Next Article in Journal
Therapeutic Effect of Erythropoietin on Alzheimer’s Disease by Activating the Serotonin Pathway
Next Article in Special Issue
Asymmetric Non-Fullerene Small Molecule Acceptor with Unidirectional Non-Fused π-Bridge and Extended Terminal Group for High-Efficiency Organic Solar Cells
Previous Article in Journal
Four Novel PAX9 Variants and the PAX9-Related Non-Syndromic Tooth Agenesis Patterns
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic Carbon Dioxide Conversion by Structurally and Materially Modified Titanium Dioxide Nanostructures

1
Department of Photonics, National Sun Yat-sen University, No. 70, Lien-Hai Rd, Kaohsiung 80424, Taiwan
2
Department of Physics, SRM Institute of Science and Technology, Ramapuram Campus, Chennai 600089, Tamil Nadu, India
3
Semiconducting Oxide Materials, Nanostructures and Tailored Heterojunction (SOMNaTH) Lab, Functional Oxides Research Group (FORG) and 2D Materials and Innovation Centre, Department of Physics, IIT Madras, Chennai 600036, Tamil Nadu, India
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(15), 8143; https://doi.org/10.3390/ijms23158143
Submission received: 24 June 2022 / Revised: 21 July 2022 / Accepted: 22 July 2022 / Published: 24 July 2022

Abstract

:
TiO2 has aroused considerable attentions as a promising photocatalytic material for decades due to its superior material properties in several fields such as energy and environment. However, the main dilemmas are its wide bandgap (3–3.2 eV), that restricts the light absorption in limited light wavelength region, and the comparatively high charge carrier recombination rate of TiO2, is a hurdle for efficient photocatalytic CO2 conversion. To tackle these problems, lots of researches have been implemented relating to structural and material modification to improve their material, optical, and electrical properties for more efficient photocatalytic CO2 conversion. Recent studies illustrate that crystal facet engineering could broaden the performance of the photocatalysts. As same as for nanostructures which have advantages such as improved light absorption, high surface area, directional charge transport, and efficient charge separation. Moreover, strategies such as doping, junction formation, and hydrogenation have resulted in a promoted photocatalytic performance. Such strategies can markedly change the electronic structure that lies behind the enhancement of the solar spectrum harnessing. In this review, we summarize the works that have been carried out for the enhancement of photocatalytic CO2 conversion by material and structural modification of TiO2 and TiO2-based photocatalytic system. Moreover, we discuss several strategies for synthesis and design of TiO2 photocatalysts for efficient CO2 conversion by nanostructure, structure design of photocatalysts, and material modification.

Graphical Abstract

1. Introduction

Plants preserve carbon cycle in nature via photosynthesis process as they convert CO2 into carbohydrates, but several activities such as industrialization, urbanization, deforestation and excessive consumption of fossil fuels affected this cycle considerably [1]. The accumulation of CO2 in the atmosphere, beyond the capacity of the nature to handle, has created severe global warming phenomena that result in climate change. Now, the devastating effects from unusual weather patterns are visible all around us. In the recent years, therefore, CO2 capturing and utilization is a focal subject for the scientific community [2,3], as it addresses the energy crisis and global warming without hindering the development plans or urbanization [4]. In particular, recycling of CO2 into carbon containing and value-added chemicals regenerates fuel within the present hydrocarbons-based energy infrastructure.
CO2 is a linear molecule, with weak electron affinity, chemically stable and the nucleophilic approach at the carbon atom governs its conversion reaction [5]. C=O bond dissociation requires more than 750 kJ/mol of energy [6]. Thermodynamically, this reaction is an uphill one, hence input of energy is required to break C=O bond. In nature, the photosensitizers (e.g., chlorophyll) capture energy from sunlight to use it in the endothermic reactions.
To mimic these natural solar energy conversion, several techniques such as photoelectrochemical (PEC) [7], biochemical [8], photocatalysis [9], radiolysis [10], thermo-catalysis [11], and electrocatalysis [12] have been proposed for CO2 conversion into useful chemicals. Photocatalytic (PC) methods has been identified as one of the most suitable approaches to convert CO2 into different gaseous (ethane, methane, etc.) and liquid (ethanol, methanol, formate etc.) products under solar light irradiation at ambient temperature and pressure [13]. The most attractive aspect of such approach is the utilization of renewable solar radiation to draw the energy required for driving the catalytic process. In the case of a pure photocatalytic approach (without any electric field), this is called artificial photo-synthesis that imitates the energy cycle of nature [14]. In addition, use of renewable solar energy brings in associated advantages such as environmental compatibility, economic feasibility, and product selectivity [15].
Up to now, several different types of materials such as metal complexes, organic molecules, precious metals, ions, organic molecules, and semiconductors have been explored to enhance the efficiency of photocatalytic processes. Among these materials, semiconductors in general and metal oxides in particular have shown immense potential over the others with regard to stability, toxicity, and feasibility of the fabrication etc., which makes them desirable in photocatalytic applications [16].
The general understanding for an efficient photocatalytic process by semiconductor materials, the electrons and holes in the semiconductor materials should have sufficient energy to ionize the CO2 and H2O, respectively. In other words, energies of the photogenerated electron and hole (e/h+) must be higher than the overpotential of H2O/O2 (0.82 vs. NHE at pH 7) and lower than the overpotential of CO2/HCOOH (−0.61 V vs. NHE at pH 7), respectively [17,18]. Performing photo-assisted redox reactions such as CO2 reduction or water oxidation therefore requires a photocatalyst with an optimum bandgap (at least 2.88 eV) [19] so as to generate electrons and holes with sufficient energies. Simultaneously, the photocatalyst must be able to absorb large fraction of the solar spectrum even from visible and near IR light absorption (in addition to UV) [20].
Historically, TiO2 is the most extensively investigated semiconductor material for photocatalytic CO2 conversion since the successful demonstration of the photoelectrochemical CO2 conversion by Inoue et al. in 1979 [21]. Moreover, owing to superior advantages such as material robustness, chemical and thermal stability, nontoxicity, corrosion resistance, and ease of synthesis, TiO2-based photocatalytic systems became popular and related applications have been focal target for researchers due to its tremendous potential in different solar driven processes such as water splitting, pollutant degradation and CO2 conversion [21,22]. It is well known TiO2 has three crystalline phases, anatase, rutile, and brookite possessing the bandgap values in the range of 3.0–3.6 eV. However, the bandgap is relatively large to absorb whole solar spectrum and this limits the light absorption of TiO2 only to UV region, which is ~5% of total solar energy available. Moreover, the charge recombination in TiO2 is a serious issue and this makes the photocatalytic reaction by TiO2 inefficient. Hence, increasing the optical absorption and decreasing the charge recombination of the TiO2 are of another prime importance. To overcome these limitations, lots of research has proposed on material modification, low dimensional introduction, and design of photocatalytic structure for TiO2 to achieve highly efficient photocatalytic CO2 conversion.
Therefore, in this review, we summarize the works that have been carried out for the enhancement of photocatalytic CO2 conversion by material and structural modification of TiO2 and TiO2-based photocatalytic system. We provide several strategies and their details: (1) crystal facet engineering (2) nanostructures (3) junction formation (4) material modification, in particular, by hydrogenation (5) single atom photocatalysts (6) metal-organic framework.

2. Strategy I: Crystal Facet Engineering of TiO2 Photocatalyst

Titanium dioxide is a multifunctional material that has attracted researchers due to its applicability to various fields such as photocatalysis, photovoltaics, and biomedical applications [23,24]. TiO2 nanoparticles and nanocrystals have been obtained via different preparation methods such as flame spray pyrolysis, hydrothermal treatment and sol-gel method. In particular, crystallites with selective facets are helpful in catalysis. In addition to wet chemistry approaches, powders produced from flame pyrolysis resulted in nanocrystals with thermodynamically stable (101) surface [25,26] and (100)/(010) surface planes, and some reactive (001) facets (Figure 1 shows the different types) [27,28].
While, the size of TiO2 nanoparticles differs remarkably depending on the synthesis process [29], the preparation techniques affect the crystal shapes significantly [30]. Choosing a specific synthesis method depends on the required morphology of TiO2 material, which further depends on the target application. The methods of synthesis significantly affect the morphology, crystallinity, and phase, which have their influence on the physical and chemical properties. The synthesis strategies are chosen, considering the ability to tailor particle-size, the integration with other structures or phases [31], flexibility of self-assembly [32], possibility of doping with other elements [33,34], also inlaying heteroatoms, heterostructures or quantum dots [35,36] in order to manipulate the electronic and optical properties.
Anatase is the most investigated photoactive polymorph of TiO2, with its thermodynamically stable (101) facets (>94%), dominating in consonance with the Wulff construction. The morphology, crystal growth, and facets can be controlled by using a shape and growth controller such as HF [28]. In addition, in another approach, a mixture of HF/alcohol has been employed for establishing metastable surfaces reaching 98.7% of (001) and 1.3% of (100) facets [37]. Using F along with citric acid or hydroxyl acids results in continuous curvature of rutile and anatase, where F ions play not only the role of a stabilizer for (001) facets growth but also as an etching reagent [38,39]. The morphology of the anatase nanoparticles can be manipulated via the relative concentrations of OH and F [40]. Hydroxyl groups boost the isotropic growth, whereas F eliminates (001) surfaces supporting the TiO2 crystals lateral growth. However, an excessive concentration of F reduces the size of particles significantly as a result of TiOF2 formation.
The facet engineering effect was reported by Wang et al. [41]; where they found that the Schottky barrier height of Au/TiO2 (101) interface is lower than its counterparts from the interface (001). Which enhanced the electrons transfer from CB of TiO2 to Au and significantly enhances the photocatalytic performance in producing CO and CH4 in comparison with other samples containing Au/TiO2 (001) interfaces. In another study, Dong et al. [42] showed that crystal facets engineering of TiO2 loaded with Cr2O3, (2HF-TiO2/0.2Cr2O3) exhibited 30-fold increment in CO2 conversion compared to the TiO2 without facet engineering.
Studies and theories illustrated that (101) facets are the most stable facets thermodynamically for anatase TiO2, whilst other facets (100) and (001) are active and possess high surface energy [43]. Anatase crystal has an equilibrium shape consisted of slightly truncated tetragonal bipyramid enclosed with two (001) and eight (101) facets [44]. Tetragonal nanorods of anatase crystals consisted of high ratio of lateral (100) facets have been fabricated via immersing alkali titanate nanotubes in basic solution followed with hydrothermal transformation [45]. In addition, Anatase crystals with elongated truncated tetragonal bipyramids that contains high ratio of (100) facets; enclosed via different types of facets (001), (100) and (101) were fabricated via hydrothermal reaction in aqueous HF solution [46]. However, still a dilemma to fabricate tetragonal cuboid of anatase crystals enclosed only with (001) and (100) facets. Keeping in mind that anatase nanosheets were fabricated by using solvothermal route in 1-butanol solvent containing HF; these nanosheets consists of (100) and (001) facets, 1.3% and 98.7, respectively [47,48]. Preparing anatase cuboids and manipulating facets percentage (100) and (001) may help in evaluating facet reactivity throughout the photocatalytic reactions. Hence, it is for crucial significance to synthesize anatase cuboids enclosed by (001) and (100) facets over a wide size range, as in lithium-ion batteries and solar cells where several studies [49,50,51,52] relied on the exposed facets of anatase effect on its optical and electrochemical features to be applied in such applications. Photocatalytic materials can be formed via the integration of multiple components, adjusting the facet of each component. This method enables to boost the photocatalytic performance but, since the facet engineering in the multiple component system, in general, are complicated for its delicate control, the combination of two pre-synthesized components is more preferable for its usage, rather than a new component on the one pre-synthesized material [53]. For example, tailoring the surface facets of TiO2 seeds into nanocrystals with graphene oxide can be seen in Figure 2a. capping agent deficiency results in octahedral TiO2 nanocrystals enclosed by TiO2-101-G. TiO2-001-G nanosheets and TiO2-100-G nanorods were formed on graphene with F and S O 4 2 Semployed as capping agents, as shown in Figure 2b–g).

3. Strategy II: Nanostructured TiO2

Although TiO2 in its bulk form has been investigated for decades, its nanostructured morphologies also have been investigated for photocatalysis and other applications [55], focusing on one dimensional (1D) [56], two dimensional (2D) [57], or three dimensional (3D) structures [58]. Nanostructures or nanomaterials are defined as features where at least one dimension is smaller than 100 nm [59]. However, in a more technically elaborated way, a nanomaterial is defined as where charge carriers are quantum-mechanically confined as evidenced by the consequent modifications in electronic and optical properties. However, even before the electronic or optical properties are investigated, the nanomaterials are characterized by a huge enhancement in the surface to volume which enhances the interaction with the surrounding environment [60]. Each of these nanostructures have their own characteristics in terms of light scattering, aspect ratio, recyclability, major surface area, stability, and directional transfer of photogenerated charges that decreases recombination rate. Synthesis method for 1D nanostructures involves tailored and directional growth, which is achieved either by the structure directing agents or by the use of templates. The directional growth is not required for 2D and 3D morphologies, but the synthesis process may involve multiple steps and highly optimized conditions to obtain structures with minimal defects. In the case of 0D nanostructures such as quantum dots, although there have been studies reported in the literature, the enhancement in band gap caused by the quantum confinement effect reduces the effectiveness of use in photocatalysis. Following chapters provide fundamental features of each nanostructure and its application for CO2 conversion.

3.1. Zero-Dimensional Nanostructured TiO2

The nanoparticles or 0D nanostructures offer a lot of advantages and a high degree of flexibility in terms of use in photocatalysis. In particular, nanoparticles can be used to functionalize various surfaces of materials that help in forming either heterojunctions or modifying light absorption. Further, tailoring the size of the nanoparticles changes the surface area, which, in turn, influence the catalytic reactions. Moreover, when formed into quantum dots, the optical and electronic properties are modified drastically to help in catalytic reactions. These nanostructures are cheap, stable, recyclable, and biocompatible [61]. TiO2 nanoparticles with their special morphology and acid-base sites have been useful in many catalytic reactions under mild conditions compared to other metal oxides (e.g., CuO, ZnO, etc.) [62]. Liu et al. [63] constructed 0D nanoparticles/2D CoP nanosheets heterojunction and the results showed improvement in photocatalytic H2 evolution rate in comparison to pure TiO2. Another study [64] reported that In-doped TiO2 nanoparticles improved the photocatalytic activity for CO2 reduction where CO was detected and CH4 yield increased remarkably. In addition, Wada et al. [65] employed rutile TiO2 nanoparticles as a modifier to enhance the charge transfer, where RuRe/TiO2/NS-C3N4 showed capability in converting CO2 into CO with high selectivity under visible light (λ > 400). Tseng et al. [66] used sol-gel method in synthesizing Cu/TiO2 nanoparticles and measured the photoreduction that showed methanol yield much higher than those resulted from sol-gel TiO2 and Degussa P25, as well. Moreover, Pt/TiO2 nanoparticles composites yield CH4, H2 and C2H6 under visible light irradiation with increment of 3.7 times in comparison with Pt/P25 [67]. Perovskite quantum dots also showed its potential for photocatalytic CO2 conversion with TiO2 [68,69].

3.2. One-Dimensional Nanostructured TiO2

The one-dimensional morphologies such as nanowires, nanorods, nanobelts and nanotubes have several interesting properties including directional charge transport, improved light absorption by high aspect ratio and widen surface area [70]. A lot of research reports further modified properties of nanostructures through doping or decoration of other materials such as graphene derivatives or noble metals to enhance photocatalytic CO2 conversion with high aspect ratio [71,72]. The integration of the specific geometry and the high aspect ratio yield dramatical enhancement in charge carrier generation, separation, and transport which boost the conversion efficiency [73]. For instance, TiO2 nanorods (TNRs) have a single-crystalline structure and small boundary resistance [74], that reduces the impact from grain boundaries and supplies fast electron transportation [75,76]. It was reported that TNRs as shown in Figure 3 shows higher photocatalytic activity than the nanoparticles because of the increment in active sites and the influence of the crystal plane [77]. Moreover, the comparison between TiO2 nanorods and the counterparts of nanoparticles showed that the recombination rate of the nanorods are lower that enhanced the photocatalyst photocatalytic activity [78].
Wang et al. [80] fabricated a one-dimensional TiO2 single crystal with ultrafine Pt nanoparticles (0.5–2.0 nm) by versatile gas-phase deposition. This film showed extremely high efficiency in CO2 photoreduction with selective formation of methane compared to pristine P25. In addition, TiO2 nanotubes fabricated by Ping et al. [81] reduced CO2 with H2 into methanol and ethanol with photocatalytic performance higher than that of TiO2 nanoparticles. Another study used microwave solvothermal approach in deposition of Pt nanoparticles on TiO2 nanotubes (TNT). This composite promoted the photocatalytic conversion of CO2 with water into methane [82]. TiO2 1D nanostructured by alkaline hydrothermal method also exhibited its promising CO2 conversion performance via the heterostructure formation with some materials such as Nb2O5, CNM and Bi2S3 under visible light irradiation [83]. The results showed improved efficiency attributed to the enhanced light absorption and the charge separation [83]. The band edges of the aforementioned materials embedded well over the band edge of TiO2; hence, applying light irradiation excites electrons, which, in turn, flow to the conduction band of the TiO2 and react with the absorbed CO2 species. Although deposition of these materials reduces the surface area and CO2 adsorption of the TNT, enhancing charge transfer kinetics brought advantage over the reduction of surface area. In addition, the 1D structures showed excellent performances, combining with various techniques and materials: for examples, TiO2 nanoflower films modified with Cu, depositing CdS, grafting CoOx nanoparticles on TNTs with defects via hydrogenation through the heterostructure by N2/H2 annealing [84], TiO2 nanowires (TNWs) loaded with noble metal nanoparticles such as Au [85] or Ag [86], TNT with electrodeposited Ag nanoparticles [87], TNW with Pd nanoparticles [88], TiO2 nanobelts (TNB) forming heterostructures with ZnFe2O4 nanoparticles [89], and TNT covered with rGO sheets with embedded TiO2 nanoparticles [90]. All the techniques and materials implemented onto 1D TiO2 nanostructures enhanced the photocatalytic performance by improving the product (CO, CH4, CH3OH, CH, CF) yield under light irradiation because of all or some of these factors: (i) increasing active surface area, (ii) increasing CO2 adsorption on the surface (iii) trapping electrons via oxygen defects (iv) enhancing charge separation.

3.3. Two-Dimensional Nanostructured TiO2

Ultrathin two-dimensional nanomaterials possess sheet-like structure with a thickness of few atoms (less than 5 nm), their widths are larger than several hundred nanometers [91,92]. Their superb physical and chemical properties have led significant attention for diverse lateral structured applications [93]. In comparison with 0D and 1D, 2D nanomaterials have extraordinary advantages; support them with optimistic potential for photocatalytic applications.
  • First and foremost, 2D materials possess larger surface-to-volume ratio over their bulk counterparts [94]. Hence, 2D materials have more active sites on their surface that can enhance their photocatalytic performance significantly.
  • Second, their atomic thickness benefits mass transport and light energy harvest [93]. The ultrathin structure minimizes the distance of the charge migration from the bulk to the surface, decreasing carrier recombination and enhancing the photocatalytic activity [95].
  • Third, the high fraction of coordinated unsaturated centers can work as active centers and interact with the substrate intimately [96].
Hence, they perform stellar platforms to prepare multicomponent photocatalysts. The aforementioned extraordinary properties provide diversified number of opportunities with high activity and selectivity for CO2 reduction. In particular, synthesizing fine-tuned and strong photocatalysts that fulfil the requirements of CO2 reduction applications [97]. These properties encouraged Tu et al. to fabricate 2D sandwich-like hybrid nanosheet out of graphene and TiO2 in Figure 4, where TiO2 nanoparticles uniformly were loaded onto graphene nanosheet to prevent their breakdown and restacking.
2D structures provide large active surface area, enhanced surface adsorption, enhanced interfacial charge transfer, and selectivity caused by the band edge alignment. Zhou et al. loaded with TiO2 nanoparticles onto layered nanosheets of g-C3N for CO2 conversion [99]. Urea served as a source of N doping as well as g-C3N4 precursor. Low urea formed an N-doped TiO2 and resulted in CH4 during the photocatalytic CO2 reduction while high urea formed composite of g-C3N4 and N-TiO2 that produced CO. The reason behind the selectivity of products is alignment of band edges with respect to the redox potentials of the possible products. The photocatalytic yield increased due to average surface area, enhanced light absorption, promoted charge transfer and well-aligned band edges with respect of product redox potentials. Likewise, TiO2-g-C3N4 nanosheets heterostructure (TNS-CNN) has been synthesized via in situ pyrolysis approach [100]. TNS-CNN used H2O and H2 as reducing agents in CO2 conversion process where its CO yield was very high compared to pristine TNS. This performance is attributed to the surface area increment, enhanced charge transfer kinetics, role played by H2 and light absorption. Ultrathin TiO2 nanosheets also play an essential role for efficient photocatalytic CO2 conversion, when prepared from the lamella structure of TiO2-Octylamine [101]. The conversion efficiency is a result of several factors such as increasing CO2 adsorption sites, enormous increasing in surface area, which in turn, increased the light absorption. Moreover, the fluorescence lifespan of the generated charges into the ultrathin TiO2 nanosheets is higher when compared to their counterparts in bulk material. Thus, these ultrathin nanosheets provide efficient charge separation within its 2D channels. In another report, TiO2 ultrathin nanosheets (TiO2-U) were synthesized by hydrothermal method followed by photochemical deposition of Pt nanoparticles [102]. Moreover, an interesting study suggested growing the photocatalytic material onto a 2D conductive substrate. Recently, Ti3C2 MXenes (TT) has been synthesized and covered with TiO2 nanoparticles [103]. Upon calcinating TT at 550 °C, the TiO2 nanoparticles formed at the edges and the surfaces of TT layers which improves the surface area by making it rougher. Applying higher temperature than 550 °C decreased the photocatalytic performance due to the decreased proportion of the conductive TT. Hence, TT offers an efficient charge separation that improves the performance and the surface area contributed significantly in providing more reactive sites for CO2 adsorption for conversion process, as well. Another report described the 2D nanostructure of Bi2WO6-TiO2 bi-nanosheet (BT) for CO2 conversion into CH4 and CO [104]. This report provided an approach concerning carbonaceous intermediates or surface species in the value-added chemicals generation. It was found that BT resulted in improved CO and CH4 yield compared to pristine material due to the enhanced charge transfer and the Z-scheme mechanism.
As known, TiO2 has three common polymorphs anatase, brookite, and rutile. Brookite is the least used one while rutile is the most common one. It is well known that TiO2 has a wide bandgap of 3.0 eV for rutile. In addition to this, it was reported that the bandgap of anatase is about 3.2 eV, while that of the brookite is anywhere between 3.0–3.6 eV. Anatase nanoparticles with grain size (5–10 nm) shows a blue shift in the absorption edge of 10 nm, and their bandgap is about 3.3–3.4 eV in comparison with the commercial sample with a crystal size 39 nm which bandgap is about 3.2 eV [105]. Another study reported that the bandgap of pure TiO2 nanoparticles is 3.7–3.9 eV, as shown in Figure 5a [106].
Studies proved that as the material shrinks to the nanometer scale, the bandgap starts to increase due to the quantum size effect [107]. Hence, TiO2 nanowires (TNWs) have higher bandgap compared to the bulk materials, for instance growing the TNWs along with (001) direction results in a blue shift and this shift depends on the size of the TNRs [107]. In addition, the bandgap of titanate nanotubes is almost 3.84 eV ascribed to quantum confinement effect, shown in Figure 5b [60]. Moreover, in a comparison between TiO2 nanosheets (TNSs) and bulk materials, the spectrum of the TNSs showed significant blue-shift that indicates increasing in the bandgap [108].

4. Strategy III: Formation of the Junction with TiO2

Photocatalysis CO2 conversion is closely related to the management of charge carriers, namely, electrons and holes, that are necessary for CO2 reduction and water oxidation, mainly governed by photon absorption, photocarrier generation, and charge separation and transfer. These reactions rely on several factors: bandgap, Eg of photocatalyst, photo carrier generation rate, charge transfer kinetics, and recombination rate. Photocarrier generation rate depends on the conditions of the irradiation and the optical absorption properties of photocatalysts while the charge transfer depends on the reaction occurrence whether it takes place at the surface or within photocatalysts. The recombination can be minimized by improving the crystallinity, increasing transfer kinetics, or the control of the crystallite size [109]. Highly crystallized material minimizes the presence of impurities, surface or bulk defects. Crystallite size control stimulates the point defects of particles that trap e or h+ which in turn delays the recombination for a few micro or nanoseconds [95]. In addition, the reaction efficiency can be enhanced by increasing the adsorption capability of the surface [110].
To achieve these improvement and enhancement in properties and performance of TiO2 photocatalysts, we can make an effective junction with other materials. As mentioned, since TiO2 has disadvantages for efficient light absorption and charge transfer, making a junction with efficient light absorbers or/and charge transporting materials would be beneficial for the enhancement of the performance. The subchapters below provide the features and examples of several junctions that can be made with TiO2.

4.1. A Semiconductor-Semiconductor Heterojunction

The most typical semiconductor-semiconductor heterojunction can be formed in p-n junction. The p-n junction, where an n-type and a p-type semiconductor are in intimate contact (for charge transport across the interface), has several advantages in photocatalysis. When two such semiconductors make a junction, a depletion layer is formed at the interface and internal electric field is established at the interface due to the relative positions of the Fermi levels. This built-in electric field at this junction helps in separating the photogenerated electron-hole pairs thus minimizing the recombination. Further, the junction can be designed in such a way that the light absorption range of one semiconductor can be extended by choosing the other semiconductor with a smaller band gap. Therefore, two fundamental challenges, light absorption and charge carrier separation, in TiO2-based photocatalysts n photocatalysis- can be tackled by creating p-n junction with materials that have suitable band gaps and intimate energy band structure [111].
Moreover, non-p-n junction, that is, n-n junction between the materials also helps in photocatalysis in a similar manner. In such a heterostructure, the band edge offsets at conduction and valance bands endow a driving force for charge transfer and separation and, with a proper selection of materials, enhanced light absorption can be expected. Such a configuration, called a type II heterojunction is highly useful in photocatalysis [112].
These semiconductor-semiconductor heterojunction structure has been demonstrated in lots of work in the field of photocatalysts. Guo et al. [113] reported that Bi2WO6/TiO2 heterojunction photocatalysts showed strong adsorption ability and improved visible light photocatalytic activities. Shang et al. [114] also studied Bi2WO6/TiO2 photocatalytic activity as it demonstrated enhanced photocatalytic activity by 8 orders compared to the bare Bi2WO6. MgO covered TNT via alkaline hydrothermal reaction also demonstrated for CO2 conversion [115]. This heterojunction enhanced the CO2 photoreduction as indicated by higher amounts of CO and CH4 compared to that from sole TNT photocatalysts. This enhancement is attributed to the chemisorption of CO2 and subsequent conversion into MgCO3 which is more reactive than the linear molecules of CO2. Li et al. [83] fabricated TNT via hydrothermal reaction followed by constructing CdS/TiO2 and Bi2S3/TiO2, the photocatalytic performance of both of them enhanced CO2 reduction into CH3OH under visible light irradiation. In addition, Li et al. [115] presented a study showing the development of MgO/TNTs films in CO2 photoreduction activity into methane in comparison with TiO2 films where MgO played a crucial role in CO2 methanization.

4.2. Semiconductor-Metal Heterojunction

TiO2 photocatalysts forms a junction with various metal materials to enhance the photocatalytic performance. A metal in contact with a semiconductor creates Schottky or ohmic contact and these contacts influence on electric field or charge concentration distribution at interface. As in the semiconductor-semiconductor junction above, the metal materials can affect the electrical properties of the photocatalytic structures for charge separation and transfer of the semiconductor materials forming a junction together. In addition, some special metals such as Ag, Au, or Cu help in extending light absorption with suitable plasmonic effect when they make a junction with TiO2 [116,117,118,119]. Further, metals with suitable surface energies also help in enhancing the adsorption of the gaseous/liquid species under catalytic reaction [18]. Wang et al. [120] successfully fabricated Au/TiO2 heterojunction that resulted in reduction products CH4 and CO, with 80% CH4 selectivity. Saraev et al. [121] modified TiO2 with Pt and Cu/CuOx and reported that this design is an efficient photocatalyst for CO2 conversion as it shifted the working range to the visible light and produced CH4. Another study reported that Au0.25Pt0.75/TiO2 nanofiber showed higher activity of CO2 photoreduction into CH4 under UV-vis irradiation [122]. In addition, Mankidy et al. [123] reported that Ag-Pt bimetallic and core sell Ag@SiO2 onto TiO2 showed a significant development in the photoreduction of CO2 with H2O into CH4.

4.3. Semiconductor-Carbon Heterojunction

Similarly, to the metal-semiconductor junctions, carbon materials are also used to form heterostructures with photoactive semiconductors. Several carbon materials such as activated carbon, graphene, and graphene oxide have been employed in heterojunctions preparation. These materials demonstrate high surface area that improves molecules adsorption and the photocatalysis performance of the material [124]. The porous morphology demonstrated by carbon materials adsorbs gaseous species for catalytic reactions. Further, graphene and carbon nanotube (CNT) have a metallic structure and forms a Schottky barrier with a semiconductor material at the interface that improve charge transfer and alleviate the recombination rate by the established built-in electric field. Simultaneously, the metallic nature of graphene or CNTs efficiently collect photogenerated charges from the semiconductor [125,126]. Photogenerated electrons move by the built-in electrical field from conduction band of semiconductor to CNTs in order to balance Fermi levels, while holes exist in the semiconductor share in redox reaction [127]. Padmanabhan et al. [128] prepared a study showing that TiO2/graphene is more active for photocatalysis than sole TiO2. In the report, the graphene sheet acted similar to an electron acceptor facilitating the transfer and separation of the generated electrons during irradiation, thus reducing the e/h recombination. Carbon quantum dots also led the improvement for photocatalytic reaction with TiO2. A CQD/TNT nanocomposite yields more than two times higher production rates both for CO and CH4 compared to those of bare TNTs [129]. Another study presented by Morawski et al. [130] stated that combining commercial P25 with carbon spheres then depositing this composite on glass fiber fabric showed high efficiency and selectivity in CO2 reduction into CO.

5. Strategy IV: Modified TiO2 Nanostructures by Hydrogenation

The next strategy applicable for efficient TiO2 photocatalysts is to modify its properties. Several approaches have been explored to modify properties of TiO2 through conventional doping process. Doping of metal ions (Cr, Sn, Zn, W, etc.) and nonmetallic ions (C, P, I, N etc.) into TiO2 to create discrete or midgap energy states within its bandgap has been attempted to enhance its electro- and photo-response [131,132]. However, the doping elements are also known to create recombination center simultaneously [131]. While, nonmetallic ion doping is difficult due to the diverse chemical properties of the dopant ions and the existence of O2− in TiO2, Nitrogen is one of the elements that has been extensively doped into TiO2 for visible light absorption [133,134].
Another way of modifying the properties of TiO2 is hydrogenation that is different from the conventional doping method. The hydrogenation is the way of introducing oxygen vacancy by hydrogen element that results in the incorporation of Ti3+ to TiO2 [135,136]. Namely, Hydrogenation reduces TiO2 through the conversion of Ti4+ to Ti3+ or other states. Depending on the degree or method of hydrogenation, the colors of the hydrogenated TiO2 can be varied: black, blue, or brown [137]. Hydrogenation generally results in surface modifications in few nm and leads to the modification of energy band structure of TiO2, forming additional energy states located under the conduction band edge [137]. These modified properties of TiO2 by the hydrogenation can contribute to many advantageous characteristics for more efficient photocatalytic performance: enhanced light absorptions and control of bandgap. To implement the hydrogenation to TiO2, a number of techniques such as electrochemical reduction, metal reduction, NaBH4 reduction, laser ablation, microwave radiation, ultrasonication, ion thermal process, and oxidation have been carried out [138,139]. The characteristics of the modified TiO2 are affected by various factors, including not only experiment conditions such as reactants, temperature, concentration and pressure of hydrogen and reaction time but also material conditions such as the surface morphology, defect content, shape, and size [140,141,142]. Here this section provides the examples of hydrogenated TiO2 synthesis and characteristics, and investigates its applicability to TiO2 nanostructures.
The several approaches are valid for the formation of the hydrogenated TiO2 to modify nanotubes. The hydrogenated TiO2 is basically a reduced form of TiO2 and can have extended light absorption region due to the creation of midgap states. In one of the reports, black TiO2 has been synthesized via solvothermal method using ethylenediamine followed with calcination of the nanotubes (NT) at 600 °C in a hydrogen atmosphere [143]. The photoelectron spectra of the surface showed Ti3+ and oxygen deficiencies (it is also named as self-doping defects) which contributed in bandgap reduction [143]. In another report, 2D TiO2 nanosheets have been synthesized via evaporation-induced self-assembly, followed by solvothermal treatment and ethylenediamine reflux [144]. Further, TiO2 nanospheres has been obtained via combining hydrogenation with surfactant-induced solvothermal method [145] to achieve reduced TiO2 in a controlled manner. These helped the morphology of TiO2 to mitigate aggregation and to have low surface energy [145]. Furthermore, 1D TiO2 nanotubes have been fabricated via applying hydrogenation and facile solvothermal method [146]. All these reports indicate that controlled hydrogenation leads to the formation of the reduced TiO2 that helps in photocatalytic processes [146].
To obtain decent hydrogenation, the experimental conditions in the synthesis process play a significant role in controlling properties of the materials. It has been found that the reactor materials impact the properties: stainless-steel reactor resulted in black powder, while in quartz reactor blue powder was obtained indicating different extents of reduction [147,148]. The initial powder was prepared via mixing 2 g of TiO2 in 50 mL NaOH at 120 °C for 48 h, then washed in water and HCl followed by drying overnight at 110 °C to yield titanate nanotubes [142]. Moreover, hydrogen in the atmosphere can easily reduce TiO2, and processing in hydrogen ambient resulted in gray TiO2. Depending on the degree of hydrogenation and processing conditions, various colors appear on TiO2. In another report, the color of protonated titanates converted into brown through calcination at 500 °C in H2 atmosphere (N2:5%) for 4 h [142]. Hydrogenation of anatase nanowire microspheres demonstrated high visible light absorption and contained Ti-H and O-H bonds which in turn leads to stabilization of surface disordered layer [142]. In addition, the pressure can be also an important factor for the hydrogenation. The hydrogenation of TiO2 was implemented using H2 pressure at lower temperature with small amount of Pt. In this hydrogenation, hydrogen flow was directed from platinum to TiO2, which is known as an advanced reduction procedure [149].
Hydride processes are also applied for the hydrogenation of TiO2. When hydrogenation process leads to reduced TiO2, the annealing and thermal conditions may pose safety hazards. To avoid these problems, hydride processing has been proposed, in which, hydrides are used in modest conditions to release molecular hydrogen that works as a safe reductant to obtain black TiO2 [150]. Both dry and sol-gel methods can be used in hydride reduction. In the dry method, pristine TiO2 and hydride are mixed and annealed in Ar and N2 atmosphere [151]. In the sol-gel process, NaBH4 is used as the reduction agent. NaBH4 is added to a mixture of two solutions that is made of (EtOH/HNO3) and (EtOH/titanium tetra butoxide), until the gel is reformed. The gel is calcined in a muffle furnace for 3 h to obtain hydrogenated TiO2 nanoparticles [152]. Although a completion of hydrogenation may lead to black TiO2, other colors is also obtainable by adjusting the temperature and the reduction duration [153]. The synthesized black TiO2 nanoparticles were able to absorb > 80% of the sunlight [145]. Nevertheless, the reaction between CaH2 and TiO2 may result in TiO3 and Magneli phase if process is allowed to continue for 240 h. Table 1 summarize hydrogenated TiO2 materials by various techniques, their features fabrication, and performance measure for CO2 conversion.

6. Strategy V: Single Atom Photocatalysts

In the past few years, single atom catalysts (SACs) have nominated as potent photocatalysts that can be employed in CO2 reduction efficiently, owing to their compelling properties. First, SACs have high activity and selectivity caused by their distinctive electron structure and unsaturated coordination sites. Second, they have a notable reduction in metal usage brought by the maximum atom utilization. Third, they possess clear reaction mechanisms endowed by the well-defined active sites. Fourth, they help in understanding and realizing the structure and activity relationship due to their atomic scale structure [172,173]. Theoretically, the valence of a single-atom on a support surface supposed to be zero; but practically the value is different. These atoms are being stabilized depending on the covalent coordination or ionic interaction with the supporting surface atoms, hence it possesses partial charge provided via the metal support interactions. In photocatalytic systems, the photogenerated electrons reduces the stabilized ions into metal ions during the photocatalytic reaction [33]. However, Zhang et al. [174] reported a mass production method of a single atom cobalt to be used in photocatalytic CO2 reduction. As well as, Xiong et al. [175] presented a study of CO2 reduction over Ni single atoms supported on defect-rich ZrO2. This strategy requires further studies to define its influencing factors. It is true that throughout the past decades, physicochemical characteristics and functions of these photocatalysts have been attained. Still, more studies required to tailor the electronic and chemical structures in order to widen its use in photocatalytic applications.

7. Strategy VI: Metal Organic Framework

Metal organic frameworks (MOFs) are microporous or mesoporous crystalline solid where the lattice is being formed via linking metallic nodes with rigid organic linkers possessing two or more coordination positions; the metallic nodes comprise of metal cations or clusters of few ions of metals. In addition, they can be called porous coordination polymers (PCPs) attributed to the nature of the interaction between the metallic nodes and the organic linkers. MOFs includes almost all the di-, tri- and tetra-positive ions mentioned in the periodic table. The materials that can be used in structure and binding groups of the organic linkers are abundant, as well. Yet, the most renowned ones are organo-phosphorous compounds, aromatic polycarboxylates and nitrogenated heterocycles [176].
MOFs has shown promising potentials in photocatalysis applications and energy conversion [177]. Titanium-based MOFs (Ti-MOFs) are attractive for practical applications, especially tetravalent cation due to its good redox activity, rigid framework and strong metal-ligand bonding [178]. Ti-MOFs represent an exemplary role in MOF family for their rich content, low toxicity, excellent structural topologies, and fascinating photocatalytic activity [179]. The variety of Ti-MOFs has been expanded by manipulating the synthesis parameters of organic ligands and titanium precursors. Nowadays, Ti-MOF derived materials showed high capability in the fields of energy conversion because of their stability, porosity, and regular component arrangement [177]. For instance, preparing Au/TiO2 by pyrolyzed Au/NH2-MIL-125 boosted CO2 reduction into CH4 [180]. It was reported that NH2-MIL-125 (Ti) reduced CO2 into HCOO in presence of TEOA as electron donor [181]. Zhang et al. synthesized Cu-NH2-MIL-125 (Ti) that showed improved light absorption and ameliorated the charge separation, supported with an extended stability throughout four photocatalysis cycles [182]. The photocatalytic conversion rate of CdS-MIL-125 (Ti) was enhanced by the improved light absorption and e/h pairs separation [183]. Moreover, the reports showed that coupling narrow band-gap semiconductors with Ti-MOFs boosts the photosensitive impact and enhances light absorption capacity [184]. Yang et al. proposed a study about ternary heterostructured MIL-125/Ag/g-C3N4 nanocomposites that showed efficient photoreduction in visible light [185]. Many other applications in different fields of renewable energy have received extensive attention due to the promising results via reducing the recombination centers, controlling reactive sites and enhancing the light absorption.

8. Conclusions

TiO2 has been one of the most investigated materials in photocatalysis and studies are in progress among the scientific community to address the unresolved issues. Crystal facet engineering is an important strategy for optimizing both reactivity and selectivity, Researchers have investigated several fabrication routes to control the crystal facets’ type and density. Studies have been able to synthesize photocatalysts with defined facets and their effects on selectivity of product formation have been investigated both theoretically and experimentally, controlling the ratio of different facets and achieving a dominant orientation still remains a challenge.
The nanostructures of TiO2 in its several types and morphologies have been fabricated to explore the novel electronic and the optical properties. Anodization is the most common technique to synthesize vertically aligned nanotubes, however, several other of the 1D and 2D structures have been synthesized using various chemical approaches such as solution growth, hydrothermal and sol-gel process. In particular, designing anatase nanocrystals with the most active facet of (001) helped in scrutinizing catalytic activity in a systematic way.
These morphologies and preparation methods directly control the photocatalytic performance of the synthesized nanomaterials in the form of band gap, electronic structure, light absorption, and surface adsorption. These property enhancements have been explained by the existence of Ti3+, oxygen vacancies, active surface area, charge separation and photocatalysis response. Moreover, owing to the reduction of the e-h recombination, the subsequent lifetime enhancement, the charge transfer kinetics is enhanced as well.
Nevertheless, there are still several challenges remain with the TiO2 nanostructures to achieve higher and better efficiency in photocatalysis. Taking into consideration the different modification approaches of these structures through the synthesis via different methods, there is no unambiguous approach to achieve all properties in the best possible values and hence the synthesis approaches depend on desired morphology or application. However, more investigations are required to achieve in-depth understating of the structure-property relationships in various types and forms of TiO2, and to address the issues such as recombination of photo generated charges and extending the light absorption. The heterojunction approach is expected to play a big role in scaling up and commercially viable technologies. By a suitable selection of the heterojunction component (metal or semiconductor), several drawbacks of TiO2 such as narrow light absorption, higher recombination, surface area for molecular adsorption can be addressed. Further, the regeneration of the photocatalysts and slowing the rate of degradation are severe challenges for practical and scalable implementation. Hence, more theoretical and experimental studies are required to reach comprehensive understanding of TiO2 nanostructures and further research needed to promote its properties.

Author Contributions

Conceptualization, T.F. and H.L.; writing—original draft preparation, T.F., S.C.R., S.R. and H.L.; writing—review and editing, T.F., S.C.R., S.R. and H.L.; visualization, T.F.; supervision, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Ministry of Science and Technology of Taiwan, grant no. 110-2221-E-110-042.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mardani, A.; Streimikiene, D.; Cavallaro, F.; Loganathan, N.; Khoshnoudi, M. Carbon dioxide (CO2) emissions and economic growth: A systematic review of two decades of research from 1995 to 2017. Sci. Total Environ. 2017, 649, 31–49. [Google Scholar] [CrossRef] [PubMed]
  2. Pipes, R.; Bhargav, A.; Manthiram, A. Phenyl Disulfide Additive for Solution-Mediated Carbon Dioxide Utilization in Li−CO2 Batteries. Adv. Energy Mater. 2019, 9, 1900453. [Google Scholar] [CrossRef]
  3. Song, C.; Liu, Q.; Deng, S.; Li, H.; Kitamura, Y. Cryogenicbased CO2 capture technologies: State-of-the-art developments and current challenges. Renew. Sustain. Energy Rev. 2019, 101, 265–278. [Google Scholar] [CrossRef]
  4. Duan, Y.-X.; Meng, F.-L.; Liu, K.-H.; Yi, S.-S.; Li, S.-J.; Yan, J.-M.; Jiang, Q. Amorphizing of Cu nanoparticles toward highly efficient and robust electrocatalyst for CO2 reduction to liquid fuels with high faradaic efficiencies. Adv. Mater. 2018, 30, 1706194. [Google Scholar] [CrossRef] [PubMed]
  5. Sun, Z.; Ma, T.; Tao, H.; Fan, Q.; Han, B. Fundamentals and challenges of electrochemical CO2 reduction using two-dimensional materials. Chem 2017, 3, 560–587. [Google Scholar] [CrossRef] [Green Version]
  6. Zhang, N.; Long, R.; Gao, C.; Xiong, Y. Recent progress on advanced design for photoelectrochemical reduction of CO2 to fuels. Sci. China Mater. 2018, 61, 771–805. [Google Scholar] [CrossRef] [Green Version]
  7. Yang, D.; Yu, H.; He, T.; Zuo, S.; Liu, X.; Yang, H.; Ni, B.; Li, H.; Gu, L.; Wang, D.; et al. Visible-light-switched electron transfer over single porphyrin-metal atom center for highly selective electroreduction of carbon dioxide. Nat. Commun. 2019, 10, 3844. [Google Scholar] [CrossRef] [Green Version]
  8. Liu, X.; Kang, F.; Hu, C.; Wang, L.; Xu, Z.; Zheng, D.; Gong, W.; Lu, Y.; Ma, Y.; Wang, J. A genetically encoded photosensitizer protein facilitates the rational design of a miniature photocatalytic CO2-reducing enzyme. Nat. Chem. 2018, 10, 1201. [Google Scholar] [CrossRef]
  9. 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]
  10. Lisovskaya, A.; Bartels, D.M. Reduction of CO2 by hydrated electrons in high temperature water. Radiat. Phys. Chem. 2019, 158, 61–63. [Google Scholar] [CrossRef]
  11. Ferrah, D.; Haines, A.R.; Galhenage, R.P.; Bruce, J.P.; Babore, A.D.; Hunt, A.; Waluyo, I.; Hemminger, J.C. Wet Chemical Growth and Thermocatalytic Activity of Cu-Based Nanoparticles Supported on TiO2 Nanoparticles/HOPG: In Situ Ambient Pressure XPS Study of the CO2 Hydrogenation Reaction. ACS Catal. 2019, 9, 6783–6802. [Google Scholar] [CrossRef]
  12. Huang, J.; Buonsanti, R. Colloidal nanocrystals as heterogeneous catalysts for electrochemical CO2 conversion. Chem. Mater. 2019, 31, 13–25. [Google Scholar] [CrossRef] [Green Version]
  13. Kang, M.J.; Kim, C.W.; Pawar, A.U.; Cha, H.G.; Ji, S.; Cai, W.-B.; Kang, Y.S. Selective Alcohol on Dark Cathode by Photoelectrochemical CO2 Valorization and Their in-situ Characterization. ACS Energy Lett. 2019, 4, 1549–1555. [Google Scholar] [CrossRef]
  14. Ji, Y.; Luo, Y. Direct Donation of Protons from H2O to CO2 in Artificial Photosynthesis on the Anatase TiO2 (101) Surface. J. Phys. Chem. Commun. 2019, 123, 3019–3023. [Google Scholar] [CrossRef]
  15. Xu, S.; Carter, E.A. Theoretical insights into heterogeneous (Photo) electrochemical CO2 reduction. Chem. Rev. 2019, 119, 6631–6669. [Google Scholar] [CrossRef]
  16. Raizada, P.; Soni, V.; Kumar, A.; Singh, P.; Khan, A.A.P.; Asiri, A.M.; Thakur, V.K.; Nguyen, V.-H. Surface defect engineering of metal oxides photocatalyst for energy application and water treatment. J. Mater. 2021, 7, 388–418. [Google Scholar] [CrossRef]
  17. Wang, P.; Wang, S.; Wang, H.; Wu, Z.; Wang, L. Recent progress on photo-electrocatalytic reduction of carbon dioxide. Part. Part. Syst. Charact. 2018, 35, 1700371. [Google Scholar] [CrossRef]
  18. Wang, L.; Chen, W.; Zhang, D.; Du, Y.; Amal, R.; Qiao, S.; Wu, J.; Yin, Z. Surface strategies for catalytic CO2 reduction: From two-dimensional materials to nanoclusters to single atoms. Chem. Soc. Rev. 2019, 48, 5310–5349. [Google Scholar] [CrossRef]
  19. Kong, L.; Qiao, J.; Ruan, Q.; Wang, H.; Xi, X.; Zha, W.; Zhou, Z.; He, W.; Zhang, W.; Sun, Z. A very low charge potential for zinc-air battery promoted by photochemical effect of triazine-based conjugated polymer nanolayer coated TiO2. J. Power Sources 2022, 536, 231507. [Google Scholar] [CrossRef]
  20. Pawar, A.U.; Kim, C.W.; Nguyen-Le, M.-T.; Kang, Y.S. General review on the components and parameters of photoelectrochemical system for CO2 reduction with in situ analysis. ACS Sustain. Chem. Eng. 2019, 7, 7431–7455. [Google Scholar] [CrossRef]
  21. 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]
  22. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  23. Tamgadge, R.M.; Shukla, A. Fluorine-doped anatase for improved supercapacitor electrode. Electrochim. Acta 2018, 289, 342–353. [Google Scholar] [CrossRef]
  24. Hussain, I.; Chowdhury, A.R.; Jaksik, J.; Grissom, G.; Touhami, A.; Ibrahim, E.E.; Schauer, M.; Okoli, O.; Uddin, M.J. Conductive glass free carbon nanotube micro yarn based perovskite solar cells. Appl. Surf. Sci. 2019, 478, 327–333. [Google Scholar] [CrossRef]
  25. Mino, L.; Ferrari, A.M.; Lacivita, V.; Spoto, G.; Bordiga, S.; Zecchina, A. CO adsorption on anatase nanocrystals: A combined experimental and periodic DFT study. J. Phys. Chem. C 2011, 115, 7694–7700. [Google Scholar] [CrossRef]
  26. Das, S.; Dhara, S. Chemical Solution Synthesis for Materials Design and Thin Film Device Applications; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
  27. Mino, L.; Zecchina, A.; Martra, G.; Rossi, A.M.; Spoto, G. A surface science approach to TiO2 P25 photocatalysis: An in situ FTIR study of phenol photodegradation at controlled water coverages from sub-monolayer to multilayer. Appl. Catal. B Environ. 2016, 196, 135–141. [Google Scholar] [CrossRef]
  28. Li, G.; Fang, K.; Ou, Y.; Yuan, W.; Yang, H.; Zhang, Z.; Wang, Y. Surface study of the reconstructed anatase TiO2 (001) surface. Prog. Nat. Sci. Mater. Int. 2021, 31, 1–13. [Google Scholar] [CrossRef]
  29. Liu, Y.; Du, Y.E.; Bai, Y.; An, J.; Li, J.; Yang, X.; Feng, Q. Facile Synthesis of {101},{010} and [111]-Faceted Anatase-TiO2 Nanocrystals Derived from Porous Metatitanic Acid H2TiO3 for Enhanced Photocatalytic Performance. ChemistrySelect 2018, 3, 2867–2876. [Google Scholar] [CrossRef]
  30. Mollavali, M.; Rohani, S.; Elahifard, M.; Behjatmanesh-Ardakani, R.; Nourany, M. Band gap reduction of (Mo+ N) co-doped TiO2 nanotube arrays with a significant enhancement in visible light photo-conversion: A combination of experimental and theoretical study. Int. J. Hydrog. Energy 2021, 46, 21475–21498. [Google Scholar] [CrossRef]
  31. Grissom, G.; Jaksik, J.; McEntee, M.; Durke, E.M.; Aishee, S.T.; Cua, M.; Okoli, O.; Touhami, A.; Moore, H.J.; Uddin, M.J. Three-dimensional carbon nanotube yarn based solid state solar cells with multiple sensitizers exhibit high energy conversion efficiency. Solar Energy 2018, 171, 16–22. [Google Scholar] [CrossRef]
  32. Likodimos, V. Photonic crystal-assisted visible light activated TiO2 photocatalysis. Appl. Catal. B Environ. 2018, 230, 269–303. [Google Scholar] [CrossRef]
  33. Cravanzola, S.; Cesano, F.; Gaziano, F.; Scarano, D. Sulfur-doped TiO2: Structure and surface properties. Catalysts 2017, 7, 214. [Google Scholar] [CrossRef] [Green Version]
  34. Humayun, M.; Raziq, F.; Khan, A.; Luo, W. Modification strategies of TiO2 for potential applications in photocatalysis: A critical review. Green Chem. Lett. Rev. 2018, 11, 86–102. [Google Scholar] [CrossRef] [Green Version]
  35. Uddin, M.J.; Daramola, D.E.; Velasquez, E.; Dickens, T.J.; Yan, J.; Hammel, E.; Cesano, F.; Okoli, O.I. A high efficiency 3D photovoltaic microwire with carbon nanotubes (CNT)-quantum dot (QD) hybrid interface. Phys. Status Solidi (RRL)–Rapid Res. Lett. 2014, 8, 898–903. [Google Scholar] [CrossRef]
  36. Jia, S.; Li, X.; Zhang, B.; Yang, J.; Zhang, S.; Li, S.; Zhang, Z. TiO2/CuS heterostructure nanowire array photoanodes toward water oxidation: The role of CuS. Appl. Surf. Sci. 2019, 463, 829–837. [Google Scholar] [CrossRef]
  37. Uddin, M.J.; Cesano, F.; Chowdhury, A.R.; Trad, T.; Cravanzola, S.; Martra, G.; Mino, L.; Zecchina, A.; Scarano, D. Surface structure and phase composition of TiO2 P25 particles after thermal treatments and HF etching. Front. Mater. 2020, 7, 192. [Google Scholar] [CrossRef]
  38. Peng, Y.-K.; Chou, H.-L.; Tsang, S.C.E. Differentiating surface titanium chemical states of anatase TiO2 functionalized with various groups. Chem. Sci. 2018, 9, 2493–2500. [Google Scholar] [CrossRef] [Green Version]
  39. Zou, Y.; Gao, G.; Wang, Z.; Shi, J.-W.; Wang, H.; Ma, D.; Fan, Z.; Chen, X.; Wang, Z.; Niu, C. Formation mechanism of rectangular-ambulatory-plane TiO2 plates: An insight into the role of hydrofluoric acid. Chem. Commun. 2018, 54, 7191–7194. [Google Scholar] [CrossRef]
  40. Ma, M.J.; Li, W.; Dambournet, D. Solution-Based Synthesis of Nano-Sized TiO2 Anatase in Fluorinating Media. In Modern Synthesis Processes and Reactivity of Fluorinated Compounds; Elsevier: Amsterdam, The Netherlands, 2017; pp. 651–669. [Google Scholar]
  41. Wang, A.; Wu, S.; Dong, J.; Wang, R.; Wang, J.; Zhang, J.; Zhong, S.; Bai, S. Interfacial facet engineering on the Schottky barrier between plasmonic Au and TiO2 in boosting the photocatalytic CO2 reduction under ultraviolet and visible light irradiation. Chem. Eng. J. 2021, 404, 127145. [Google Scholar] [CrossRef]
  42. Dong, J.; Wang, Z.; Cao, H.; Xue, J.; Liu, C.; Sun, S.; Gao, C.; Zhu, X.; Bao, J. Independent Cr2O3 functions as efficient cocatalyst on the crystal facets engineered TiO2 for photocatalytic CO2 reduction. Appl. Surf. Sci. 2021, 554, 149634. [Google Scholar] [CrossRef]
  43. Butburee, T.; Kotchasarn, P.; Hirunsit, P.; Sun, Z.; Tang, Q.; Khemthong, P.; Sangkhun, W.; Thongsuwan, W.; Kumnorkaew, P.; Wang, H. New understanding of crystal control and facet selectivity of titanium dioxide ruling photocatalytic performance. J. Mater. Chem. A 2019, 7, 8156–8166. [Google Scholar] [CrossRef]
  44. Li, X. Synthesis and Metal-Insulator Transition Properties of Vanadium Dioxide Nanostructures; The Pennsylvania State University: State College, PA, USA, 2019. [Google Scholar]
  45. Du, Y.-E.; Niu, X.; He, X.; Hou, K.; Liu, H.; Zhang, C. Synthesis and photocatalytic activity of TiO2/CdS nanocomposites with co-exposed anatase highly reactive facets. Molecules 2021, 26, 6031. [Google Scholar] [CrossRef] [PubMed]
  46. Du, Y.-E.; Niu, X.; He, J.; Liu, L.; Liu, Y.; Chen, C.; Yang, X.; Feng, Q. Hollow square rodlike microtubes composed of anatase nanocuboids with coexposed {100},{010}, and {001} facets for improved photocatalytic performance. ACS Omega 2020, 5, 14147–14156. [Google Scholar] [CrossRef] [PubMed]
  47. Wen, C.Z.; Zhou, J.Z.; Jiang, H.B.; Hu, Q.H.; Qiao, S.Z.; Yang, H.G. Synthesis of micro-sized titanium dioxide nanosheets wholly exposed with high-energy {001} and {100} facets. Chem. Commun. 2011, 47, 4400–4402. [Google Scholar] [CrossRef] [Green Version]
  48. Li, W. Sol-Gel Synthesis of TiO2 Anatase in a Fluorinated Medium and Its Applications as Negative Electrode for Li+ and Na+ Batteries. Ph.D. Thesis, Paris 6, Doctoral School of Physical Chemistry and Analytical Chemistry of Paris Center, Paris, France, 2015. [Google Scholar]
  49. Bokare, A.; Erogbogbo, F. Photocatalysis and Li-Ion Battery Applications of {001} Faceted Anatase TiO2-Based Composites. J 2021, 4, 500–530. [Google Scholar] [CrossRef]
  50. Du, D.-J.; Du, Y.-E.; Yue, W.-B.; Yang, X.-J. Lithium storage performance of {010}-faceted and [111]-faceted anatase TiO2 nanocrystals. J. Cent. South Univ. 2019, 26, 1530–1539. [Google Scholar] [CrossRef]
  51. Maitani, M.M.; Tateyama, A.; Boix, P.P.; Han, G.; Nitta, A.; Ohtani, B.; Mathews, N.; Wada, Y. Effects of energetics with {001} facet-dominant anatase TiO2 scaffold on electron transport in CH3NH3PbI3 perovskite solar cells. Electrochim. Acta 2019, 300, 445–454. [Google Scholar] [CrossRef]
  52. Liu, X.; Du, G.; Li, M. True photoreactivity origin of Ti3+-doped anatase TiO2 crystals with respectively dominated exposed {001},{101}, and {100} facets. ACS Omega 2019, 4, 14902–14912. [Google Scholar] [CrossRef] [Green Version]
  53. Liu, G.; Wang, T.; Zhou, W.; Meng, X.; Zhang, H.; Liu, H.; Kako, T.; Ye, J. Crystal-facet-dependent hot-electron transfer in plasmonic-Au/semiconductor heterostructures for efficient solar photocatalysis. J. Mater. Chem. C 2015, 3, 7538–7542. [Google Scholar] [CrossRef] [Green Version]
  54. Liu, L.; Liu, Z.; Liu, A.; Gu, X.; Ge, C.; Gao, F.; Dong, L. Engineering the TiO2–graphene interface to enhance photocatalytic H2 production. ChemSusChem 2014, 7, 618–626. [Google Scholar] [CrossRef]
  55. Pathakoti, K.; Manubolu, M.; Hwang, H.-M. Nanotechnology applications for environmental industry. In Handbook of Nanomaterials for Industrial Applications; Elsevier: Amsterdam, The Netherlands, 2018; pp. 894–907. [Google Scholar]
  56. Kozak, M.; Mazierski, P.; Żebrowska, J.; Kobylański, M.; Klimczuk, T.; Lisowski, W.; Trykowski, G.; Nowaczyk, G.; Zaleska-Medynska, A. Electrochemically obtained TiO2/CuxOy nanotube arrays presenting a photocatalytic response in processes of pollutants degradation and bacteria inactivation in aqueous phase. Catalysts 2018, 8, 237. [Google Scholar] [CrossRef] [Green Version]
  57. Qu, J.; Sha, L.; Wu, C.; Zhang, Q. Applications of mechanochemically prepared layered double hydroxides as adsorbents and catalysts: A mini-review. Nanomaterials 2019, 9, 80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Xue, Y.; Wu, Z.; He, X.; Yang, X.; Chen, X.; Gao, Z. Constructing a Z-scheme heterojunction of egg-like core@ shell CdS@ TiO2 photocatalyst via a facile reflux method for enhanced photocatalytic performance. Nanomaterials 2019, 9, 222. [Google Scholar] [CrossRef] [Green Version]
  59. Shahvaranfard, F. Modification of Low Dimensional Nanostructured TiO2 for Energy Application. Ph.D. Thesis, Friedrich-Alexander-Universität Erlangen-Nürnberg (FAU), Erlangen, Germany, 2022. [Google Scholar]
  60. Lee, K.; Mazare, A.; Schmuki, P. One-dimensional titanium dioxide nanomaterials: Nanotubes. Chem. Rev. 2014, 114, 9385–9454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Haider, A.J.; Jameel, Z.N.; Al-Hussaini, I.H. Review on: Titanium dioxide applications. Energy Procedia 2019, 157, 17–29. [Google Scholar] [CrossRef]
  62. Ranjan, S.; Ramalingam, C. Titanium dioxide nanoparticles induce bacterial membrane rupture by reactive oxygen species generation. Environ. Chem. Lett. 2016, 14, 487–494. [Google Scholar] [CrossRef]
  63. Liu, Q.; Huang, J.; Tang, H.; Yu, X.; Shen, J. Construction 0D TiO2 nanoparticles/2D CoP nanosheets heterojunctions for enhanced photocatalytic H2 evolution activity. J. Mater. Sci. Technol. 2020, 56, 196–205. [Google Scholar] [CrossRef]
  64. Tahir, M.; Amin, N.S. Indium-doped TiO2 nanoparticles for photocatalytic CO2 reduction with H2O vapors to CH4. Appl. Catal. B Environ. 2015, 162, 98–109. [Google Scholar] [CrossRef]
  65. 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]
  66. Tseng, I.-H.; Chang, W.-C.; Wu, J.C. Photoreduction of CO2 using sol–gel derived titania and titania-supported copper catalysts. Appl. Catal. B Environ. 2002, 37, 37–48. [Google Scholar] [CrossRef]
  67. Wang, Y.; Lai, Q.; Zhang, F.; Shen, X.; Fan, M.; He, Y.; Ren, S. High efficiency photocatalytic conversion of CO2 with H2O over Pt/TiO2 nanoparticles. RSC Adv. 2014, 4, 44442–44451. [Google Scholar] [CrossRef]
  68. Liu, Y.; Lee, C.-C.; Horn, M.W.; Lee, H. Toward efficient photocatalysts for light-driven CO2 reduction: TiO2 nanostructures decorated with perovskite quantum dots. Nano Express 2021, 2, 020003. [Google Scholar] [CrossRef]
  69. Xu, Y.-F.; Yang, M.-Z.; Chen, B.-X.; Wang, X.-D.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y. A CsPbBr3 perovskite quantum dot/graphene oxide composite for photocatalytic CO2 reduction. J. Am. Chem. Soc. 2017, 139, 5660–5663. [Google Scholar] [CrossRef] [PubMed]
  70. Cheng, M.; Yang, S.; Chen, R.; Zhu, X.; Liao, Q.; Huang, Y. Copper-decorated TiO2 nanorod thin films in optofluidic planar reactors for efficient photocatalytic reduction of CO2. Int. J. Hydrog. Energy 2017, 42, 9722–9732. [Google Scholar] [CrossRef]
  71. Devi, A.D.; Pushpavanam, S.; Singh, N.; Verma, J.; Kaur, M.P.; Roy, S.C. Enhanced methane yield by photoreduction of CO2 at moderate temperature and pressure using Pt coated, graphene oxide wrapped TiO2 nanotubes. Results Eng. 2022, 14, 100441. [Google Scholar] [CrossRef]
  72. 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]
  73. Ouyang, W.; Teng, F.; He, J.H.; Fang, X. Enhancing the photoelectric performance of photodetectors based on metal oxide semiconductors by charge-carrier engineering. Adv. Funct. Mater. 2019, 29, 1807672. [Google Scholar] [CrossRef]
  74. Wang, X.; Wang, Z.; Jiang, X.; Tao, J.; Gong, Z.; Cheng, Y.; Zhang, M.; Yang, L.; Lv, J.; He, G. Silver-decorated TiO2 nanorod array films with enhanced photoelectrochemical and photocatalytic properties. J. Electrochem. Soc. 2016, 163, H943. [Google Scholar] [CrossRef]
  75. Qu, J.; Lai, C. One-dimensional nanostructures as photoanodes for dye-sensitized solar cells. J. Nanomater. 2013, 2013, 2. [Google Scholar] [CrossRef]
  76. Subramanian, A.; Pan, Z.; Li, H.; Zhou, L.; Li, W.; Qiu, Y.; Xu, Y.; Hou, Y.; Muzi, C.; Zhang, Y. Synergistic promotion of photoelectrochemical water splitting efficiency of TiO2 nanorods using metal-semiconducting nanoparticles. Appl. Surf. Sci. 2017, 420, 631–637. [Google Scholar] [CrossRef]
  77. Jiang, L.; He, J.; Yang, Y.; Mao, D.; Chen, D.; Wang, W.; Chen, Y.; Sharma, V.K.; Wang, J. Enhancing visible-light photocatalytic activity of hard-biotemplated TiO2: From macrostructural morphology replication to microstructural building units design. J. Alloys Compd. 2022, 898, 162886. [Google Scholar] [CrossRef]
  78. Ghosh, M.; Liu, J.; Chuang, S.S.; Jana, S.C. Fabrication of hierarchical V2O5 nanorods on TiO2 nanofibers and their enhanced photocatalytic activity under visible light. ChemCatChem 2018, 10, 3305–3318. [Google Scholar] [CrossRef]
  79. Attar, A.S.; Ghamsari, M.S.; Hajiesmaeilbaigi, F.; Mirdamadi, S.; Katagiri, K.; Koumoto, K. Sol–gel template synthesis and characterization of aligned anatase-TiO2 nanorod arrays with different diameter. Mater. Chem. Phys. 2009, 113, 856–860. [Google Scholar] [CrossRef]
  80. 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]
  81. Ping, G.; Wang, C.; Chen, D.; Liu, S.; Huang, X.; Qin, L.; Huang, Y.; Shu, K. Fabrication of self-organized TiO2 nanotube arrays for photocatalytic reduction of CO2. J. Solid State Electrochem. 2013, 17, 2503–2510. [Google Scholar] [CrossRef]
  82. Feng, X.; Sloppy, J.D.; LaTempa, T.J.; Paulose, M.; Komarneni, S.; Bao, N.; Grimes, C.A. Synthesis and deposition of ultrafine Pt nanoparticles within high aspect ratio TiO2 nanotube arrays: Application to the photocatalytic reduction of carbon dioxide. J. Mater. Chem. 2011, 21, 13429–13433. [Google Scholar] [CrossRef]
  83. Li, X.; Liu, H.; Luo, D.; Li, J.; Huang, Y.; Li, H.; Fang, Y.; Xu, Y.; Zhu, L. Adsorption of CO2 on heterostructure CdS (Bi2S3)/TiO2 nanotube photocatalysts and their photocatalytic activities in the reduction of CO2 to methanol under visible light irradiation. Chem. Eng. J. 2012, 180, 151–158. [Google Scholar] [CrossRef]
  84. Li, Y.; Wang, C.; Song, M.; Li, D.; Zhang, X.; Liu, Y. TiO2−x/CoOx photocatalyst sparkles in photothermocatalytic reduction of CO2 with H2O steam. Appl. Catal. B Environ. 2019, 243, 760–770. [Google Scholar] [CrossRef]
  85. 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]
  86. Tahir, M.; Tahir, B.; Amin, N.A.S.; Zakaria, Z.Y. Photo-induced reduction of CO2 to CO with hydrogen over plasmonic Ag-NPs/TiO2 NWs core/shell hetero-junction under UV and visible light. J. CO2 Util. 2017, 18, 250–260. [Google Scholar] [CrossRef]
  87. 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]
  88. Su, K.-Y.; Chen, C.-Y.; Wu, R.-J. Preparation of Pd/TiO2 nanowires for the photoreduction of CO2 into renewable hydrocarbon fuels. J. Taiwan Inst. Chem. Eng. 2019, 96, 409–418. [Google Scholar] [CrossRef]
  89. Song, G.; Xin, F.; Yin, X. Photocatalytic reduction of carbon dioxide over ZnFe2O4/TiO2 nanobelts heterostructure in cyclohexanol. J. Colloid Interface Sci. 2015, 442, 60–66. [Google Scholar] [CrossRef] [PubMed]
  90. Razzaq, A.; Grimes, C.A.; In, S.-I. Facile fabrication of a noble metal-free photocatalyst: TiO2 nanotube arrays covered with reduced graphene oxide. Carbon 2016, 98, 537–544. [Google Scholar] [CrossRef]
  91. Zhang, H. Ultrathin two-dimensional nanomaterials. ACS Nano 2015, 9, 9451–9469. [Google Scholar] [CrossRef]
  92. Madkour, L.H. Carbon Nanomaterials and Two-Dimensional Transition Metal Dichalcogenides (2D TMDCs). In Nanoelectronic Materials; Springer: Berlin/Heidelberg, Germany, 2019; pp. 165–245. [Google Scholar]
  93. Tan, C.; Cao, X.; Wu, X.-J.; He, Q.; Yang, J.; Zhang, X.; Chen, J.; Zhao, W.; Han, S.; Nam, G.-H. Recent advances in ultrathin two-dimensional nanomaterials. Chem. Rev. 2017, 117, 6225–6331. [Google Scholar] [CrossRef]
  94. Late, D.J.; Bhat, A.; Rout, C.S. Fundamentals and properties of 2D materials in general and sensing applications. In Fundamentals and Sensing Applications of 2D Materials; Elsevier: Amsterdam, The Netherlands, 2019; pp. 5–24. [Google Scholar]
  95. Chen, F.; Ma, T.; Zhang, T.; Zhang, Y.; Huang, H. Atomic-level charge separation strategies in semiconductor-based photocatalysts. Adv. Mater. 2021, 33, 2005256. [Google Scholar] [CrossRef]
  96. Liu, Y.; Zou, J.; Guo, B.; Ren, Y.; Wang, Z.; Song, Y.; Yu, Y.; Wu, L. Selective photocatalytic oxidation of thioanisole on DUT-67 (Zr) mediated by surface coordination. Langmuir 2020, 36, 2199–2208. [Google Scholar] [CrossRef]
  97. Sadeghfar, F.; Zalipour, Z.; Taghizadeh, M.; Taghizadeh, A.; Ghaedi, M. Photodegradation processes. In Interface Science and Technology; Elsevier: Amsterdam, The Netherlands, 2021; Volume 32, pp. 55–124. [Google Scholar]
  98. 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]
  99. Zhou, S.; Liu, Y.; Li, J.; Wang, Y.; Jiang, G.; Zhao, Z.; Wang, D.; Duan, A.; Liu, J.; Wei, Y. Facile in situ synthesis of graphitic carbon nitride (g-C3N4)-N-TiO2 heterojunction as an efficient photocatalyst for the selective photoreduction of CO2 to CO. Appl. Catal. B Environ. 2014, 158, 20–29. [Google Scholar] [CrossRef]
  100. 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 Environ. 2019, 242, 369–378. [Google Scholar] [CrossRef]
  101. 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]
  102. 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 Environ. 2019, 244, 919–930. [Google Scholar] [CrossRef]
  103. 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]
  104. Yuan, L.; Lu, K.-Q.; Zhang, F.; Fu, X.; Xu, Y.-J. Unveiling the interplay between light-driven CO2 photocatalytic reduction and carbonaceous residues decomposition: A case study of Bi2WO6-TiO2 binanosheets. Appl. Catal. B Environ. 2018, 237, 424–431. [Google Scholar] [CrossRef]
  105. Reddy, K.M.; Manorama, S.V.; Reddy, A.R. Bandgap studies on anatase titanium dioxide nanoparticles. Mater. Chem. Phys. 2003, 78, 239–245. [Google Scholar] [CrossRef]
  106. Munir, S.; Shah, S.M.; Hussain, H. Effect of carrier concentration on the optical band gap of TiO2 nanoparticles. Mater. Des. 2016, 92, 64–72. [Google Scholar] [CrossRef]
  107. Wang, X.; Li, Z.; Shi, J.; Yu, Y. One-dimensional titanium dioxide nanomaterials: Nanowires, nanorods, and nanobelts. Chem. Rev. 2014, 114, 9346–9384. [Google Scholar] [CrossRef]
  108. Lee, M.; Seo, Y.; Shin, H.S.; Jo, C.; Ryoo, R. Anatase TiO2 nanosheets with surface acid sites for Friedel–Crafts alkylation. Microporous Mesoporous Mater. 2016, 222, 185–191. [Google Scholar] [CrossRef]
  109. Xu, C.; Zhang, Z.; Zhang, S.; Si, H.; Ma, S.; Fan, W.; Xiong, Z.; Liao, Q.; Sattar, A.; Kang, Z. Manipulation of perovskite crystallization kinetics via Lewis base additives. Adv. Funct. Mater. 2021, 31, 2009425. [Google Scholar] [CrossRef]
  110. Sheng, Y.; Wei, Z.; Miao, H.; Yao, W.; Li, H.; Zhu, Y. Enhanced organic pollutant photodegradation via adsorption/photocatalysis synergy using a 3D g-C3N4/TiO2 free-separation photocatalyst. Chem. Eng. J. 2019, 370, 287–294. [Google Scholar] [CrossRef]
  111. Yang, K.; Cheng, G.; Chen, R.; Zhao, K.; Liang, Y.; Li, W.; Han, C. Simply Coupling TiO2 Nanospheres with Cu2O Particles to Boost the Photocatalytic Hydrogen Evolution through p–n Heterojunction-Induced Charge Transfer. Energy Technol. 2022, 10, 2100259. [Google Scholar] [CrossRef]
  112. Liu, Y.; Zheng, X.; Yang, Y.; Li, J.; Liu, W.; Shen, Y.; Tian, X. Photocatalytic Hydrogen Evolution Using Ternary-Metal-Sulfide/TiO2 Heterojunction Photocatalysts. ChemCatChem 2022, 14, e202101439. [Google Scholar] [CrossRef]
  113. Guo, Q.; Huang, Y.; Xu, H.; Luo, D.; Huang, F.; Gu, L.; Wei, Y.; Zhao, H.; Fan, L.; Wu, J. The effects of solvent on photocatalytic properties of Bi2WO6/TiO2 heterojunction under visible light irradiation. Solid State Sci. 2018, 78, 95–106. [Google Scholar] [CrossRef]
  114. Shang, M.; Wang, W.; Zhang, L.; Sun, S.; Wang, L.; Zhou, L. 3D Bi2WO6/TiO2 hierarchical heterostructure: Controllable synthesis and enhanced visible photocatalytic degradation performances. J. Phys. Chem. C 2009, 113, 14727–14731. [Google Scholar] [CrossRef]
  115. Li, Q.; Zong, L.; Li, C.; Yang, J. Reprint of “Photocatalytic reduction of CO2 on MgO/TiO2 nanotube films”. Appl. Surf. Sci. 2014, 319, 16–20. [Google Scholar] [CrossRef]
  116. Ran, H.; Fan, J.; Zhang, X.; Mao, J.; Shao, G. Enhanced performances of dye-sensitized solar cells based on Au-TiO2 and Ag-TiO2 plasmonic hybrid nanocomposites. Appl. Surf. Sci. 2018, 430, 415–423. [Google Scholar] [CrossRef]
  117. Yao, G.-Y.; Zhao, Z.-Y.; Liu, Q.-L.; Dong, X.-D.; Zhao, Q.-M. Theoretical calculations for localized surface plasmon resonance effects of Cu/TiO2 nanosphere: Generation, modulation, and application in photocatalysis. Sol. Energy Mater. Sol. Cells 2020, 208, 110385. [Google Scholar] [CrossRef]
  118. Wang, Z.; Lee, H.; Chen, J.; Wu, M.; Leung, D.Y.; Grimes, C.A.; Lu, Z.; Xu, Z.; Feng, S.-P. Synergistic effects of Pd–Ag bimetals and g-C3N4 photocatalysts for selective and efficient conversion of gaseous CO2. J. Power Sources 2020, 466, 228306. [Google Scholar] [CrossRef]
  119. Taghizadeh, A.; Taghizadeh, M.; Sabzehmeidani, M.M.; Sadeghfar, F.; Ghaedi, M. Electronic structure: From basic principles to photocatalysis. In Interface Science and Technology; Elsevier: Amsterdam, The Netherlands, 2021; Volume 32, pp. 1–53. [Google Scholar]
  120. 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]
  121. Saraev, A.A.; Kurenkova, A.Y.; Gerasimov, E.Y.; Kozlova, E.A. Broadening the Action Spectrum of TiO2-Based Photocatalysts to Visible Region by Substituting Platinum with Copper. Nanomaterials 2022, 12, 1584. [Google Scholar] [CrossRef]
  122. 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]
  123. Mankidy, B.D.; Joseph, B.; Gupta, V.K. Photo-conversion of CO2 using titanium dioxide: Enhancements by plasmonic and co-catalytic nanoparticles. Nanotechnology 2013, 24, 405402. [Google Scholar] [CrossRef] [PubMed]
  124. Tayebi, M.; Kolaei, M.; Tayyebi, A.; Masoumi, Z.; Belbasi, Z.; Lee, B.-K. Reduced graphene oxide (RGO) on TiO2 for an improved photoelectrochemical (PEC) and photocatalytic activity. Sol. Energy 2019, 190, 185–194. [Google Scholar] [CrossRef]
  125. Mondal, A.; Prabhakaran, A.; Gupta, S.; Subramanian, V.R. Boosting photocatalytic activity using reduced graphene oxide (RGO)/semiconductor nanocomposites: Issues and future scope. ACS Omega 2021, 6, 8734–8743. [Google Scholar] [CrossRef]
  126. Shin, D.H.; Choi, S.-H. Graphene-based semiconductor heterostructures for photodetectors. Micromachines 2018, 9, 350. [Google Scholar] [CrossRef] [Green Version]
  127. Sun, P.; Zhou, S.; Yang, Y.; Liu, S.; Cao, Q.; Wang, Y.; Wågberg, T.; Hu, G. Artificial chloroplast-like phosphotungstic acid—Iron oxide microbox heterojunctions penetrated by carbon nanotubes for solar photocatalytic degradation of tetracycline antibiotics in wastewater. Adv. Compos. Hybrid Mater. 2022, 1–18. [Google Scholar] [CrossRef]
  128. Padmanabhan, N.T.; Thomas, N.; Louis, J.; Mathew, D.T.; Ganguly, P.; John, H.; Pillai, S.C. Graphene coupled TiO2 photocatalysts for environmental applications: A review. Chemosphere 2021, 271, 129506. [Google Scholar] [CrossRef]
  129. Zhang, J.; Xu, J.; Tao, F. Interface Modification of TiO2 Nanotubes by Biomass-Derived Carbon Quantum Dots for Enhanced Photocatalytic Reduction of CO2. ACS Appl. Energy Mater. 2021, 4, 13120–13131. [Google Scholar] [CrossRef]
  130. Morawski, A.W.; Ćmielewska, K.; Witkowski, K.; Kusiak-Nejman, E.; Pełech, I.; Staciwa, P.; Ekiert, E.; Sibera, D.; Wanag, A.; Gano, M. CO2 Reduction to Valuable Chemicals on TiO2-Carbon Photocatalysts Deposited on Silica Cloth. Catalysts 2021, 12, 31. [Google Scholar] [CrossRef]
  131. Rajaraman, T.S.; Parikh, S.P.; Gandhi, V.G. Black TiO2: A review of its properties and conflicting trends. Chem. Eng. 2020, 389, 123918. [Google Scholar] [CrossRef]
  132. Mezzat, F.; Zaari, H.; El Kenz, A.; Benyoussef, A. Effect of metal and non metal doping of TiO2 on photocatalytic activities: Ab initio calculations. Opt. Quantum Electron. 2021, 53, 1–14. [Google Scholar] [CrossRef]
  133. Jeon, J.P.; Kweon, D.H.; Jang, B.J.; Ju, M.J.; Baek, J.B. Enhancing the photocatalytic activity of TiO2 catalysts. Adv. Sustain. Syst. 2020, 4, 2000197. [Google Scholar] [CrossRef]
  134. Xu, T.; Wang, M.; Wang, T. Effects of N doping on the microstructures and optical properties of TiO2. J. Wuhan Univ. Technol. Mater. Sci. Ed. 2019, 34, 55–63. [Google Scholar] [CrossRef]
  135. Yamazaki, Y.; Toyonaga, T.; Doshita, N.; Mori, K.; Kuwahara, Y.; Yamazaki, S.; Yamashita, H. Crystal Facet Engineering and Hydrogen Spillover-Assisted Synthesis of Defective Pt/TiO2–x Nanorods with Enhanced Visible Light-Driven Photocatalytic Activity. ACS Appl. Mater. Interfaces 2021, 14, 2291–2300. [Google Scholar] [CrossRef]
  136. Chen, X.; Peng, X.; Jiang, L.; Yuan, X.; Fei, J.; Zhang, W. Photocatalytic removal of antibiotics by MOF-derived Ti3+-and oxygen vacancy-doped anatase/rutile TiO2 distributed in a carbon matrix. Chem. Eng. J. 2022, 427, 130945. [Google Scholar] [CrossRef]
  137. Sorcar, S.; Hwang, Y.; Grimes, C.A.; In, S.-I. Highly enhanced and stable activity of defect-induced titania nanoparticles for solar light-driven CO2 reduction into CH4. Mater. Today 2017, 20, 507–515. [Google Scholar] [CrossRef]
  138. Nas, M.S.; Calimli, M.H. Recent Development of Nanoparticle by Green-Conventional Methods and Applications for Corrosion and Fuel Cells. Curr. Nanosci. 2021, 17, 525–539. [Google Scholar] [CrossRef]
  139. Behera, A. Nanomaterials. In Advanced Materials; Springer: Berlin/Heidelberg, Germany, 2022; pp. 77–125. [Google Scholar]
  140. Gupta, T.; Cho, J.; Prakash, J. Hydrothermal synthesis of TiO2 nanorods: Formation chemistry, growth mechanism, and tailoring of surface properties for photocatalytic activities. Mater. Today Chem. 2021, 20, 100428. [Google Scholar] [CrossRef]
  141. Wang, Y.-H.; Rahman, K.H.; Wu, C.-C.; Chen, K.-C. A review on the pathways of the improved structural characteristics and photocatalytic performance of titanium dioxide (TiO2) thin films fabricated by the magnetron-sputtering technique. Catalysts 2020, 10, 598. [Google Scholar] [CrossRef]
  142. Li, Z.; Wang, S.; Wu, J.; Zhou, W. Recent progress in defective TiO2 photocatalysts for energy and environmental applications. Renew. Sustain. Energy Rev. 2022, 156, 111980. [Google Scholar] [CrossRef]
  143. Zhang, X.; Hu, W.; Zhang, K.; Wang, J.; Sun, B.; Li, H.; Qiao, P.; Wang, L.; Zhou, W. Ti3+ self-doped black TiO2 nanotubes with mesoporous nanosheet architecture as efficient solar-driven hydrogen evolution photocatalysts. ACS Sustain. Chem. Eng. 2017, 5, 901. [Google Scholar] [CrossRef]
  144. Wu, J.; Qiao, P.; Li, H.; Xu, Y.; Yang, W.; Yang, F.; Lin, K.; Pan, K.; Zhou, W. Engineering surface defects on two-dimensional ultrathin mesoporous anatase TiO2 nanosheets for efficient charge separation and exceptional solar-driven photocatalytic hydrogen evolution. Mater. Chem. 2020, 8, 82. [Google Scholar] [CrossRef]
  145. Li, Z.; Wang, S.; Xie, Y.; Yang, W.; Tao, B.; Lu, J.; Wu, J.; Qu, Y.; Zhou, W. Surface defects induced charge imbalance for boosting charge separation and solar-driven photocatalytic hydrogen evolution. Colloid Interface Sci. 2021, 596, 21. [Google Scholar] [CrossRef] [PubMed]
  146. Yang, W.T.; Li, M.; Pan, K.; Guo, L.P.; Wu, J.X.; Li, Z.J.; Yang, F.; Lin, K.; Zhou, W. Surface engineering of mesoporous anatase titanium dioxide nanotubes for rapid spatial charge separation on horizontal-Vertical dimensions and efficient solar-driven photocatalytic hydrogen evolution. Colloid Interface Sci. 2021, 586, 83. [Google Scholar] [CrossRef] [PubMed]
  147. Liu, X.; Zhu, G.; Wang, X.; Yuan, X.; Lin, T.; Huang, F. Progress in black titania: A new material for advanced photocatalysis. Adv. Energy Mater. 2016, 6, 1600452. [Google Scholar] [CrossRef]
  148. Shoneye, A.; Sen Chang, J.; Chong, M.N.; Tang, J. Recent progress in photocatalytic degradation of chlorinated phenols and reduction of heavy metal ions in water by TiO2-based catalysts. Int. Mater. Rev. 2022, 67, 47–64. [Google Scholar] [CrossRef]
  149. Janczarek, M.; Kowalska, E. Defective dopant-free TiO2 as an efficient visible light-active photocatalyst. Catalysts 2021, 11, 978. [Google Scholar] [CrossRef]
  150. Dobrosielska, M.; Zieliński, M.; Frydrych, M.; Pietrowski, M.; Marciniak, P.; Martyła, A.; Sztorch, B.; Przekop, R.E. Sol–Gel Approach for Design of Pt/Al2O3-TiO2 System—Synthesis and Catalytic Tests. Ceramics 2021, 4, 667–680. [Google Scholar] [CrossRef]
  151. Lluna-Galán, C.; Izquierdo-Aranda, L.; Adam, R.; Cabrero-Antonino, J.R. Catalytic Reductive Alcohol Etherifications with Carbonyl-Based Compounds or CO2 and Related Transformations for the Synthesis of Ether Derivatives. ChemSusChem 2021, 14, 3744–3784. [Google Scholar] [CrossRef]
  152. Tian, J.; Hu, X.; Yang, H.; Zhou, Y.; Cui, H.; Liu, H. High yield production of reduced TiO2 with enhanced photocatalytic activity. Appl. Surf. Sci. 2016, 360, 43. [Google Scholar] [CrossRef]
  153. Fang, W.; Xing, M.; Zhang, J. A new approach to prepare Ti3+ self-doped TiO2 via NaBH4 reduction and hydrochloric acid treatment. Appl. Catal. B Environ. 2014, 240, 160–161. [Google Scholar] [CrossRef]
  154. Tan, H.; Zhao, Z.; Niu, M.; Mao, C.; Cao, D.; Cheng, D.; Feng, P.; Sun, Z. A facile and versatile method for preparation of colored TiO2 with enhanced solar-driven photocatalytic activity. Nanoscale 2014, 6, 23. [Google Scholar] [CrossRef]
  155. Wang, Z.; Yang, C.; Lin, T.; Yin, H.; Chen, P.; Wan, D.; Xu, F.; Huang, F.; Lin, J.; Xie, X.; et al. H-doped black titania with very high solar absorption and excellent photocatalysis enhanced by localized surface plasmon resonance. Adv. Funct Mater. 2013, 23, 50. [Google Scholar] [CrossRef]
  156. Teng, F.; Li, M.; Gao, C.; Zhang, G.; Zhang, P.; Wang, Y.; Chen, L.; Xie, E. Preparation of black TiO2 by hydrogen plasma assisted chemical vapor deposition and its photocatalytic activity. Appl. Catal. B Environ. 2014, 339, 148–149. [Google Scholar] [CrossRef]
  157. Khan, M.M.; Ansari, S.A.; Pradhan, D.; Ansari, M.O.; Han, D.H.; Lee, J.; Cho, M.H. Band gap engineered TiO2 nanoparticles for visible light induced photoelectrochemical and photocatalytic studies. Mater. Chem. 2014, 2, 44. [Google Scholar]
  158. Liu, N.; Haublein, V.; Zhou, X.M.; Venkatesan, U.; Hartmann, M.; Mackovic, M.; Nakajima, T.; Spiecker, E.; Osvet, A.; Frey, L.; et al. “Black” TiO2 nanotubes formed by high-energy proton implantation show noble-metal-co-catalyst free photocatalytic H2-evolution. Nano Lett. 2015, 15, 20. [Google Scholar] [CrossRef] [Green Version]
  159. Wang, Z.; Yang, C.; Lin, T.; Yin, H.; Chen, P.; Wan, D.; Xu, F.; Huang, F.; Lin, J.; Xie, X.; et al. Visible-light photocatalytic, solar thermal and photoelectrochemical properties of aluminium-reduced black titania. Energy Environ. Sci. 2013, 6, 14. [Google Scholar] [CrossRef]
  160. Zhu, G.; Lin, T.; Lü, X.; Zhao, W.; Yang, C.; Wang, Z.; Yin, H.; Liu, Z.; Huang, F.; Lin, J. Black brookite titania with high solar absorption and excellent photocatalytic performance. Mater. Chem. 2013, 1, 3. [Google Scholar] [CrossRef]
  161. Yin, H.; Lin, T.; Yang, C.; Wang, Z.; Zhu, G.; Xu, T.; Xie, X.; Huang, F.; Jiang, M. Gray TiO2 nanowires synthesized by aluminum-mediated reduction and their excellent photocatalytic activity for water cleaning. Chem. Eur. J. 2013, 19, 6. [Google Scholar] [CrossRef]
  162. Yang, C.; Wang, Z.; Lin, T.; Yin, H.; Lü, X.; Wan, D.; Xu, T.; Zheng, C.; Lin, J.; Huang, F.; et al. Core-shell nanostructured “black” rutile titania as excellent catalyst for hydrogen production enhanced by sulfur doping. Am. Chem. Soc. 2013, 135, 8. [Google Scholar] [CrossRef] [PubMed]
  163. Lin, T.; Yang, C.; Wang, Z.; Yin, H.; Lü, X.; Huang, F.; Lin, J.; Xie, X.; Jiang, M. Effective nonmetal incorporation in black titania with enhanced solar energy utilization. Energy Environ. Sci. 2014, 7, 72. [Google Scholar] [CrossRef]
  164. Zhao, Z.; Tan, H.; Zhao, H.; Lv, Y.; Zhou, L.-J.; Song, Y.; Sun, Z. Reduced TiO2 rutile nanorods with well-defined facets and their visible-light photocatalytic activity. Chem. Commun. 2014, 50, 7. [Google Scholar]
  165. Sinhamahapatra, A.; Jeon, J.-P.; Yu, J.-S. A new approach to prepare highly active and stable black titania for visible light-assisted hydrogen production. Energy Env. Sci. 2015, 8, 44. [Google Scholar] [CrossRef] [Green Version]
  166. Zou, X.; Liu, J.; Su, J.; Zuo, F.; Chen, J.; Feng, P. Facile synthesis of thermal- and photostable titania with paramagnetic oxygen vacancies for visible-light photocatalysis. Chem. Eur. J. 2013, 19, 73. [Google Scholar] [CrossRef]
  167. Zuo, F.; Wang, L.; Wu, T.; Zhang, Z.Y.; Borchardt, D.; Feng, P.Y. Self-doped Ti3+ enhanced photocatalyst for hydrogen production under visible light. Am. Chem. Soc. 2010, 132, 7. [Google Scholar] [CrossRef]
  168. Chen, X.; Liu, L.; Huang, F. Black titanium dioxide (TiO2) nanomaterials. Chem. Soc. Rev. 2015, 44, 85. [Google Scholar] [CrossRef] [PubMed]
  169. Li, H.; Chen, Z.; Tsang, C.K.; Li, Z.; Ran, X.; Lee, C.; Nie, B.; Zheng, L.; Hung, T.; Lu, J.; et al. Electrochemical doping of anatase TiO2 in organic electrolytes for high-performance supercapacitors and photocatalyst. Mater. Chem. 2014, 2, 36. [Google Scholar]
  170. Liu, X.; Gao, S.; Xu, H.; Lou, Z.; Wang, W.; Huang, B.; Dai, Y. Green synthetic approach for Ti3+ self-doped TiO2−x nanoparticles with efficient visible light photocatalytic activity. Nanoscale 2013, 5, 5. [Google Scholar] [CrossRef]
  171. Grabstanowicz, L.R.; Gao, S.; Li, T.; Rickard, R.M.; Rajh, T.; Liu, D.J.; Xu, T. Facile oxidative conversion of TiH2 to high-concentration Ti3+-self-doped rutile TiO2 with visible-light photoactivity. Inorg. Chem. 2013, 52, 90. [Google Scholar] [CrossRef]
  172. Liu, X.; Xu, H.; Grabstanowicz, L.R.; Gao, S.; Lou, Z.; Wang, W.; Huang, B.; Dai, Y.; Xu, T. Ti3+ self-doped TiO2−x anatase nanoparticles via oxidation of TiH2 in H2O2. Catal. Today 2014, 225, 9. [Google Scholar] [CrossRef]
  173. Li, X.; Yang, X.; Huang, Y.; Zhang, T.; Liu, B. Supported noble-metal single atoms for heterogeneous catalysis. Adv. Mater. 2019, 31, 1902031. [Google Scholar] [CrossRef] [PubMed]
  174. Huang, P.; Huang, J.; Pantovich, S.A.; Carl, A.D.; Fenton, T.G.; Caputo, C.A.; Grimm, R.L.; Frenkel, A.I.; Li, G. Selective CO2 reduction catalyzed by single cobalt sites on carbon nitride under visible-light irradiation. J. Am. Chem. Soc. 2018, 140, 16042–16047. [Google Scholar] [CrossRef] [PubMed]
  175. Zhang, P.; Zhan, X.; Xu, L.; Fu, X.; Zheng, T.; Yang, X.; Xu, Q.; Wang, D.; Qi, D.; Sun, T. Mass production of a single-atom cobalt photocatalyst for high-performance visible-light photocatalytic CO2 reduction. J. Mater. Chem. A 2021, 9, 26286–26297. [Google Scholar] [CrossRef]
  176. Xiong, X.; Mao, C.; Yang, Z.; Zhang, Q.; Waterhouse, G.I.; Gu, L.; Zhang, T. Photocatalytic CO2 reduction to CO over Ni single atoms supported on defect-rich zirconia. Adv. Energy Mater. 2020, 10, 2002928. [Google Scholar] [CrossRef]
  177. Dhakshinamoorthy, A.; Li, Z.; Garcia, H. Catalysis and photocatalysis by metal organic frameworks. Chem. Soc. Rev. 2018, 47, 8134–8172. [Google Scholar] [CrossRef]
  178. Chen, X.; Peng, X.; Jiang, L.; Yuan, X.; Yu, H.; Wang, H.; Zhang, J.; Xia, Q. Recent advances in titanium metal–organic frameworks and their derived materials: Features, fabrication, and photocatalytic applications. Chem. Eng. J. 2020, 395, 125080. [Google Scholar] [CrossRef]
  179. Yang, S.; Peng, L.; Bulut, S.; Queen, W.L. Recent Advances of MOFs and MOF-Derived Materials in Thermally Driven Organic Transformations. Chem. A Eur. J. 2019, 25, 2161–2178. [Google Scholar] [CrossRef]
  180. Zhu, J.; Li, P.-Z.; Guo, W.; Zhao, Y.; Zou, R. Titanium-based metal–organic frameworks for photocatalytic applications. Coord. Chem. Rev. 2018, 359, 80–101. [Google Scholar] [CrossRef]
  181. Khaletskaya, K.; Pougin, A.; Medishetty, R.; Rosler, C.; Wiktor, C.; Strunk, J.; Fischer, R.A. Fabrication of gold/titania photocatalyst for CO2 reduction based on pyrolytic conversion of the metal–organic framework NH2-MIL-125 (Ti) loaded with gold nanoparticles. Chem. Mater. 2015, 27, 7248–7257. [Google Scholar] [CrossRef]
  182. Bueken, B.; Vermoortele, F.; Vanpoucke, D.E.; Reinsch, H.; Tsou, C.C.; Valvekens, P.; De Baerdemaeker, T.; Ameloot, R.; Kirschhock, C.E.; Van Speybroeck, V. A Flexible Photoactive Titanium Metal–Organic Framework Based on a [TiIV3 (μ3-O)(O) 2 (COO) 6] Cluster. Angew. Chem. Int. Ed. 2015, 54, 13912–13917. [Google Scholar] [CrossRef] [PubMed]
  183. Ao, D.; Zhang, J.; Liu, H. Visible-light-driven photocatalytic degradation of pollutants over Cu-doped NH2-MIL-125 (Ti). J. Photochem. Photobiol. A Chem. 2018, 364, 524–533. [Google Scholar] [CrossRef]
  184. Rahmani, A.; Emrooz, H.B.M.; Abedi, S.; Morsali, A. Synthesis and characterization of CdS/MIL-125 (Ti) as a photocatalyst for water splitting. Mater. Sci. Semicond. Processing 2018, 80, 44–51. [Google Scholar] [CrossRef]
  185. An, X.; Jimmy, C.Y.; Wang, F.; Li, C.; Li, Y. One-pot synthesis of In2S3 nanosheets/graphene composites with enhanced visible-light photocatalytic activity. Appl. Catal. B Environ. 2013, 129, 80–88. [Google Scholar] [CrossRef]
Figure 1. Ball and stick models of anatase TiO2 low-index surfaces: (a) (101) surface; (b) (100) surface, and (c) (001) surface (red: oxygen; grey: titanium). (Taken with permission from [28]).
Figure 1. Ball and stick models of anatase TiO2 low-index surfaces: (a) (101) surface; (b) (100) surface, and (c) (001) surface (red: oxygen; grey: titanium). (Taken with permission from [28]).
Ijms 23 08143 g001
Figure 2. (a) Schematic illustrating the synthesis of TiO2-graphene nanocomposites with controllable TiO2 crystal facets. (bg) transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images of the as-prepared (b,e) TiO2-101-G, (c,f) TiO2-001-G, and (d,g) TiO2-100-G. (Taken with permission from [54]).
Figure 2. (a) Schematic illustrating the synthesis of TiO2-graphene nanocomposites with controllable TiO2 crystal facets. (bg) transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images of the as-prepared (b,e) TiO2-101-G, (c,f) TiO2-001-G, and (d,g) TiO2-100-G. (Taken with permission from [54]).
Ijms 23 08143 g002
Figure 3. Field emission scanning electron microscopy image of the synthesized TiO2 nanorod arrays. (Taken with permission from [79]).
Figure 3. Field emission scanning electron microscopy image of the synthesized TiO2 nanorod arrays. (Taken with permission from [79]).
Ijms 23 08143 g003
Figure 4. SEM image of 2D nanosheet sandwich-like graphene/TiO2. (a) The typical G2-TiO2 with remarkable structural flexibility. (b) The TiO2 nanoparticles intimate contact with graphene. (Taken with permission from [98]).
Figure 4. SEM image of 2D nanosheet sandwich-like graphene/TiO2. (a) The typical G2-TiO2 with remarkable structural flexibility. (b) The TiO2 nanoparticles intimate contact with graphene. (Taken with permission from [98]).
Ijms 23 08143 g004
Figure 5. Tauc plots of (a) TiO2 nanoparticles (b) Titanate nanotubes and rutile.
Figure 5. Tauc plots of (a) TiO2 nanoparticles (b) Titanate nanotubes and rutile.
Ijms 23 08143 g005
Table 1. Various techniques for the synthesis of hydrogenated TiO2 materials and their features in the synthesis processes.
Table 1. Various techniques for the synthesis of hydrogenated TiO2 materials and their features in the synthesis processes.
CatalystTreatmentNoteH2 Evolution Rate/Removal EfficacyReference
Black TiO2 nanoparticlesThermal plasma furnaceThe absorption increases promptly and monotonously in visible spectrum, when the wavelength is >400 nmVisible light: 83%[154]
Black TiO2 nanotubesHydrogen plasma methodNaOH solution (10M, 50 mL), to be used in heating 2g of P25 for 12 h, then being washed with water and HCL.7 µmol h−1 cm−2[155]
Black TiO2 nanoparticlesElectron beam treatmentElectron-beam-assisted high energy electron used in changing the composition of TiO2. Electron beam maximum energy 0.7 MeV. Electron beam maximum power 28 kWVisible light: 85%[156]
TiO2 nanotubes with black appearance for the proton-implanted layerProton implantationThe top of the nanotubes is being modified via high energy proton ion-implantation strategy. Then implanting the substrate with Varian 350D ion implanter. The resulted nanotubes showed high performance in aqueous solution.UV: 38%[157]
Defective TiO2Metal reductionMetals like Zn, Al, Mg are excellent reductants that for being cheap, safe and convenient in comparison with hydrogen.Solar light: 95%[142]
Black TiO2 and
TiO2 nanotubes
Aluminum reductionTiO2 and Al are being processed in a dual tube furnace below 0.5 Pa3.9 mmol g−1 h−1[158,159]
Gray TiO2 nanowiresAluminum reductionTitanate nanowires are being processed in double zone furnace in Al atmosphere for 4 hSolar light: 95%[160]
Black brookite TiO2 nanoparticlesAluminum reductionBrookite TiO2 and Al powder are being placed in dual vacuum furnace and heated for 4 h at 300–600 and 800 °C.
This process promoted the absorption of visible spectrum and IR of brookite TiO2
Solar light: 92%[159]
Black rutile TiO2 NanoparticlesMolten AluminumThe sample is being heated at 550–800 °C at a pressure of 6 × 10−4 Pa in a vacuum-double-zone furnace. And the results showed enhanced absorption.932 µmol h−1 g−1[161]
Black TiO2−x nanoparticlesAl powderAl powder and P25 (0.5 g) are being processed in a two-zone vacuum furnace. Then using thermal plasma furnace to apply hydrogen plasma for 5 h15 mmol h−1 g−1[162]
Black TiO2-N nanoparticles The material is being heated in a gas stream of NH3-ArSolar light: 85%[162]
Rutile TiO2 nanoparticlesZn reductionMixing aqueous TiCl2 (1 mL) and isopropanol (30 mL) at 180 °C in existence of Zn powder for 6 h.1.4 mmol h−1 g−1[163]
Black TiO2 photocatalystMg reductionMixing TiO2 with Mg powder resulted black TiO2. But Mg and H2 resulted in highly stable and active reduced black TiO2.440 µmol h−1 g−1[164]
Porous amorphous Vo-TiO2Organic reduction300-Xe lamp has been used as a light source. The target is aqueous methanol solution (25 vol%, 120 mL) for 8 h in UV and visible light: 5.67 mmol h−1 g−1
For 14 h, in visible light radiation: 115 µmol h−1 g−1
Visible light and UV: 5.67 mmol h−1 g−1[165]
Ti3+ doped TiO2Organic reduction300 W Xe lamp, aqueous methanol solution (25 vol%, 120 mL), for 4 h in visible light irradiation: 50 μmol h−1 g−1Visible light: 115 µmol h−1 g−1[166]
Defective TiO2−xOrganic reductionImidiazole and 2-ethylimidazole-[165,166,167]
Gray TiO2Organic reductionA TiO2 precursor exposed to UV for one hour then annealed with hydrochloric acid, and imidazole (1 g) in a muffle furnace at 450 °C115 µmol h−1 g−1[165]
Black defective TiO2 nanotubesElectrochemical reductionTiO2 were synthesized via Ti foil anodization in (4 mA for 5000 s or 80 V for 7200 s). then calcined in air.Visible light: 72%[168]
Ti3+ self-doped TiO2−x nanoparticlesChemical oxidationThe light source used is 300 W Xe lamp. Target and concentration are aqueous methanol solution (100 mL, 20%) MB (120 mL, 5 × 10−4 mol/L), for 4 h.250 µmol h−1 g−1[169]
Ti3+ self-doped rutile TiO2Chemical oxidationUsing solar simulator, MB (30 mL, 10−5 M), for 1 h-[170]
Ti3+ self-doped TiO2−x anatase nanoparticlesChemical oxidationLight source: 300 W Xe arc lamp
MB (100 mL, 1.5 × 10−5 mol/L) aqueous methanol solution (20 vol%) for 30 min
147 µmol h−1 g−1[171]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fawzi, T.; Rani, S.; Roy, S.C.; Lee, H. Photocatalytic Carbon Dioxide Conversion by Structurally and Materially Modified Titanium Dioxide Nanostructures. Int. J. Mol. Sci. 2022, 23, 8143. https://doi.org/10.3390/ijms23158143

AMA Style

Fawzi T, Rani S, Roy SC, Lee H. Photocatalytic Carbon Dioxide Conversion by Structurally and Materially Modified Titanium Dioxide Nanostructures. International Journal of Molecular Sciences. 2022; 23(15):8143. https://doi.org/10.3390/ijms23158143

Chicago/Turabian Style

Fawzi, Tarek, Sanju Rani, Somnath C. Roy, and Hyeonseok Lee. 2022. "Photocatalytic Carbon Dioxide Conversion by Structurally and Materially Modified Titanium Dioxide Nanostructures" International Journal of Molecular Sciences 23, no. 15: 8143. https://doi.org/10.3390/ijms23158143

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