production rate of 1223 μmol g−1 h−1. Remarkably, Garcia et al. fabricated Au–Cu alloy nanoparticles supported on TiO2 to achieve excellent performance [96]. The selectivity of electrons toward CH4 evolution could reach 97% with a high production rate of approximately 2200 μmol g−1 h−1 under visible-light irradiation. The enhanced visible-light-harvesting ability can be ascribed to the plasmonic effect from Au. Furthermore, the presence of Cu bonding to the CO* intermediate on the photocatalyst could lead to the high selectivity of CH4 through a “carbene pathway” with the detection of some intermediates including CO2, Cu-CO, and carbon deposits on the surface (Figure 9).

3.4. Hybrid TiO2 Nanocomposites

The employment of other materials into TiO2 photocatalysts is fruitful for CO2 reduction because of supplied high active sites surface area, facilitation of charge transfer to enhance CO2 adsorption, and impeding electron-hole recombination [98,99,100,101,102,103,104,105,106,107,108,109,110].

3.4.1. Carbon-Containing Composites

Carbon-based materials possess many applicable features such as large surface areas for active sites, broad electronic properties, and various architectural nanostructures, leading to numerous applications in the previous decades. Thus, carbon materials including graphitic carbon derivatives, graphene, graphene oxide (GO), and carbon nanotubes have been incorporated with semiconductor-based materials for facial synthesis and for lowering the fabrication cost while attaining the light absorption in the visible light and enhanced charge separation and transport [98,99,100,101,102,103,104,105,106,107,108]. Liang et al. reported solvent-exfoliated graphene on GO-TiO2 systems to functionalize titania for the photochemical reduction of CO2 [99]. An increment in the visible-light harvesting was observed on the less defective solvent-exfoliated graphene caused by the fast-electrical mobility, accelerating the photoinduced electron transport into reactive sites.
Next, Zou et al. prepared a hollow spherical structure from titania Ti0.91O2 and graphene nanosheets [100]. The hollow spherical structure and the unique composition enhanced the photocatalytic activity because of the fast charge transfer, long-lasting lifetime of charge carriers, and enhancement of light absorption. In agreement with this study, carbon@TiO2 hollow spheres were recently reported to increase the photocatalytic activity to produce CH4 and CH3OH [101]. Importantly, the carbon content was found to be correlated to the charge transfer efficiency. In their subsequent study, Zou et al. developed an in situ reduction-hydrolysis method to synthesize TiO2-graphene nanosheets [102]. Using this technique, GO was reduced to graphene simultaneously with the formation of TiO2 from hydrolysis of Ti4+ dihydroxy is to create a 2D sandwich-like structure. Therein, the TiO2 deposited on graphene acted as a stabilizing agent for graphene nanosheets, whereas the Ti3+ observed on TiO2 NPs could trap the photoexcited electrons coupled with the suppression of the electron-hole pair recombination. Interestingly, the photocatalysts could catalyze the photoreduction of CO2 toward C2H6 with a production rate of 16.8 μmol g−1 h−1, which is different from other reported catalysts, proving that the architectural design of hybrid nanocomposites can vary product selectivity.
Among the carbon-based candidates, graphitic carbon nitrides (g-C3N4) with an increased concentration of nitrogen on the surface and a narrower band gap (2.7 eV) than TiO2 have attracted increasing attention because of the ability for CO2 activation. Cu-TiO2 was dispersed on g-C3N4 to enhance the photocatalytic performance of CO2 conversion to CH4 [103]. In another study, the charge transfer between semiconducting C3N4 nanosheets and a Ru(II)–Re(I) binuclear complex (RuRe) was improved using rutile TiO2 nanocrystals as a modifier [104]. The RuRe/TiO2/NS-C3N4 hybrid could promote both CO formation rate and turnover number under visible-light irradiation as a result of the advanced lifetime of photoinduced electrons. Zhao et al. wrapped Pt/TiO2-nanocrystals with reduced GO (rGO) to synthesize a core-shell structure. The surface residual hydroxyl on the rGO shell and the extended π bonds proved to increase the CO2 adsorption and activation, while the whole hybrid core-shell structure enhanced the electron transfer and the separation efficiency to highly selective production of CH4 up to 99.1% with a production rate of 41.3 μmol g−1 h−1 [105].
Recently, Petit et al. employed carbon nitride nanosheets (CNNS) from g-C3N4 and grew TiO2 and control its facets on CNNS, favoring the formation of {001} facets because of their enhanced photocatalytic activity [106]. The hybrid TiO2/CNNS heterostructures exhibited superior CO2 adsorption and charge transfer acceleration, resulting in the availability of photoexcited electrons under UV-Vis illumination, which is 10-fold higher compared with that exhibited by pristine materials. Transient absorption spectroscopy analyses showed that the hole transfer from TiO2 to CN was observed. The interfacial charge transfer via the heterojunction could inhibit the recombination of paired charges (Figure 10). This revealed that the fine control of TiO2 facets could lead to a high activity for photoreduction of CO2 associated with a hybrid nanocomposite structure with carbon nitride materials, thus enabling charge transfer to enhance CO2 adsorption, and impeding electron-hole recombination can be a promising method to increase the production rate.
Lately, Barbieri et al. reported a continuous operating system employing C3N4-TiO2 Nafion photocatalytic membrane reactors to examine the effects of crucial parameters including reaction pressure, H2O/CO2 feeding ratio, or contact time on the performance and selectivity of various detected products from photoreduction of CO2 (MeOH and HCHO) [107].
Similarly, in another work, P-O linked Z-scheme g-C3N4/TiO2 nanotube composites were fabricated to enhance the visible-light harvesting capability and the charge separation caused by g-C3N4 and P-O links [108]. CO2 photoreduction yielded 46.9 mg L−1 h−1, 38.2 mg L−1 h−1, and 28.8 mg L−1 h−1 of acetic, methanol, and formic acid, respectively, which was ~3.3, 3.5, and ~3.8 times the production from TiO2 nanotubes.

3.4.2. Other Composites

Xu et al. employed an in situ hydrothermal method to deposit CuInS2 nanoplates onto TiO2 nanofibers to achieve the Z-scheme TiO2/CuInS2 heterostructure; excellent photoreduction activity of CO2 to CH4 and CH3OH under light irradiation was observed [109]. A mechanism was proposed for the high activity of TiO2/CuInS2 based on the direct Z-scheme heterojunction, in which photoexcited electrons from the TiO2 CB could be transferred and recombined with the hole of CuInS2 VB, whereas electrons generated on CuInS2 were catalyzed for the photoreduction of CO2 (Figure 11).
The introduction of two-dimensional inorganic compounds, MXene, as a supporter for TiO2 was investigated. Yu et al. grew TiO2 NPs on Ti3C2 in situ by calcination. An ultrathin fluffy rice crust-like structure was observed on the composite, which is beneficial for high surface for active sites. In addition, the highly conductive Ti3C2 assisted photoexcited electron transfer and prevented paired charge recombination. As a result, the TiO2/Ti3C2 composite achieved a 3.7 times higher CH4-production rate than commercial TiO2 (P25) [110].

4. Conclusions and Perspective

Photoreduction of CO2 is an attractive approach to utilize CO2 gas for producing value-added carbon chemicals to achieve multiple purposes. This method has several advantages such as contributing to the energy demands using CO2 as a significant carbon resource, diminishing environmental issues from CO2, and leveraging clean, renewable energy from sunlight. To mimic natural photosynthesis, in which solar energy is used to convert CO2 and H2O to carbohydrates and oxygen, artificial photosynthesis was performed to develop photocatalysts, which play an identical and sophisticated role like terrestrial plants to efficiently harvest sunlight and separate the photogenerated charges for reactions for the conversion of CO2 and H2O into energetic chemicals. Overall, highly efficient photoreduction of CO2 to fuels can be achieved by a combination of some critical strategies, including: High adsorption and activation of CO2, large surface areas for active sites, efficient sunlight harvesting, generation of photoexcited electron-hole pairs, accelerated charge transfer to reactive site, elongated lifetime of charge carriers, and effective inhibition of charge recombination. The present review reveals the significant recent progress on TiO2-based photocatalysts and decisive factors for photocatalytic performance via some critical strategies in Table 2.
The architectural structure design of photocatalysts, including crystal phases, oriented facets, and dispersion of TiO2 active sites, can play a crucial role in driving the CO2 reduction pathways and improving photocatalytic activity. Recently, tremendous achievements have been reported with respect to the influential structure of TiO2 rutile and anatase phases, morphology, size, various facets including {101}, {001}, {110}, and {111}, and alternative methods for TiO2 dispersion on porous materials to improve their intrinsic photoactivities. Moreover, the introduction of oxygen defects/vacancies and functional basic sites into the TiO2 surface can modify intermediate affinity and significantly improve CO2 adsorption and activation ability for CO2 reduction. Importantly, the utilization of cocatalysts such as noble metals, transition metals, and their oxides via plasmonic effects or/and heterojunctions improves visible-light harvesting, increases charge separation efficiency, and impedes electron-hole recombination. Furthermore, TiO2-based hybrid nanocomposites with advanced performances have emerged recently by manipulating several efficient approaches mentioned above combined with the hybridization with other materials (e.g., graphene, GO, rGO, g-C3N4 and its derivatives, sensitized-materials, and 2D MXene) to overcome the limitations of Ti and TiO2-based materials, exhibiting a CO2 photoreduction production rate of some thousands of micromoles per gram per hour.
Nonetheless, further improvement of CO2 adsorption, activation, and visible-light harvesting needs to be examined to enhance the sunlight harvesting yield and rate of production, which is, however, still far from practical applications. Insights on the reaction mechanism should also be investigated via in situ measurements with several powerful techniques for understanding and controlling reaction pathways, resulting in high selectivity of major products. In addition, most of studies reported the popular distribution of C1 products (CH4, HCOOH, CH3OH), especially CH4, from photoreduction of CO2 while there is a lack of the formation of C2+ products with higher price and demand in industry although a few studies exhibited the formation of C2H4, C2H6, and C3H8 as mentioned above (see Table 2). However, the production of C2+ products is much more complex, which requires a deeper understanding of mechanism and control. Furthermore, mass transfer and charge transfer should be comprehensively considered by examining the influence of operating parameters, structure of photocatalysts, and reactor design and operation to enable advanced yield and efficiency of sustainable CO2 photoreduction.

Author Contributions

Conceptualization, T.P.N., D.L.T.N. and S.Y.C.; methodology, T.P.N., D.L.T.N. and V.-H.N.; formal analysis, T.-H.L.; investigation, D.-V.N.V. and Q.T.T.; writing—review and editing, D.L.T.N., T.P.N., S.Y.C. and Q.V.L.; supervision, S.Y.C., S.Y.K. and Q.V.L.; project administration, S.-R.B. and S.Y.K.; funding acquisition, S.Y.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded in part by the Bio & Medical Technology Development Program (2018M3A9H1023141) of the National Research Foundation of Korea (NRF) funded by the Korean government and in part by a Korea Agency for Infrastructure Technology Advancement grant (17IFIP-B133622-01), funded by the Ministry of Land, Infrastructure and Transport and in part by Korea University Grant.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Meinshausen, M.; Meinshausen, N.; Hare, W.; Raper, S.C.B.; Frieler, K.; Knutti, R.; Frame, D.J.; Allen, M.R. Greenhouse-gas emission targets for limiting global warming to 2  °C. Nature 2009, 458, 1158–1162. [Google Scholar] [CrossRef]
  2. Li, X.; Yu, J.; Jaroniec, M.; Chen, X. Cocatalysts for Selective Photoreduction of CO2 into Solar Fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef] [PubMed]
  3. Solomon, S.; Plattner, G.-K.; Knutti, R.; Friedlingstein, P. Irreversible climate change due to carbon dioxide emissions. Proc. Natl. Acad. Sci. USA 2009, 106, 1704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Goeppert, A.; Czaun, M.; Jones, J.-P.; Surya Prakash, G.K.; Olah, G.A. Recycling of carbon dioxide to methanol and derived products—closing the loop. Chem. Soc. Rev. 2014, 43, 7995–8048. [Google Scholar] [CrossRef] [PubMed]
  5. 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]
  6. Marques, M.F.; Kim, D.H. From CO2 methanation to ambitious long-chain hydrocarbons: Alternative fuels paving the path to sustainability. Chem. Soc. Rev. 2019, 48, 205–259. [Google Scholar] [CrossRef]
  7. Kondratenko, E.V.; Mul, G.; Baltrusaitis, J.; Larrazábal, G.O.; Pérez-Ramírez, J. Status and perspectives of CO2 conversion into fuels and chemicals by catalytic, photocatalytic and electrocatalytic processes. Energy Environ. Sci. 2013, 6, 3112–3135. [Google Scholar] [CrossRef] [Green Version]
  8. Zhu, S.; Wang, D. Photocatalysis: Basic Principles, Diverse Forms of Implementations and Emerging Scientific Opportunities. Adv. Energy Mater. 2017, 7, 1700841. [Google Scholar] [CrossRef] [Green Version]
  9. Kibria, M.G.; Edwards, J.P.; Gabardo, C.M.; Dinh, C.-T.; Seifitokaldani, A.; Sinton, D.; Sargent, E.H. Electrochemical CO2 Reduction into Chemical Feedstocks: From Mechanistic Electrocatalysis Models to System Design. Adv. Mater. 2019, 31, 1807166. [Google Scholar] [CrossRef]
  10. Marques, M.F.; Nguyen, D.L.T.; Lee, J.-E.; Piao, H.; Choy, J.-H.; Hwang, Y.J.; Kim, D.H. Toward an Effective Control of the H2 to CO Ratio of Syngas through CO2 Electroreduction over Immobilized Gold Nanoparticles on Layered Titanate Nanosheets. ACS Catal. 2018, 8, 4364–4374. [Google Scholar] [CrossRef]
  11. Nguyen, D.L.T.; Kim, Y.; Hwang, Y.J.; Won, D.H. Progress in development of electrocatalyst for CO2 conversion to selective CO production. Carbon Energy 2019. [Google Scholar] [CrossRef] [Green Version]
  12. Zheng, T.; Jiang, K.; Wang, H. Recent Advances in Electrochemical CO2-to-CO Conversion on Heterogeneous Catalysts. Adv. Mater. 2018, 30, 1802066. [Google Scholar] [CrossRef]
  13. Nguyen, D.L.T.; Jee, M.S.; Won, D.H.; Jung, H.; Oh, H.-S.; Min, B.K.; Hwang, Y.J. Selective CO2 Reduction on Zinc Electrocatalyst: The Effect of Zinc Oxidation State Induced by Pretreatment Environment. ACS Sustain. Chem. Eng. 2017, 5, 11377–11386. [Google Scholar] [CrossRef]
  14. Hasani, A.; Tekalgne, M.; Le, Q.V.; Jang, H.W.; Kim, S.Y. Two-dimensional materials as catalysts for solar fuels: Hydrogen evolution reaction and CO2 reduction. J. Mater. Chem. A 2019, 7, 430–454. [Google Scholar] [CrossRef]
  15. Chae, S.Y.; Choi, J.Y.; Kim, Y.; Nguyen, D.L.T.; Joo, O.-S. Photoelectrochemical CO2 Reduction with a Rhenium Organometallic Redox Mediator at Semiconductor/Aqueous Liquid Junction Interfaces. Angew. Chem. 2019, 131, 16547–16551. [Google Scholar] [CrossRef]
  16. Shaner, M.R.; Atwater, H.A.; Lewis, N.S.; McFarland, E.W. A comparative technoeconomic analysis of renewable hydrogen production using solar energy. Energy Environ. Sci. 2016, 9, 2354–2371. [Google Scholar] [CrossRef] [Green Version]
  17. Lewis, N.S. Research opportunities to advance solar energy utilization. Science 2016, 351, aad1920. [Google Scholar] [CrossRef] [Green Version]
  18. Nguyen, V.-H.; Wu, J.C.S. Recent developments in the design of photoreactors for solar energy conversion from water splitting and CO2 reduction. Appl. Catal. A Gen. 2018, 550, 122–141. [Google Scholar] [CrossRef]
  19. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  20. Park, J.H.; Kim, S.; Bard, A.J. Novel Carbon-Doped TiO2 Nanotube Arrays with High Aspect Ratios for Efficient Solar Water Splitting. Nano Lett. 2006, 6, 24–28. [Google Scholar] [CrossRef]
  21. Khan, S.U.M.; Al-Shahry, M.; Ingler, W.B. Efficient Photochemical Water Splitting by a Chemically Modified n-TiO2. Science 2002, 297, 2243. [Google Scholar] [CrossRef] [PubMed]
  22. Gaya, U.I.; Abdullah, A.H. Heterogeneous photocatalytic degradation of organic contaminants over titanium dioxide: A review of fundamentals, progress and problems. J. Photochem. Photobiol. C 2008, 9, 1–12. [Google Scholar] [CrossRef]
  23. Ochiai, T.; Fujishima, A. Photoelectrochemical properties of TiO2 photocatalyst and its applications for environmental purification. J. Photochem. Photobiol. C 2012, 13, 247–262. [Google Scholar] [CrossRef]
  24. Epling, G.A.; Lin, C. Photoassisted bleaching of dyes utilizing TiO2 and visible light. Chemosphere 2002, 46, 561–570. [Google Scholar] [CrossRef]
  25. Bet-moushoul, E.; Mansourpanah, Y.; Farhadi, K.; Tabatabaei, M. TiO2 nanocomposite based polymeric membranes: A review on performance improvement for various applications in chemical engineering processes. Chem. Eng. J. 2016, 283, 29–46. [Google Scholar] [CrossRef]
  26. Hwang, Y.J.; Hahn, C.; Liu, B.; Yang, P. Photoelectrochemical Properties of TiO2 Nanowire Arrays: A Study of the Dependence on Length and Atomic Layer Deposition Coating. ACS Nano 2012, 6, 5060–5069. [Google Scholar] [CrossRef]
  27. Hwang, Y.J.; Boukai, A.; Yang, P. High Density n-Si/n-TiO2 Core/Shell Nanowire Arrays with Enhanced Photoactivity. Nano Lett. 2009, 9, 410–415. [Google Scholar] [CrossRef] [Green Version]
  28. 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]
  29. Dhakshinamoorthy, A.; Navalon, S.; Corma, A.; Garcia, H. Photocatalytic CO2 reduction by TiO2 and related titanium containing solids. Energy Environ. Sci. 2012, 5, 9217–9233. [Google Scholar] [CrossRef]
  30. Low, J.; Cheng, B.; Yu, J. Surface modification and enhanced photocatalytic CO2 reduction performance of TiO2: A review. Appl. Surf. Sci. 2017, 392, 658–686. [Google Scholar] [CrossRef]
  31. Shehzad, N.; Tahir, M.; Johari, K.; Murugesan, T.; Hussain, M. A critical review on TiO2 based photocatalytic CO2 reduction system: Strategies to improve efficiency. J. CO2 Util. 2018, 26, 98–122. [Google Scholar] [CrossRef]
  32. Lo, C.-C.; Hung, C.-H.; Yuan, C.-S.; Wu, J.-F. Photoreduction of carbon dioxide with H2 and H2O over TiO2 and ZrO2 in a circulated photocatalytic reactor. Sol. Energy Mater. Sol. Cells 2007, 91, 1765–1774. [Google Scholar] [CrossRef]
  33. Sarkar, A.; Gracia-Espino, E.; Wågberg, T.; Shchukarev, A.; Mohl, M.; Rautio, A.-R.; Pitkänen, O.; Sharifi, T.; Kordas, K.; Mikkola, J.-P. Photocatalytic reduction of CO2 with H2O over modified TiO2 nanofibers: Understanding the reduction pathway. Nano Res. 2016, 9, 1956–1968. [Google Scholar] [CrossRef]
  34. Marszewski, M.; Cao, S.; Yu, J.; Jaroniec, M. Semiconductor-based photocatalytic CO2 conversion. Mater. Horiz. 2015, 2, 261–278. [Google Scholar] [CrossRef]
  35. Wang, W.-N.; An, W.-J.; Ramalingam, B.; Mukherjee, S.; Niedzwiedzki, D.M.; Gangopadhyay, S.; Biswas, P. Size and Structure Matter: Enhanced CO2 Photoreduction Efficiency by Size-Resolved Ultrafine Pt Nanoparticles on TiO2 Single Crystals. J. Am. Chem. Soc. 2012, 134, 11276–11281. [Google Scholar] [CrossRef]
  36. Xu, H.; Ouyang, S.; Liu, L.; Wang, D.; Kako, T.; Ye, J. Porous-structured Cu2O/TiO2 nanojunction material toward efficient CO2 photoreduction. Nanotechnology 2014, 25, 165402. [Google Scholar] [CrossRef]
  37. An, X.; Li, K.; Tang, J. Cu2O/Reduced Graphene Oxide Composites for the Photocatalytic Conversion of CO2. ChemSusChem 2014, 7, 1086–1093. [Google Scholar] [CrossRef] [Green Version]
  38. Hsu, H.-C.; Shown, I.; Wei, H.-Y.; Chang, Y.-C.; Du, H.-Y.; Lin, Y.-G.; Tseng, C.-A.; Wang, C.-H.; Chen, L.-C.; Lin, Y.-C.; et al. Graphene oxide as a promising photocatalyst for CO2 to methanol conversion. Nanoscale 2013, 5, 262–268. [Google Scholar] [CrossRef]
  39. Yu, J.; Jin, J.; Cheng, B.; Jaroniec, M. A noble metal-free reduced graphene oxide—CdS nanorod composite for the enhanced visible-light photocatalytic reduction of CO2 to solar fuel. J. Mater. Chem. A 2014, 2, 3407–3416. [Google Scholar] [CrossRef]
  40. In, S.-I.; Vaughn Ii, D.D.; Schaak, R.E. Hybrid CuO-TiO2−xNx Hollow Nanocubes for Photocatalytic Conversion of CO2 into Methane under Solar Irradiation. Angew. Chem. Int. Ed. 2012, 51, 3915–3918. [Google Scholar] [CrossRef]
  41. He, H.; Zapol, P.; Curtiss, L.A. Computational screening of dopants for photocatalytic two-electron reduction of CO2 on anatase (101) surfaces. Energy Environ. Sci. 2012, 5, 6196–6205. [Google Scholar] [CrossRef]
  42. Xiong, Z.; Lei, Z.; Li, Y.; Dong, L.; Zhao, Y.; Zhang, J. A review on modification of facet-engineered TiO2 for photocatalytic CO2 reduction. J. Photochem. Photobiol. C Photochem. Rev. 2018, 36, 24–47. [Google Scholar] [CrossRef]
  43. Wang, Y.; He, D.; Chen, H.; Wang, D. Catalysts in electro-, photo- and photoelectrocatalytic CO2 reduction reactions. J. Photochem. Photobiol. C Photochem. Rev. 2019, 40, 117–149. [Google Scholar] [CrossRef]
  44. Kawahara, T.; Konishi, Y.; Tada, H.; Tohge, N.; Nishii, J.; Ito, S. A Patterned TiO2(Anatase)/TiO2(Rutile) Bilayer-Type Photocatalyst: Effect of the Anatase/Rutile Junction on the Photocatalytic Activity. Angew. Chem. 2002, 114, 2935–2937. [Google Scholar] [CrossRef]
  45. Ohno, T.; Tokieda, K.; Higashida, S.; Matsumura, M. Synergism between rutile and anatase TiO2 particles in photocatalytic oxidation of naphthalene. Appl. Catal. A Gen. 2003, 244, 383–391. [Google Scholar] [CrossRef]
  46. Kandiel, T.A.; Dillert, R.; Feldhoff, A.; Bahnemann, D.W. Direct Synthesis of Photocatalytically Active Rutile TiO2 Nanorods Partly Decorated with Anatase Nanoparticles. J. Phys. Chem. C 2010, 114, 4909–4915. [Google Scholar] [CrossRef]
  47. Tan, L.-L.; Ong, W.-J.; Chai, S.-P.; Mohamed, A.R. Visible-light-activated oxygen-rich TiO2 as next generation photocatalyst: Importance of annealing temperature on the photoactivity toward reduction of carbon dioxide. Chem. Eng. J. 2016, 283, 1254–1263. [Google Scholar] [CrossRef]
  48. Reñones, P.; Moya, A.; Fresno, F.; Collado, L.; Vilatela, J.J.; de la Peña O’Shea, V.A. Hierarchical TiO2 nanofibres as photocatalyst for CO2 reduction: Influence of morphology and phase composition on catalytic activity. J. CO2 Util. 2016, 15, 24–31. [Google Scholar] [CrossRef] [Green Version]
  49. Liu, S.; Yu, J.; Jaroniec, M. Tunable Photocatalytic Selectivity of Hollow TiO2 Microspheres Composed of Anatase Polyhedra with Exposed {001} Facets. J. Am. Chem. Soc. 2010, 132, 11914–11916. [Google Scholar] [CrossRef]
  50. He, H.; Zapol, P.; Curtiss, L.A. A Theoretical Study of CO2 Anions on Anatase (101) Surface. J. Phys. Chem. C 2010, 114, 21474–21481. [Google Scholar] [CrossRef]
  51. Han, X.; Kuang, Q.; Jin, M.; Xie, Z.; Zheng, L. Synthesis of Titania Nanosheets with a High Percentage of Exposed (001) Facets and Related Photocatalytic Properties. J. Am. Chem. Soc. 2009, 131, 3152–3153. [Google Scholar] [CrossRef] [PubMed]
  52. Yu, J.; Low, J.; Xiao, W.; Zhou, P.; Jaroniec, M. Enhanced Photocatalytic CO2-Reduction Activity of Anatase TiO2 by Coexposed {001} and {101} Facets. J. Am. Chem. Soc. 2014, 136, 8839–8842. [Google Scholar] [CrossRef] [PubMed]
  53. Xu, Q.; Yu, J.; Zhang, J.; Zhang, J.; Liu, G. Cubic anatase TiO2 nanocrystals with enhanced photocatalytic CO2 reduction activity. Chem. Commun. 2015, 51, 7950–7953. [Google Scholar] [CrossRef] [PubMed]
  54. He, Z.; Wen, L.; Wang, D.; Xue, Y.; Lu, Q.; Wu, C.; Chen, J.; Song, S. Photocatalytic Reduction of CO2 in Aqueous Solution on Surface-Fluorinated Anatase TiO2 Nanosheets with Exposed {001} Facets. Energy Fuels 2014, 28, 3982–3993. [Google Scholar] [CrossRef]
  55. Jiao, W.; Wang, L.; Liu, G.; Lu, G.Q.; Cheng, H.-M. Hollow Anatase TiO2 Single Crystals and Mesocrystals with Dominant {101} Facets for Improved Photocatalysis Activity and Tuned Reaction Preference. ACS Catal. 2012, 2, 1854–1859. [Google Scholar] [CrossRef]
  56. Kar, P.; Zeng, S.; Zhang, Y.; Vahidzadeh, E.; Manuel, A.; Kisslinger, R.; Alam, K.M.; Thakur, U.K.; Mahdi, N.; Kumar, P.; et al. High rate CO2 photoreduction using flame annealed TiO2 nanotubes. Appl. Catal. B 2019, 243, 522–536. [Google Scholar] [CrossRef]
  57. Liu, J.; Liu, B.; Ren, Y.; Yuan, Y.; Zhao, H.; Yang, H.; Liu, S. Hydrogenated nanotubes/nanowires assembled from TiO2 nanoflakes with exposed {111} facets: Excellent photo-catalytic CO2 reduction activity and charge separation mechanism between (111) and ( 1 ¯ 1 ¯ 1 ¯ ) polar surfaces. J. Mater. Chem. A 2019, 7, 14761–14775. [Google Scholar] [CrossRef]
  58. Anpo, M.; Yamashita, H.; Ichihashi, Y.; Fujii, Y.; Honda, M. Photocatalytic Reduction of CO2 with H2O on Titanium Oxides Anchored within Micropores of Zeolites:  Effects of the Structure of the Active Sites and the Addition of Pt. J. Phys. Chem. B 1997, 101, 2632–2636. [Google Scholar] [CrossRef]
  59. Anpo, M.; Yamashita, H.; Ikeue, K.; Fujii, Y.; Zhang, S.G.; Ichihashi, Y.; Park, D.R.; Suzuki, Y.; Koyano, K.; Tatsumi, T. Photocatalytic reduction of CO2 with H2O on Ti-MCM-41 and Ti-MCM-48 mesoporous zeolite catalysts. Catal. Today 1998, 44, 327–332. [Google Scholar] [CrossRef]
  60. Anpo, M. Photocatalytic reduction of CO2 with H2O on highly dispersed Ti-oxide catalysts as a model of artificial photosynthesis. J. CO2 Util. 2013, 1, 8–17. [Google Scholar] [CrossRef]
  61. Hwang, J.-S.; Chang, J.-S.; Park, S.-E.; Ikeue, K.; Anpo, M. Photoreduction of Carbondioxide on Surface Functionalized Nanoporous Catalysts. Top. Catal. 2005, 35, 311–319. [Google Scholar] [CrossRef]
  62. Xie, K.; Umezawa, N.; Zhang, N.; Reunchan, P.; Zhang, Y.; Ye, J. Self-doped SrTiO3−δ photocatalyst with enhanced activity for artificial photosynthesis under visible light. Energy Environ. Sci. 2011, 4, 4211–4219. [Google Scholar] [CrossRef]
  63. Liu, L.; Zhao, H.; Andino, J.M.; Li, Y. Photocatalytic CO2 Reduction with H2O on TiO2 Nanocrystals: Comparison of Anatase, Rutile, and Brookite Polymorphs and Exploration of Surface Chemistry. ACS Catal. 2012, 2, 1817–1828. [Google Scholar] [CrossRef]
  64. Lee, J.; Sorescu, D.C.; Deng, X. Electron-Induced Dissociation of CO2 on TiO2(110). J. Am. Chem. Soc. 2011, 133, 10066–10069. [Google Scholar] [CrossRef] [PubMed]
  65. Liu, J.-Y.; Gong, X.-Q.; Alexandrova, A.N. Mechanism of CO2 Photocatalytic Reduction to Methane and Methanol on Defected Anatase TiO2 (101): A Density Functional Theory Study. J. Phys. Chem. C 2019, 123, 3505–3511. [Google Scholar] [CrossRef]
  66. Liu, L.; Jiang, Y.; Zhao, H.; Chen, J.; Cheng, J.; Yang, K.; Li, Y. Engineering Coexposed {001} and {101} Facets in Oxygen-Deficient TiO2 Nanocrystals for Enhanced CO2 Photoreduction under Visible Light. ACS Catal. 2016, 6, 1097–1108. [Google Scholar] [CrossRef]
  67. Fang, W.; Khrouz, L.; Zhou, Y.; Shen, B.; Dong, C.; Xing, M.; Mishra, S.; Daniele, S.; Zhang, J. Reduced {001}-TiO2−x photocatalysts: Noble-metal-free CO2 photoreduction for selective CH4 evolution. Phys. Chem. Chem. Phys. 2017, 19, 13875–13881. [Google Scholar] [CrossRef] [Green Version]
  68. Liu, L.; Zhao, C.; Li, Y. Spontaneous Dissociation of CO2 to CO on Defective Surface of Cu(I)/TiO2–x Nanoparticles at Room Temperature. J. Phys. Chem. C 2012, 116, 7904–7912. [Google Scholar] [CrossRef]
  69. Liu, L.; Gao, F.; Zhao, H.; Li, Y. Tailoring Cu valence and oxygen vacancy in Cu/TiO2 catalysts for enhanced CO2 photoreduction efficiency. Appl. Catal. B 2013, 134–135, 349–358. [Google Scholar] [CrossRef]
  70. Zhao, J.; Li, Y.; Zhu, Y.; Wang, Y.; Wang, C. Enhanced CO2 photoreduction activity of black TiO2−coated Cu nanoparticles under visible light irradiation: Role of metallic Cu. Appl. Catal. A Gen. 2016, 510, 34–41. [Google Scholar] [CrossRef]
  71. Pham, T.-D.; Lee, B.-K. Novel capture and photocatalytic conversion of CO2 into solar fuels by metals co-doped TiO2 deposited on PU under visible light. Appl. Catal. A Gen. 2017, 529, 40–48. [Google Scholar] [CrossRef]
  72. Pham, T.-D.; Lee, B.-K. Novel photocatalytic activity of Cu@V co-doped TiO2/PU for CO2 reduction with H2O vapor to produce solar fuels under visible light. J. Catal. 2017, 345, 87–95. [Google Scholar] [CrossRef]
  73. Meng, X.; Ouyang, S.; Kako, T.; Li, P.; Yu, Q.; Wang, T.; Ye, J. Photocatalytic CO2 conversion over alkali modified TiO2 without loading noble metal cocatalyst. Chem. Commun. 2014, 50, 11517–11519. [Google Scholar] [CrossRef] [PubMed]
  74. Xie, S.; Wang, Y.; Zhang, Q.; Fan, W.; Deng, W.; Wang, Y. Photocatalytic reduction of CO2 with H2O: Significant enhancement of the activity of Pt–TiO2 in CH4 formation by addition of MgO. Chem. Commun. 2013, 49, 2451–2453. [Google Scholar] [CrossRef]
  75. Xie, S.; Wang, Y.; Zhang, Q.; Deng, W.; Wang, Y. MgO- and Pt-Promoted TiO2 as an Efficient Photocatalyst for the Preferential Reduction of Carbon Dioxide in the Presence of Water. ACS Catal. 2014, 4, 3644–3653. [Google Scholar] [CrossRef]
  76. Feng, X.; Pan, F.; Zhao, H.; Deng, W.; Zhang, P.; Zhou, H.-C.; Li, Y. Atomic layer deposition enabled MgO surface coating on porous TiO2 for improved CO2 photoreduction. Appl. Catal. B 2018, 238, 274–283. [Google Scholar] [CrossRef]
  77. Liao, Y.; Cao, S.-W.; Yuan, Y.; Gu, Q.; Zhang, Z.; Xue, C. Efficient CO2 Capture and Photoreduction by Amine-Functionalized TiO2. Chem. Eur. J. 2014, 20, 10220–10222. [Google Scholar] [CrossRef]
  78. Liu, S.; Xia, J.; Yu, J. Amine-Functionalized Titanate Nanosheet-Assembled Yolk@Shell Microspheres for Efficient Cocatalyst-Free Visible-Light Photocatalytic CO2 Reduction. ACS Appl. Mater. Interfaces 2015, 7, 8166–8175. [Google Scholar] [CrossRef]
  79. Tanaka, K.; White, J.M. Dissociative adsorption of carbon dioxide on oxidized and reduced platinum/titanium dioxide. J. Phys. Chem. 1982, 86, 3977–3980. [Google Scholar] [CrossRef]
  80. Raskó, J. FTIR study of the photoinduced dissociation of CO2 on titania-supported noble metals. Catal. Lett. 1998, 56, 11–15. [Google Scholar] [CrossRef]
  81. Lan, Y.; Xie, Y.; Chen, J.; Hu, Z.; Cui, D. Selective photocatalytic CO2 reduction on copper–titanium dioxide: A study of the relationship between CO production and H2 suppression. Chem. Commun. 2019, 55, 8068–8071. [Google Scholar] [CrossRef] [PubMed]
  82. Li, N.; Liu, M.; Yang, B.; Shu, W.; Shen, Q.; Liu, M.; Zhou, J. Enhanced Photocatalytic Performance toward CO2 Hydrogenation over Nanosized TiO2-Loaded Pd under UV Irradiation. J. Phys. Chem. C 2017, 121, 2923–2932. [Google Scholar] [CrossRef]
  83. Liu, Y.; Miao, C.; Yang, P.; He, Y.; Feng, J.; Li, D. Synergetic promotional effect of oxygen vacancy-rich ultrathin TiO2 and photochemical induced highly dispersed Pt for photoreduction of CO2 with H2O. Appl. Catal. B 2019, 244, 919–930. [Google Scholar] [CrossRef]
  84. Gonell, F.; Puga, A.V.; Julián-López, B.; García, H.; Corma, A. Copper-doped titania photocatalysts for simultaneous reduction of CO2 and production of H2 from aqueous sulfide. Appl. Catal. B 2016, 180, 263–270. [Google Scholar] [CrossRef] [Green Version]
  85. Fang, B.; Xing, Y.; Bonakdarpour, A.; Zhang, S.; Wilkinson, D.P. Hierarchical CuO–TiO2 Hollow Microspheres for Highly Efficient Photodriven Reduction of CO2 to CH4. ACS Sustain. Chem. Eng. 2015, 3, 2381–2388. [Google Scholar] [CrossRef]
  86. Jiang, Z.; Sun, W.; Miao, W.; Yuan, Z.; Yang, G.; Kong, F.; Yan, T.; Chen, J.; Huang, B.; An, C.; et al. Living Atomically Dispersed Cu Ultrathin TiO2 Nanosheet CO2 Reduction Photocatalyst. Adv. Sci. 2019, 6, 1900289. [Google Scholar] [CrossRef] [Green Version]
  87. Zhai, Q.; Xie, S.; Fan, W.; Zhang, Q.; Wang, Y.; Deng, W.; Wang, Y. Photocatalytic Conversion of Carbon Dioxide with Water into Methane: Platinum and Copper(I) Oxide Co-catalysts with a Core–Shell Structure. Angew. Chem. Int. Ed. 2013, 52, 5776–5779. [Google Scholar] [CrossRef]
  88. Tan, D.; Zhang, J.; Shi, J.; Li, S.; Zhang, B.; Tan, X.; Zhang, F.; Liu, L.; Shao, D.; Han, B. Photocatalytic CO2 Transformation to CH4 by Ag/Pd Bimetals Supported on N-Doped TiO2 Nanosheet. ACS Appl. Mater. Interfaces 2018, 10, 24516–24522. [Google Scholar] [CrossRef]
  89. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. TiO2–MnOx–Pt Hybrid Multiheterojunction Film Photocatalyst with Enhanced Photocatalytic CO2-Reduction Activity. ACS Appl. Mater. Interfaces 2019, 11, 5581–5589. [Google Scholar] [CrossRef]
  90. Kang, Q.; Wang, T.; Li, P.; Liu, L.; Chang, K.; Li, M.; Ye, J. Photocatalytic Reduction of Carbon Dioxide by Hydrous Hydrazine over Au–Cu Alloy Nanoparticles Supported on SrTiO3/TiO2 Coaxial Nanotube Arrays. Angew. Chem. Int. Ed. 2015, 54, 841–845. [Google Scholar] [CrossRef]
  91. Kim, S.M.; Lee, S.W.; Moon, S.Y.; Park, J.Y. The effect of hot electrons and surface plasmons on heterogeneous catalysis. J. Phys. Condens. Matter 2016, 28, 254002. [Google Scholar] [CrossRef] [PubMed]
  92. Choi, C.H.; Chung, K.; Nguyen, T.-T.H.; Kim, D.H. Plasmon-Mediated Electrocatalysis for Sustainable Energy: From Electrochemical Conversion of Different Feedstocks to Fuel Cell Reactions. ACS Energy Lett. 2018, 3, 1415–1433. [Google Scholar] [CrossRef]
  93. Jang, Y.H.; Jang, Y.J.; Kim, S.; Quan, L.N.; Chung, K.; Kim, D.H. Plasmonic Solar Cells: From Rational Design to Mechanism Overview. Chem. Rev. 2016, 116, 14982–15034. [Google Scholar] [CrossRef] [PubMed]
  94. 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]
  95. Zhang, Z.; Wang, Z.; Cao, S.-W.; Xue, C. Au/Pt Nanoparticle-Decorated TiO2 Nanofibers with Plasmon-Enhanced Photocatalytic Activities for Solar-to-Fuel Conversion. J. Phys. Chem. C 2013, 117, 25939–25947. [Google Scholar] [CrossRef]
  96. Tahir, M. Synergistic effect in MMT-dispersed Au/TiO2 monolithic nanocatalyst for plasmon-absorption and metallic interband transitions dynamic CO2 photo-reduction to CO. Appl. Catal. B 2017, 219, 329–343. [Google Scholar] [CrossRef]
  97. Neaţu, Ş.; Maciá-Agulló, J.A.; Concepción, P.; Garcia, H. Gold–Copper Nanoalloys Supported on TiO2 as Photocatalysts for CO2 Reduction by Water. J. Am. Chem. Soc. 2014, 136, 15969–15976. [Google Scholar] [CrossRef]
  98. Jiang, Z.; Zhang, X.; Yuan, Z.; Chen, J.; Huang, B.; Dionysiou, D.D.; Yang, G. Enhanced photocatalytic CO2 reduction via the synergistic effect between Ag and activated carbon in TiO2/AC-Ag ternary composite. Chem. Eng. J. 2018, 348, 592–598. [Google Scholar] [CrossRef]
  99. Liang, Y.T.; Vijayan, B.K.; Gray, K.A.; Hersam, M.C. Minimizing Graphene Defects Enhances Titania Nanocomposite-Based Photocatalytic Reduction of CO2 for Improved Solar Fuel Production. Nano Lett. 2011, 11, 2865–2870. [Google Scholar] [CrossRef]
  100. Tu, W.; Zhou, Y.; Liu, Q.; Tian, Z.; Gao, J.; Chen, X.; Zhang, H.; Liu, J.; Zou, Z. Robust Hollow Spheres Consisting of Alternating Titania Nanosheets and Graphene Nanosheets with High Photocatalytic Activity for CO2 Conversion into Renewable Fuels. Adv. Funct. Mater. 2012, 22, 1215–1221. [Google Scholar] [CrossRef]
  101. Wang, W.; Xu, D.; Cheng, B.; Yu, J.; Jiang, C. Hybrid carbon@TiO2 hollow spheres with enhanced photocatalytic CO2 reduction activity. J. Mater. Chem. A 2017, 5, 5020–5029. [Google Scholar] [CrossRef]
  102. Tu, W.; Zhou, Y.; Liu, Q.; Yan, S.; Bao, S.; Wang, X.; Xiao, M.; Zou, Z. An In Situ Simultaneous Reduction-Hydrolysis Technique for Fabrication of TiO2-Graphene 2D Sandwich-Like Hybrid Nanosheets: Graphene-Promoted Selectivity of Photocatalytic-Driven Hydrogenation and Coupling of CO2 into Methane and Ethane. Adv. Funct. Mater. 2013, 23, 1743–1749. [Google Scholar] [CrossRef]
  103. Jin, B.; Yao, G.; Jin, F.; Hu, Y.H. Photocatalytic conversion of CO2 over C3N4-based catalysts. Catal. Today 2018, 316, 149–154. [Google Scholar] [CrossRef]
  104. Wada, K.; Ranasinghe, C.S.K.; Kuriki, R.; Yamakata, A.; Ishitani, O.; Maeda, K. Interfacial Manipulation by Rutile TiO2 Nanoparticles to Boost CO2 Reduction into CO on a Metal-Complex/Semiconductor Hybrid Photocatalyst. ACS Appl. Mater. Interfaces 2017, 9, 23869–23877. [Google Scholar] [CrossRef] [PubMed]
  105. Zhao, Y.; Wei, Y.; Wu, X.; Zheng, H.; Zhao, Z.; Liu, J.; Li, J. Graphene-wrapped Pt/TiO2 photocatalysts with enhanced photogenerated charges separation and reactant adsorption for high selective photoreduction of CO2 to CH4. Appl. Catal. B 2018, 226, 360–372. [Google Scholar] [CrossRef]
  106. Crake, A.; Christoforidis, K.C.; Godin, R.; Moss, B.; Kafizas, A.; Zafeiratos, S.; Durrant, J.R.; Petit, C. Titanium dioxide/carbon nitride nanosheet nanocomposites for gas phase CO2 photoreduction under UV-visible irradiation. Appl. Catal. B 2019, 242, 369–378. [Google Scholar] [CrossRef]
  107. Brunetti, A.; Pomilla, F.R.; Marcì, G.; Garcia-Lopez, E.I.; Fontananova, E.; Palmisano, L.; Barbieri, G. CO2 reduction by C3N4-TiO2 Nafion photocatalytic membrane reactor as a promising environmental pathway to solar fuels. Appl. Catal. B 2019, 255, 117779. [Google Scholar] [CrossRef]
  108. 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]
  109. Xu, F.; Zhang, J.; Zhu, B.; Yu, J.; Xu, J. CuInS2 sensitized TiO2 hybrid nanofibers for improved photocatalytic CO2 reduction. Appl. Catal. B 2018, 230, 194–202. [Google Scholar] [CrossRef]
  110. Low, J.; Zhang, L.; Tong, T.; Shen, B.; Yu, J. TiO2/MXene Ti3C2 composite with excellent photocatalytic CO2 reduction activity. J. Catal. 2018, 361, 255–266. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of photoreduction of CO2 into fuels on TiO2-based photocatalysts.
Figure 1. Schematic illustration of photoreduction of CO2 into fuels on TiO2-based photocatalysts.
Nanomaterials 10 00337 g001
Figure 2. (a) conductive band (CB), valence band (VB), and bandgap energies of several semiconductors summarized from the references [34,35,36,37,38,39,40]. Reproduced with permission from [34]. Copyright 2015, Royal Society of Chemistry. (b) Configurations of CO2 molecule on TiO2 surface. Reproduced with permission from [41]. Copyright 2012, American Chemical Society.
Figure 2. (a) conductive band (CB), valence band (VB), and bandgap energies of several semiconductors summarized from the references [34,35,36,37,38,39,40]. Reproduced with permission from [34]. Copyright 2015, Royal Society of Chemistry. (b) Configurations of CO2 molecule on TiO2 surface. Reproduced with permission from [41]. Copyright 2012, American Chemical Society.
Nanomaterials 10 00337 g002aNanomaterials 10 00337 g002b
Figure 3. (a) Heterojunctions induced from rutile and anatase crystal phases of TiO2. (b) TEM images of anatase {101} and rutile phases {110}. (a,b) Reproduced with permission [47]. Copyright 2016, Royal Society of Chemistry. (c) Schematic illustration of the spatial separation of redox sites on TiO2 prepared by alternating HF amounts. (d) Comparison of the CH4 production rate of P25 and the TiO2 samples prepared by varying HF amounts. (c,d) Reproduced with permission [52]. Copyright 2014, American Chemical Society. (e) Transformation of the anatase phase of TiO2 nanotubes into the rutile phase with {110} as the dominant plane using the flame annealing method. (e) Reproduced with permission [56]. Copyright 2019, Elsevier.
Figure 3. (a) Heterojunctions induced from rutile and anatase crystal phases of TiO2. (b) TEM images of anatase {101} and rutile phases {110}. (a,b) Reproduced with permission [47]. Copyright 2016, Royal Society of Chemistry. (c) Schematic illustration of the spatial separation of redox sites on TiO2 prepared by alternating HF amounts. (d) Comparison of the CH4 production rate of P25 and the TiO2 samples prepared by varying HF amounts. (c,d) Reproduced with permission [52]. Copyright 2014, American Chemical Society. (e) Transformation of the anatase phase of TiO2 nanotubes into the rutile phase with {110} as the dominant plane using the flame annealing method. (e) Reproduced with permission [56]. Copyright 2019, Elsevier.
Nanomaterials 10 00337 g003
Figure 4. (a) SEM and (b) TEM images of the catalyst prepared in the TOBT-CH3COOH-H2O-H2SO4 system. (c) HRTEM image and (d) the corresponding FFT pattern from the box in (b). (e) Structural illustration of the TiO2 nanotubes/nanowires constructed with nanoflakes with exposed {111} crystal facets. (f) The schematic illustrations of charge separation under Es and photocatalytic reaction. (g) Photocatalytic CO2 reduction activity and stability in recycling the photocatalyst. (ag) Reproduced with permission [57]. Copyright 2019, Royal Society of Chemistry.
Figure 4. (a) SEM and (b) TEM images of the catalyst prepared in the TOBT-CH3COOH-H2O-H2SO4 system. (c) HRTEM image and (d) the corresponding FFT pattern from the box in (b). (e) Structural illustration of the TiO2 nanotubes/nanowires constructed with nanoflakes with exposed {111} crystal facets. (f) The schematic illustrations of charge separation under Es and photocatalytic reaction. (g) Photocatalytic CO2 reduction activity and stability in recycling the photocatalyst. (ag) Reproduced with permission [57]. Copyright 2019, Royal Society of Chemistry.
Nanomaterials 10 00337 g004
Figure 5. (a) Schematic representation showing Vo, OHb, and CO2 molecule adsorbed at a Vo site on reduced TiO2. (b) Scanning Tunneling Microscope (STM) image of the TiO2 {110} surface after adsorption of CO2 at 55 K. (a,b) Reproduced with permission from [64]. Copyright 2011, American Chemical Society. (ce) Configurations of adsorbed CO2 on Ov site: (c) CO2(H); (d) CO2(V); (e): Adsorption and dissociation of CO2 on the oxygen vacancy site. The Ti, O, and O in CO2 and C are shown in silver, red, pink, and gray colors, respectively. (ce) Reproduced with permission from [65]. Copyright 2019, American Chemical Society.
Figure 5. (a) Schematic representation showing Vo, OHb, and CO2 molecule adsorbed at a Vo site on reduced TiO2. (b) Scanning Tunneling Microscope (STM) image of the TiO2 {110} surface after adsorption of CO2 at 55 K. (a,b) Reproduced with permission from [64]. Copyright 2011, American Chemical Society. (ce) Configurations of adsorbed CO2 on Ov site: (c) CO2(H); (d) CO2(V); (e): Adsorption and dissociation of CO2 on the oxygen vacancy site. The Ti, O, and O in CO2 and C are shown in silver, red, pink, and gray colors, respectively. (ce) Reproduced with permission from [65]. Copyright 2019, American Chemical Society.
Nanomaterials 10 00337 g005
Figure 6. (a) Generation of CO on oxygen-deficient blue TiO2 nanocrystals with coexposed {101}-{001} facets in comparison with some other photocatalysts. Reproduced with permission from [66]. Copyright 2016, American Chemical Society. (b) A rational mechanism for spontaneous dissociation of CO2 on a defective Cu(I)/TiO2−x catalyst even in the dark. Reproduced with permission from [68]. Copyright 2012, American Chemical Society. (c,d) TEM images of deposition of five atomic layers of MgO on porous TiO2. (e) CO production rate of different atomic layer deposited MgO surface coating layers on porous TiO2. (ce) Reproduced with permission from [76]. Copyright 2018, Elsevier.
Figure 6. (a) Generation of CO on oxygen-deficient blue TiO2 nanocrystals with coexposed {101}-{001} facets in comparison with some other photocatalysts. Reproduced with permission from [66]. Copyright 2016, American Chemical Society. (b) A rational mechanism for spontaneous dissociation of CO2 on a defective Cu(I)/TiO2−x catalyst even in the dark. Reproduced with permission from [68]. Copyright 2012, American Chemical Society. (c,d) TEM images of deposition of five atomic layers of MgO on porous TiO2. (e) CO production rate of different atomic layer deposited MgO surface coating layers on porous TiO2. (ce) Reproduced with permission from [76]. Copyright 2018, Elsevier.
Nanomaterials 10 00337 g006
Figure 7. (a) Schematic illustration of photocatalytic reduction of CO2 on a Pt-TiO2 photocatalyst, exhibiting fast transfer of photogenerated electrons inside the highly oriented TiO2 single crystals into the Pt sites where the redox reaction converts CO2 into CH4. Reproduced with permission from [35]. Copyright 2017, American Chemical Society. (b) HRTEM images and STEM-HAADF-mapping of well-dispersed 1.1-nm Pt nanoparticles on oxygen vacancy-rich ultrathin TiO2. (c) Proposed mechanisms of the ultrathin TiO2-supported highly dispersed Pt nanoparticles for photoreduction of CO2 with H2O. (b,c) Reproduced with permission from [83]. Copyright 2019, Elsevier.
Figure 7. (a) Schematic illustration of photocatalytic reduction of CO2 on a Pt-TiO2 photocatalyst, exhibiting fast transfer of photogenerated electrons inside the highly oriented TiO2 single crystals into the Pt sites where the redox reaction converts CO2 into CH4. Reproduced with permission from [35]. Copyright 2017, American Chemical Society. (b) HRTEM images and STEM-HAADF-mapping of well-dispersed 1.1-nm Pt nanoparticles on oxygen vacancy-rich ultrathin TiO2. (c) Proposed mechanisms of the ultrathin TiO2-supported highly dispersed Pt nanoparticles for photoreduction of CO2 with H2O. (b,c) Reproduced with permission from [83]. Copyright 2019, Elsevier.
Nanomaterials 10 00337 g007
Figure 8. (a) Schematic illustrations of synthesis of hybrid TiO2-MnOx-Pt composite by photodeposition of MnOx nanosheets and Pt nanoparticles on TiO2. (b) Schematic diagram of proposed photocatalytic CO2 reduction mechanism and modified bandgaps of the ternary TiO2-MnOx-Pt photocatalyst before contact and after contact and under irradiation. Reproduced with permission from [89]. Copyright 2018, American Chemical Society.
Figure 8. (a) Schematic illustrations of synthesis of hybrid TiO2-MnOx-Pt composite by photodeposition of MnOx nanosheets and Pt nanoparticles on TiO2. (b) Schematic diagram of proposed photocatalytic CO2 reduction mechanism and modified bandgaps of the ternary TiO2-MnOx-Pt photocatalyst before contact and after contact and under irradiation. Reproduced with permission from [89]. Copyright 2018, American Chemical Society.
Nanomaterials 10 00337 g008
Figure 9. Proposed schematic illustration of the mechanism of photocatalytic CO2 reduction under UV and visible light for Au–Cu alloy NPs decorated on TiO2 as photocatalysts, demonstrating the crucial role of the irradiation wavelength range on product distribution. Reproduced with permission from [96]. Copyright 2014, American Chemical Society.
Figure 9. Proposed schematic illustration of the mechanism of photocatalytic CO2 reduction under UV and visible light for Au–Cu alloy NPs decorated on TiO2 as photocatalysts, demonstrating the crucial role of the irradiation wavelength range on product distribution. Reproduced with permission from [96]. Copyright 2014, American Chemical Society.
Nanomaterials 10 00337 g009
Figure 10. (a) Scheme of the synthesis procedure used to produce TiO2/CNNS composites. (b) Transient absorption spectroscopy spectra used to study the photoexcitation processes in powdered samples at different delay times of TiO2 nanosheets (TiO2-NS) following photoexcitation. Initial spectra at 1 μs for (c) TiO2-NS, (d) Ti-NS/CN (synthesized in the presence of HF), and (e) Ti-ISO/CN (synthesized in the absence of HF). The hole contribution is shaded in blue, and the electron contribution is shaded in red. The sum of the two is indicated as the total (green line) compared with the experimental values shown in the black line and squares. (f) Photocatalytic CO2 reduction reaction pathway for the TiO2/CNNS nanocomposites improved by hole transfer from TiO2-NS to CNNS. Reproduced with permission from [106]. Copyright 2019, Elsevier.
Figure 10. (a) Scheme of the synthesis procedure used to produce TiO2/CNNS composites. (b) Transient absorption spectroscopy spectra used to study the photoexcitation processes in powdered samples at different delay times of TiO2 nanosheets (TiO2-NS) following photoexcitation. Initial spectra at 1 μs for (c) TiO2-NS, (d) Ti-NS/CN (synthesized in the presence of HF), and (e) Ti-ISO/CN (synthesized in the absence of HF). The hole contribution is shaded in blue, and the electron contribution is shaded in red. The sum of the two is indicated as the total (green line) compared with the experimental values shown in the black line and squares. (f) Photocatalytic CO2 reduction reaction pathway for the TiO2/CNNS nanocomposites improved by hole transfer from TiO2-NS to CNNS. Reproduced with permission from [106]. Copyright 2019, Elsevier.
Nanomaterials 10 00337 g010
Figure 11. SEM images of (a,b) pristine TiO2, (c,d) sensitized CuInS2/TiO2. TEM image of (e) pristine TiO2 and (f) sensitized CuInS2/TiO2. (g) energy dispersive X-ray (EDX) element mapping of Ti, O, Cu, In, and S. (h) Scheme of the illustrated charge transfer and separation in the optimum sensitized CuInS2/TiO2 photocatalyst under simulated sunlight irradiation. Reproduced with permission from [109]. Copyright 2018, Elsevier.
Figure 11. SEM images of (a,b) pristine TiO2, (c,d) sensitized CuInS2/TiO2. TEM image of (e) pristine TiO2 and (f) sensitized CuInS2/TiO2. (g) energy dispersive X-ray (EDX) element mapping of Ti, O, Cu, In, and S. (h) Scheme of the illustrated charge transfer and separation in the optimum sensitized CuInS2/TiO2 photocatalyst under simulated sunlight irradiation. Reproduced with permission from [109]. Copyright 2018, Elsevier.
Nanomaterials 10 00337 g011
Table 1. Products, number of transferred electrons, and standard potentials vs. normal hydrogen electrode (NHE) in reduction of CO2 at pH = 7.
Table 1. Products, number of transferred electrons, and standard potentials vs. normal hydrogen electrode (NHE) in reduction of CO2 at pH = 7.
ProductReactionnEo (V vs. NHE)
Carbon monoxideCO2 + 2H+ + 2e→CO + H2O2−0.51
Formic acidCO2 + 2H+ + 2e→HCOOH2−0.58
Oxalate2CO2 + 2H+ + 2e→H2C2O42−0.87
MethanolCO2 + 6H+ + 6e→CH3OH + H2O6−0.39
MethaneCO2 + 8H+ + 8e→CH4 + 2H2O8−0.24
Ethanol2CO2 + 12H+ + 12e→C2H5OH + 3H2O12−0.33
Ethane2CO2 + 14H+ + 14e→C2H6 + 4H2O14−0.27
HydrogenH2O + 2e→2OH + H22−0.41
Table 2. Summary of photocatalytic CO2 reduction performance of various photocatalysts.
Table 2. Summary of photocatalytic CO2 reduction performance of various photocatalysts.
PhotocatalystsReaction ConditionsProduction Rate
(μmol g−1 h−1)
Reference
(Year)
Anatase TiO2 by coexposed {001} and {101} facets300 W Xe lamp
TiO2-NaHCO3-HCl
(solid-liquid)
CH4: 1.35[52] (2014)
Cubic anatase TiO2 nanocrystals (100 ± 13 nm)300 W Xe lamp
TiO2-H2O-CO2
(solid-liquid)
CH4: 4.56
CH3OH: 1.48
[53] (2015)
TiO2 nanosheets with exposed {001} facet2 × 18 W Hg lamps
TiO2-H2O-CO2-NaOH
(solid-liquid)
CH4: 0.2
CO: 0.12
CH3OH: 0.19
HCHO: 0.066
[54] (2014)
Flame-annealed TiO2 nanotubes formed in aqueous electrolyteAM 1.5G, 100 mW cm−2
Catalysts-H2O-CO2
(solid-gas)
CH4: 156.5[56] (2019)
Nanotubes/nanowires assembled from TiO2 nanoflakes with {111} facet300-W Xe lamp
TiO2-H2O-CO2-(CH3)2CHOH
(solid-liquid)
CH4: 1708.1
CO: 463.2
[57] (2019)
Highly dispersed titanium oxide on mesoporous SBA-15 (Ti-SBA-15)100-W Hg lamp (>250 nm)
TiO2-H2O-CO2
(solid-gas)
CH4: 106
CH3OH: 27.7
[61] (2005)
Oxygen-deficient blue TiO2 nanocrystals with coexposed {101} and {001} facets100-W Hg lamp
450-W Xe lamp
TiO2-H2O-CO2
(solid-gas)
CO: 55 (UV-VIS)
CO: 27 (Visible light)
[66] (2016)
Cu and V co-doped TiO2 (P25) deposited on polyurethane
Cu@V/TiO2-PU
Visible light (2 × 20 W white bulbs)
Catalysts-H2O-CO2
(solid-gas)
CH4: 933
CO: 588
[72] (2017)
3% NaOH-surface modification TiO2 (ST01)300-W Xe lamp
Catalysts-H2O-CO2
(solid-gas)
CH4: 8.7[73] (2014)
5 ultrathin MgO layers deposited on porous TiO2 mixed anatase-rutile phases by atomic layer deposition450-W Xe lamp
Catalysts-H2O-CO2
(solid-gas)
CO: 54[76] (2019)
TiO2-0.5% Ag100-W Xe lamp (320–780 nm)
Catalysts-H2O-CO2
(solid-gas)
CH4: 2.1[75] (2014)
TiO2-0.5% Au100-W Xe lamp (320–780 nm)
Catalysts-H2O-CO2
(solid-gas)
CH4: 3.1[75] (2014)
TiO2-0.5% Rh100-W Xe lamp (320–780 nm)
Catalysts-H2O-CO2
(solid-gas)
CH4: 3.5[75] (2014)
TiO2-0.5% Pd100-W Xe lamp (320–780 nm)
Catalysts-H2O-CO2
(solid-gas)
CH4: 4.3[75] (2014)
TiO2-0.5% Pt100-W Xe lamp (320–780 nm)
Catalysts-H2O-CO2
(solid-gas)
CH4: 4.3[75] (2014)
1-D nanostructured TiO2 single crystals loaded with Pt nanoparticles400-W Xe lamp
Catalysts-H2O-CO2
(solid-gas)
CH4: 1361
CO: 200
[35] (2012)
Ag-Pd on N-doped TiO2 NSs300-W Xe lamp
Simulated sunlight
Catalysts-H2O-CO2
(solid-liquid)
CH4: 79[88] (2018)
Montmorillonite (MMT) dispersed Au/TiO2 nanocatalystSimulated sunlight
Catalysts-H2O-CO2
(solid-gas)
CO: 1223[95] (2017)
TiO2 powder (Degussa P25) loaded with Au–Cu alloy nanoparticles1000-W Xe lamp
Catalysts-H2O-CO2
(solid-gas)
CH4: 2200[96] (2014)
Au–Cu bimetal as cocatalyst supported on SrTiO3/TiO2300-W Xe lamp
Catalysts-H2O-CO2
(solid-gas)
CO: 3770
CH4: 421.2
C2H6: 190.1
C2H4: 73.3
C3H6: 40.8
[97] (2015)
TiO2-graphene 2D sandwich-like hybrid nanosheets500-W Xe lamp
Catalysts-H2O-CO2
(solid-gas)
C2H6: 16.8
CH4: 8
[102] (2013)
2.5% CuInS2/TiO2350-W Xe lamp
Catalysts-H2O-CO2
(solid-gas)
CH4: 2.5
CH3OH: 0.86
[109] (2018)
TiO2/carbon nitride nanosheet nanocomposites300-W Xe lamp (>325 nm)
Catalysts-H2O-CO2
(solid-gas)
CO: 1.96[106] (2019)

Share and Cite

MDPI and ACS Style

Nguyen, T.P.; Nguyen, D.L.T.; Nguyen, V.-H.; Le, T.-H.; Vo, D.-V.N.; Trinh, Q.T.; Bae, S.-R.; Chae, S.Y.; Kim, S.Y.; Le, Q.V. Recent Advances in TiO2-Based Photocatalysts for Reduction of CO2 to Fuels. Nanomaterials 2020, 10, 337. https://doi.org/10.3390/nano10020337

AMA Style

Nguyen TP, Nguyen DLT, Nguyen V-H, Le T-H, Vo D-VN, Trinh QT, Bae S-R, Chae SY, Kim SY, Le QV. Recent Advances in TiO2-Based Photocatalysts for Reduction of CO2 to Fuels. Nanomaterials. 2020; 10(2):337. https://doi.org/10.3390/nano10020337

Chicago/Turabian Style

Nguyen, Thang Phan, Dang Le Tri Nguyen, Van-Huy Nguyen, Thu-Ha Le, Dai-Viet N. Vo, Quang Thang Trinh, Sa-Rang Bae, Sang Youn Chae, Soo Young Kim, and Quyet Van Le. 2020. "Recent Advances in TiO2-Based Photocatalysts for Reduction of CO2 to Fuels" Nanomaterials 10, no. 2: 337. https://doi.org/10.3390/nano10020337

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