Next Article in Journal
Three-Dimensional Evaluation of the Cytotoxicity and Antibacterial Properties of Alpha Lipoic Acid-Capped Silver Nanoparticle Constructs for Oral Applications
Previous Article in Journal
Ultra-Violet-Assisted Scalable Method to Fabricate Oxygen-Vacancy-Rich Titanium-Dioxide Semiconductor Film for Water Decontamination under Natural Sunlight Irradiation
Previous Article in Special Issue
Structural, Optical, and Arsenic Removal Properties of Sol–Gel Synthesized Fe-Doped TiO2 Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Heterophase Polymorph of TiO2 (Anatase, Rutile, Brookite, TiO2 (B)) for Efficient Photocatalyst: Fabrication and Activity

1
Department of Chemistry, Faculty of Mathematics and Natural Sciences, Universitas Padjadjaran, Sumedang 45363, West Java, Indonesia
2
Integrated Graduate School of Medicine, Engineering, and Agricultural Sciences, University of Yamanashi, Kofu 400-8511, Japan
3
Center for Crystal Science and Technology, University of Yamanashi, Kofu 400-8511, Japan
4
Department of Physics, Faculty of Mathematics and Natural Sciences, Universitas Padjadjaran, Sumedang 45363, West Java, Indonesia
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(4), 704; https://doi.org/10.3390/nano13040704
Submission received: 14 January 2023 / Revised: 7 February 2023 / Accepted: 8 February 2023 / Published: 12 February 2023
(This article belongs to the Special Issue Structure, Synthesis and Applications of TiO2 Based Nanomaterials)

Abstract

:
TiO2 exists naturally in three crystalline forms: Anatase, rutile, brookite, and TiO2 (B). These polymorphs exhibit different properties and consequently different photocatalytic performances. This paper aims to clarify the differences between titanium dioxide polymorphs, and the differences in homophase, biphase, and triphase properties in various photocatalytic applications. However, homophase TiO2 has various disadvantages such as high recombination rates and low adsorption capacity. Meanwhile, TiO2 heterophase can effectively stimulate electron transfer from one phase to another causing superior photocatalytic performance. Various studies have reported the biphase of polymorph TiO2 such as anatase/rutile, anatase/brookite, rutile/brookite, and anatase/TiO2 (B). In addition, this paper also presents the triphase of the TiO2 polymorph. This review is mainly focused on information regarding the heterophase of the TiO2 polymorph, fabrication of heterophase synthesis, and its application as a photocatalyst.

1. Introduction

Titanium dioxide, also known as titania, is a naturally occurring transition metal oxide of titanium with the chemical formula TiO2 [1]. Titanium dioxide is the most effective and efficient semiconductor material as a photocatalyst, has good stability, and has high ultraviolet absorption compared to other materials [2,3,4,5,6]. In addition, TiO2 is a commercially available materials for applications in various fields due to its wide availability, biocompatibility, and non-toxicity [7,8,9]. TiO2 has a white colour which is used as a pigment in paints, printing inks, plastics, ceramics, and cosmetics [10,11,12,13]. TiO2 is the most effective and efficient semiconductor material, and has also been reported in the photoelectrical devices [14,15]. TiO2 also has electronic and optical properties that can be utilized in the field of photocatalysts [16], self-cleaning materials [17], and solar cells [18].
Titanium dioxide has many applications as a photocatalyst, for example, in hydrogen production [19,20,21,22], degradation of organic compounds [23,24,25,26,27], remediation of metal ions [28,29,30], and synthesis of organic compounds [31,32,33,34]. Most of the research on TiO2 has been carried out to increase its photocatalytic efficiency [35]. Many studies have also been devoted to the synthesis of various forms of nanomaterials [36,37,38], engineered with doping [39,40,41,42] or composites [43,44,45].
TiO2 exists naturally in three crystalline forms, anatase, rutile, and brookite [46,47,48,49]. Besides that, there is another polymorph, namely TiO2 (B) [50,51]. These polymorphs exhibit different properties and, consequently, different photocatalytic performances [52]. In general, many studies have stated that anatase has the best photocatalytic activity [53]. However, in some cases, rutile is more active as a photocatalyst [54]. In addition, the binary mixture of TiO2 polymorphs showed a significant increase in the rate of catalytic activity for several reactions. Among these phases, the binary phase of anatase and rutile is the most studied phase [55,56].
The main drawback of photocatalyst application is the high recombination rate of electrons and holes which will reduce the quantum efficiency and decrease the photocatalytic activity [57,58]. One way to overcome this problem is to use multiphase TiO2 because it exhibits higher photocatalytic activity than single-phase due to the possible charge transfer generated between different TiO2 polymorphs (with different levels of electronic bands), leading to an effective separation of charge carriers thereby preventing recombination of electrons and holes [59,60].
The TiO2 heterophase can effectively stimulate electron transfer from one phase to another cause superior photocatalytic performance [61,62]. For example, TiO2 P25 Degussa, which consists of ∼20% rutile and ∼80% anatase, is an excellent photocatalyst [63]. Several studies have succeeded in synthesizing TiO2 heterophase which shows better photocatalytic abilities than pure anatase, namely, anatase/rutile [64] and anatase/brookite [65]. Growing research efforts have recently been devoted to new TiO2 heterophase composites including anatase/TiO2 (B) [66,67] and rutile/brookite [68], as well as the three-phase anatase/rutile/brookite [69] and anatase/rutile/TiO2 (B) [70]. Wang et al. [71] reported a novel approach to fabricate heterophase anatase/TiO2 (B) in which heterophase can be obtained at 550 °C. The results show that although in many cases the photocatalytic activity of TiO2 (B) is lower than that of anatase, a suitable composition between anatase and TiO2 (B) will lead to increased activity [72].
As described in detail by other researchers previously, the most obvious advantage of the multiphase photocatalyst is that it can promote electron–hole separation, thereby enhancing the photocatalytic activity of materials. Therefore, in the photocatalytic system, it is very important to synthesize a photocatalyst with a multiphase structure and high degradation efficiency. This review mainly addresses reports related to TiO2 heterophase over the past 10 years, in which a wider field of research has been reported. This work aims to clarify the differences between titanium dioxide polymorphs, and the differences in homophase and heterophase (biphase and triphase) properties in various photocatalytic applications. This review includes information on TiO2 polymorph heterophase, the fabrication of synthetic heterophase, and their applications as photocatalysts. This review ends with conclusions and perspectives, which may stimulate further development of the utility of TiO2 heterophase. Review articles covering heterophase with thorough explanations have not been reported. Other reviews have focused on anatase and rutile [73,74,75] with other composite or doped materials [76,77,78,79,80].

2. Homophase TiO2

2.1. Photocatalysis Mechanism of TiO2

TiO2 is a semiconductor material, which is a substance that lies between conductors (such as metals) and insulators (such as ceramics) [81]. In semiconductor, the distance between the position of the valence band (VB) and the conduction band (CB) determine the ability of the semiconductor material in the light absorption process and its oxidation-reduction ability [82,83,84,85]. In general, the photocatalytic reaction of TiO2 includes several basic processes, such as the formation of charge carriers, separation, relaxation, capture, transfer, recombination, and transport [86,87,88]. This process must be thoroughly observed to understand the workings of TiO2 photocatalysts and is useful for the development of new photocatalysts [89,90].
In general, the mechanism of photocatalysis using TiO2 is illustrated in Figure 1. When TiO2 material is subjected to photon irradiation with an energy greater than the band gap of TiO2, the electrons in VB will be excited to CB resulting in holes in VB [91,92,93,94,95]. The process of photoexcitation of pairs of electrons (e) and holes (h+) will participate in redox reactions with adsorbed species which will form superoxide radical anions (•O2−) and hydroxyl radicals (•OH), respectively, which will play a role in the degradation of organic pollutants in water [96,97,98,99]. Only photons with energies greater than the band-gap energy can excite electrons and drive reactions to occur. The activation of TiO2 by UV light can be written as Equations (1)–(3) [100,101,102]. In this reaction, h+ and e are strong oxidizing and reducing agents.
TiO2 + e + h+
e + O2 → O2
h+ + H2O → •OH + H+
Reactive oxygen species (ROS) produced during the photocatalytic process vary greatly, mainly the ROS that is formed is the superoxide radical O2 and the hydroxyl radical •OH. However, other species such as •OOH and H2O2 may be formed through further oxidation processes, dimerization, or disproportionation [103,104,105].

2.2. Phase of TiO2

TiO2 is widely available in nature as rutile, anatase, and brookite polymorphs; these three types are octahedral TiO6 with different distortions [106,107]. The structures of anatase, rutile, and brookite are shown in Figure 2 [108]. While rutile is a stable phase, both anatase and brookite are metastable phases. In addition, this brookite is difficult to synthesize, so it is rarely studied [109,110,111,112]. There is another polymorph found from TiO2, namely TiO2 (B) [113,114,115]. The characteristics of the Ti-O bond determine the differences in the structural and electronic properties of the different TiO2 phases [116]. Table 1 shows the different properties of the TiO2 polymorph [117].
In general, anatase showed higher photocatalytic activity than rutile. However, the reason for the difference in photocatalytic activity between these phases is still being debated [118,119]. Zhang et al. [120] showed that anatase is a semiconductor with an indirect band gap, while rutile and brookite are included in the category of direct band gap semiconductors. In addition, photocatalytic effects such as organic pollutant decomposition will be successful if the semiconductor material has a band gap energy of redox potential at the hole that is VB positive enough to produce hydroxyl radicals and electrons in CB, which must be negative enough to produce superoxide radicals (E0(H2O/•OH) = 2.8 V vs. NHE) and (E0(O2/O2•−) = −0.28 V vs. NHE). Figure 3 shows the band gap energies of anatase, rutile, and brookite TiO2 [121,122,123].

2.2.1. Anatase

Anatase polymorph has better photocatalytic activity than other TiO2 polymorphs [124,125,126,127,128]. Anatase has a larger band gap than rutile. This increases the oxidizing ability and facilitates electron transfer. Recent results show that the anatase has a lower conduction band than the rutile [129]. Moreover, the indirect band gap of the anatase form is smaller than the direct band gap. Thus, the indirect band gap is used for the photocatalysis process of the anatase form. For rutile, on the other hand, the fundamental band gap of indirect band gap is very similar to the direct band gap. Thus, the rutile type tends to excite electrons in a direct band gap. Indirect band gap materials generally exhibit a longer charge carrier lifetime compared to direct band gap materials [130,131,132,133], thus making anatase have better activity in most cases compared to rutile and brookite.
Anatase has a much wider specific surface area, so it has more active sites than rutile. Anatase also has a higher oxygen vacancy concentration than rutile, which is the reason for anatase’s higher charge separation efficiency [134]. The adsorption affinity of anatase for organic compounds is higher than that of rutile, and anatase shows a lower recombination rate than that of rutile because of its higher hole-trapping rate [135]. The adsorption affinity is defined as the equilibrium ratio of the solid-phase concentration to the liquid-phase solute concentration, also known as adsorption equilibrium constant [136]. However, even though it has better photocatalytic activity, the single anatase phase has low thermodynamic stability when compared to the rutile phase, so it can only be synthesized in several types of synthesis [137,138].
Bubacz et al. [139] synthesized anatase TiO2 with TiOSO4 precursor in ammonia water as a phenol degradation and azo dye. The decreased concentrations of the dye and phenol was measured by UV/VIS spectroscopy and a TOC analyzer. Polycrystalline of anatase was produced with a crystalline size of 12.7–13.0 nm and a particle size of 195.7 nm. The material produces satisfactory activity with an optimum pH of 6.5. In addition, Etacheri et al. [140] synthesized anatase with a band gap that can be reduced (without doping) and can works in visible light. This is achieved by generating oxygen in situ via the thermal decomposition of the peroxo–titania complex (H2O2-TiO2). The increased strength of the Ti–O–Ti bond and the maximum shift of VB are responsible for the stability in high temperature and visible light activity (Figure 4).
In the report of Lv et al. [141], anatase is modified using fluorine as a shape guide for surface modification. This modification produces various morphological forms such as nanotubes, titania sheets with {001} high-energy facets, and hollow spheres. This form results in increased diverse photocatalytic activity. In this regard, regarding the effect of fluoride on the structure and photocatalytic activity of TiO2, many questions remain open. In another report, Wang et al. [142] reported the differences in the photocatalytic activity ability between anatase and rutile mesopores. Mesoporous anatase results in long-lived electron generation and allows for efficient electron transfer. Meanwhile, mesoporous rutile is substantially less efficient and more reversible than anatase.

2.2.2. Rutile

The rutile phase becomes an attractive polymorph TiO2 for photocatalytic applications because it can absorb UV light near the visible region. The band gap of rutile is 3.0 eV, lower than that of anatase (3.2 eV). However, the rutile phase has not been widely studied as a photocatalyst because of its lower photocatalytic activity than anatase. Its low specific surface area, fast recombination, and the positive location of its minimum conduction band make it have a poorer reducing ability than anatase [143,144,145,146,147].
Rutile is the most stable TiO2 polymorph and is easily produced at high temperatures. However, rutile has rarely been studied as a photocatalyst for the oxidation of dyes/organic compounds, due to poor oxygen uptake due to the lower valence band position compared to anatase [148]. In the research by Wang et al. [149], rutile was successfully synthesized using low temperatures. Pure rutile was synthesized by hydrolysis of an aqueous TiCl4 ethanol solution at 50 °C. The results show that rutile synthesized at low temperatures exhibits much higher photocatalytic activity than commercial P25 photocatalysts in the degradation of rhodamine B. Applications of rutile photocatalysis include water separation. In contrast to anatase, rutile allows preferential water oxidation, which is useful for the construction of Z-scheme water separation systems [150]. In addition, in another study, rutile nanoparticles with large specific surface areas and abundant oxygen vacancies were designed for photocatalytic nitrogen fixation [151]. Meanwhile, in a study conducted by Yurdakal et al. [152], rutile was used for the selective oxidation of aromatic alcohols to aldehydes in an aqueous suspension.
The photocatalytic abilities of rutile and anatase were compared by Jung and Kim [153] to stearic acid and measured the rate of decrease in the integrated absorbance of the ensemble of the C−H stretching vibrations using FTIR spectroscopy. The results show that although rutile nanoparticles have a higher specific surface area than anatase nanoparticles, the photocatalytic properties of rutile nanoparticles are much lower than those of anatase nanoparticles. This is attributed to the intrinsic radiative recombination of photogenerated electrons and holes in rutile. However, other studies have shown that the presence of the vacancy-oxygen defect (VO) in rutile can significantly increase the ability of photocatalytic activity. VO is known to suppress the charge recombination process [154]. Although in many studies anatase shows higher activity, in some cases, rutile can be superior [155]. For example, the research by Zhang et al. [156], compared the photocatalytic activity of anatase, rutile, and brookite for Rhodamine B (RhB) degradation. The results show that rutile provides a faster photocatalytic reaction rate than anatase for the same particle size and specific surface area.

2.2.3. Brookite

Brookite has an orthorhombic crystal structure with the Pbca space group [157]. Brookite will turn into rutile at high temperatures. This conversion can occur directly to rutile or through the formation of anatase first. This depends on several factors such as crystallite size, size distribution, and crystallite contact area. [158]. For crystal sizes larger than 11 nm, brookite is more stable than anatase, while for sizes larger than 35 nm, rutile is the most stable phase [159]. Xie et al. [160] synthesized pure-phase brookite by hydrothermal method using Ti(SO4)2 as the precursor. Phase formation is achieved by hydrothermal treatment at 180 °C [161]. Kandiel et al. [162] observed that a direct transformation of anatase and brookite to rutile was observed, while in the case of anatase–brookite mixture, anatase changed first to brookite and then to rutile.
The photocatalytic activity of brookite was studied by Khan et al. and its activity is highly dependent on the level of disability. The results of the DFT calculations show that the Ti4+ defect is the main defect in increasing the photocatalytic activity of brookite [163,164]. One of the main reasons for the different catalytic performances of the TiO2 polymorph is the depth of charge carrier trapping. Electrons in brookite are trapped in shallow traps and not photogeneration, which reduces the amount of free electron, but on the other hand, this extends the life span of hole [165].
Bellardita et al. calculated the absolute crystallinity of brookite. This value was calculated from the ratio between the full width at half maximum (FWHM) intensity peak XRD pattern (121) of brookite and peak (111) of CaF2 as an internal standard. The results showed that crystallinity had a positive effect on the oxidation of 4-nitrophenol and 4-methoxybenzyl alcohol [166]. Choi et al. [167] synthesized pure brookite TiO2 films as a photocatalyst. The brookite phase was synthesized on titanium foil using a hydrothermal reaction. The results show that the brookite phase exhibits significantly higher photoactivity among the TiO2 polymorphs, despite its smaller specific surface area compared to anatase [168]. Tetramethylammonium (TMA) can be degraded in pure anatase and brookite phases, but not in rutile, where TMA concentration was measured by ion chromatograph [167].

2.2.4. TiO2 (B)

TiO2 (B) is a polymorph of titanium dioxide discovered in 1980 which was prepared from the hydrolysis of K2Ti4O9 followed by heating at 500 °C. TiO2 (B) has a host covalent framework of NaxTiO2 bronze [169]. In 2004, Armstrong et al. [115] successfully prepared TiO2 (B) nanowires. The synthesis was carried out involving a hydrothermal reaction between NaOH and TiO2, followed by acid washing and heating at 400 °C. The synthesized TiO2 (B) has the potential to be a material in lithium-ion batteries because it has the advantage of having a relatively open structure. Therefore, it is an excellent host for Li intercalation [170,171,172,173]. TiO2 (B) nanoparticles were also developed with various sizes. The small particle size causes the dimensionality reduction to increase the amount of Li and hence the charge that can be stored [174].
The use of TiO2 (B) as a photocatalyst was carried out by Chakraborty et al. [175]. The results show that H2Ti5O11.H2O is a prerequisite for the formation of TiO2 (B) phase. The photocatalytic activity of TiO2 (B) showed 1.35 and 1.95 times higher in decomposing 4-chlorophenol than Degussa P25 and 25 nm sized anatase nanoparticle. This is due to the high crystallinity of TiO2 (B). In addition, in other studies, TiO2 (B) doped with Ti3+ provided increased photocatalytic performance in rhodamine B (RhB) decomposition, which was measured using a UV spectrophotometer and hydrogen evolution, which was measured using gas chromatography. The RhB degradation rate is ∼6.9 and the hydrogen evolution rate is 26.6 times higher compared to titanium dioxide nanoparticles [176].

2.3. Disadvantages of TiO2 Homophase Photocatalysis

TiO2 has great potential as a photocatalyst for the degradation of organic pollutants and microorganisms. However, in practice, there are many limitations, namely, high h+ and e recombination rates, low adsorption capacity, and low thermal stability of the anatase TiO2 phase (Figure 5). TiO2 has a wide band gap of 3.0−3.2 eV, meaning this photocatalyst mainly absorbs ultraviolet photons, while indoor lamps only emit visible light photons [177,178]. This causes the use of indoor photocatalysts such as disinfection to be very limited because they require UVA irradiation. Strategies that can be implemented to overcome this challenge include adding metals (Cu, Ag, Eu, Fe, Zn, and La) [179,180,181,182], adding non-metallic dopants (N, F, S, and C) [183,184,185,186], or the use of the photon induction method [187].
The crystallite size and phase of TiO2 are aspects that affect the photonic efficiency and recombination dynamics in the photocatalytic process. Samples with a single anatase phase showed a high percentage of recombination due to the influence of crystallographic and microstructural parameters [188]. Meanwhile, samples with a mixed phase, such as anatase/rutile, have a more prominent crystallography and microstructure between anatase (011) and (110) rutile, which causes a decrease in the recombination rate due to the presence of hole and electron transfers [189]. Meanwhile, in the single rutile phase, there is a decrease in the formation rate of •OH and e due to the less efficient formation of •OH species during the oxidation of water on the rutile particles, as well as the shorter lifetime and lower reactivity of e in the single rutile phase. This phenomenon indicates that the photocatalytic synergy associated with anatase/rutile mixtures is not always present, but depends on the relationship between the fermi level of the anatase and rutile particles and the characteristics of the particles [190]. Photocatalytic junctions, especially type-II heterojunctions between semiconductors, are considered a potential pathway to improve photocatalytic performance [191]. The homojunction heterophase that exploits polymorphism has potential advantages. It has homogeneous components and nearly perfect lattice matching, and they can conduct efficient charge transfer at the interface [192].

3. Heterophase TiO2 in Photocatalysts

3.1. Photocatalysis Mechanism of TiO2 Heterophase

TiO2 with mixed phases such as anatase and rutile showed better photocatalytic activity than TiO2 with 100% anatase phase. This increase in activity generally comes from the interaction between the two forms; the interaction reduces recombination due to a bulk structure of anatase/rutile. This mixed phase has an interfacial electron trapping site that enhances the photocatalytic activity of TiO2 [193]. Meanwhile, in the anatase structure, the oxidation process is limited, while in contrast to the rutile structure, the reduction process is limited. Thus, when it is TiO2 with the mixed phase, the oxidation-reduction process can be accelerated [194]. This is illustrated in Figure 6. When different phases of TiO2 are mixed, the rate of recombination is reduced, the lighting efficiency is increased, and the energy band gap can be activated by light with energies lower than UV [135]. Due to its high activity, P25 Degussa, a commercial TiO2, is frequently used as a benchmark in photocatalytic processes. It has a mixed phase of anatase and rutile with a ratio of 80% and 20%, respectively [195].
At semiconductor heterojunctions, energy bands of two different materials come together, leading to an interaction. Both band structures are positioned discontinuously from each other, causing them to align close to the interface [196]. This alignment is caused by the discontinuous band structures of the semiconductors when compared to each other and the interaction of the two surfaces at the interface [197]. In research by Scanlon et al. [198], there are signs that when there is contact, the valence band energy and conduction band energy are higher in the rutile phase than in the anatase phase. Electrons will go to the anatase because of the lower conduction band minimum energy, and holes will move to the rutile because of the higher valence band maximum energy, according to the alignment of the bending energy bands at the anatase/rutile interface. Then, the holes in the valence band will react with water to generate hydroxyl radicals, while the electrons in the conduction band will concurrently react with oxygen to generate superoxide anions [199,200]. Figure 7 shows that the energy band changes at the interface of the anatase/rutile heterojunctions can lead to electron–hole separation and band bending until the Fermi levels of the anatase and rutile are aligned, and the density of the heterojunction and the toxicity of the nanoparticles can be determined by the degree of the band bending. The degree of band bending can be used to experimentally show that the heterojunction density of the TiO2 mixed phase is theoretically proportional to the quantity of electron–holes at the heterojunction interface. Heterojunction density is the local density of state and charge density for heterojunction semiconductor (usually composites), which is calculated using computational software such as the Vienna ab initio simulation package (VASP) and Cambridge Sequential Total Energy Package (CASTEP) [201]. Furthermore, the ROS generated during the photocatalytic process and its potential in the oxidation process can be estimated by the band bending value [202]. Band bending value or degree of band bending is proportional to the amount of accumulated electrons or holes that depend on the density of heterojunction; this value can reflect the density of heterojunction in mixed-phase TiO2 [203].

3.2. Computational Study of Heterophase TiO2

Surface structures and interfacial sections are very important for photocatalytic processes because these points are considered as the focus of transfer and trapping of charge carrier species [204,205]. The possible arrangement at the interfacial point between anatase and rutile was analyzed by stacking the atomic layers of the anatase (011) plane and the (110) rutile plane [189]. In a study by Denkins et al. [206], two plates were prepared, one anatase and one rutile, each 30 Å thick, separated from each other by 5 Å. Then, MD simulations were carried out for 400 ps at 1300 K. An illustration of the initial and final structures is shown in Figure 8. As shown, a vacuum distance of about 25 Å is present to minimize interactions between non-interfacial surfaces.
The band gap of the structure of the TiO2 mixture measured using HSE06 is around 3.0 eV which comes from the alignment band between the anatase (3.5 eV) and rutile (3.2 eV) phases [207]. The quantum size effect is responsible for the distinction in the band-gap values (approximately 0.2 eV) of each phase. The energy level of the Ti d orbitals can be adjusted by various coordination configurations, which makes this plausible. Additionally, a significant component in determining the photocatalytic activity linked to geometric and electrical structures is the absorption of photons from a semiconductor photocatalyst [208]. There is no significant change for the absorption curves of the mixed structure compared to the individual components of anatase and rutile. It shows that there is little orbital overlap and little electronic transition between rutile and anatase, and that direct electronic infusion of rutile into anatase hardly ever occurs during the excitation process. The structural changes that form the interface have no significant effect on the absorption of light by mixed-phase TiO2 and the optical absorption of this structure occurs in the anatase and rutile phases, respectively [205]. In another study, anatase/rutile (A/R) and interlayer amorphous (Am) films A/Am/R were synthesized using atmospheric pressure chemical vapor deposition (APCVD) to produce the absorption and bandgap shown in Figure 9. The gap energy of the band for the anatase/rutile heterojunction (2.78 eV) appears red-shifted from the pure component [209].

3.3. Fabrication Methods of Heterophase TiO2

3.3.1. Sol–Gel Method

In general, the sol–gel process involves a system transition from a liquid ‘sol’, which is mostly in the colloidal form, to a solid ‘gel’ phase. The starting materials used in the manufacture of ‘sol’ in the synthesis of TiO2 are usually metal alkoxide compounds such as titanium tetraisopropoxide (TTIP) and inorganic metal salts such as TiCl4 [210,211,212]. Arnal et al. [213] prepared titania by the sol–gel method from TiCl4 in diethyl ether, at 110 °C, to produce anatase. Meanwhile, for TiC14 with ethanol, it leads to rutile as early as 110 °C, whereas the reaction of tert-butyl alcohol at 110 °C leads to the formation of a single brookite. The TiO2 polymorph heterophase synthesized by Castrejón-Sánchez et al. [214] which can control the anatase/rutile ratio. Initially, anatase-amorphous TiO2 powder was synthesized by the sol–gel method. Then, annealing was carried out with a time ranging from 35 to 200 min at 475 °C. By adjusting the annealing duration, it is feasible to manage the anatase/rutile ratio in nanostructured TiO2 powders from pure anatase to rutile. Figure 10 shows the sol–gel synthesis process with Ti(Obu)4 precursor using isobutyl alcohol and the addition of HNO3 to produce anatase/rutile heterophase.

3.3.2. Hydrothermal Method

The hydrothermal process has strengths over other procedures, including excellent purity and crystallinity in the synthesized materials. Additionally, this technique yields a homogenous particle size distribution with a performance of more than 90% [215,216,217,218,219]. In the hydrothermal method, the trigger for the appearance of the rutile phase is pH [220], using the precursor [221], and using certain solvents [222]. The addition of additives can also trigger the appearance of the rutile phase in TiO2, such as adding urea as a nitrogen source in g-C3N4/TiO2 [222], tartaric acid (C4H6O6) [223], or adding metal impurities [224,225,226].
In a study conducted by Wang et al. [227], mixed-phase TiO2 was prepared by the hydrothermal method from TiOSO4 and peroxide titanic acid (PTA). The findings demonstrated that anatase and rutile phases made up the mixed-phase TiO2 powder, and that the PTA sol was crucial to the development of the rutile core. By adjusting the quantity of PTA sol utilized in the synthesis process, it is simple to control the amount of rutile in the TiO2 mixed phase between 0% and 70.5%. Figure 11 shows the hydrothermal method with addition of PTA to produce anatase/rutile heterophase. Meanwhile, in a study by Yang et al. [228], rutile, anatase, and brookite TiO2 nanorods were obtained by the hydrothermal method using PTA with different pH values. Rutile forms at pH values below 10, but anatase TiO2 can form at pH values over 10. Once the pH is 10, brookite can form.

3.3.3. Sonochemical Method

Sonochemistry has been widely used to synthesize nano-sized materials [229,230,231,232,233]. Arami et al. [234] synthesized TiO2 nanoparticles with an average diameter of about 20 nm via a simple sonochemical method. The results indicated the nanoparticles consisted of a pure rutile phase. Meanwhile, for the mixed phase, anatase/rutile was prepared by a sonochemical process in which titanium(IV) isopropoxide (TTIP) was used as a precursor. The transformation of the anatase phase to the rutile phase in TiO2 powder was initially obtained by calcining the sample at 600 °C. The complete rutile phase occurs at the calcination temperature of 800 °C [235]. The best photocatalytic activity was demonstrated by an anatase/rutile composite with the proper ratio in a TiO2 photocatalyst that was calcined at 700 °C as a result of improved charge transfer at the mixed phase junction and reduced electron–hole pair recombination [236]. In addition, anatase/brookite composites have also been synthesized at 50 °C using sonochemical methods with similar crystal size and specific surface area, the photocatalytic activity of anatase/brookite for CH3CHO degradation is 5.4 times higher than in pure anatase. According to the electron energy loss spectrum, the collision of brookite and anatase crystals caused this high activity [237]. Figure 12 shows the sonochemical method with the TTIP precursor using ethanol and water as solvent to produce the anatase/rutile heterophase.

3.4. Transformation of Polymorph TiO2

The initial crystalline phase of TiO2 that forms during the various processes used to synthesize it is typically anatase. This may be owing to the usual TiO6 octahedral’s easier organization into a typical anatase structure when compared to rutile from a structural perspective [238]. From a thermodynamic perspective, the faster recrystallization of anatase can be due to the lower surface free energy compared to rutile [137]. However, some studies have been able to form rutile at room temperature conditions. The hydrothermal synthesis approach makes it possible to control the precipitation of rutile while facilitating the direct deposition of TiO2 crystals from the liquid phase. Rutile can only be obtained using this technique and high-temperature processing otherwise [239]. As the calcination changes in temperature, the proportions of rutile and anatase will progressively change as well. With higher calcination temperatures, the proportion of the rutile phase increases, which causes the proportion of the anatase phase to diminish [240]. Without any precursor or dopant modification, when making TiO2, the anatase to rutile conversion typically takes place between 600 and 700 °C [241]. Yuangpho et al. [242] examined the temperature of the anatase–rutile transformation and its impact on the material’s physical/chemical characteristics and photocatalytic activity. The photocatalytic activity of TiO2 decreases with increasing annealing temperature. These findings suggest that the phase composition influences the photocatalytic activity of TiO2 particles.

4. Photocatalytic Activity of Heterophase TiO2

4.1. Biphase of TiO2

4.1.1. Anatase/Rutile

Anatase and rutile are the most common phase of TiO2 [243]. As a single phase, anatase has better photocatalytic activity than rutile [244]. However, the mixed anatase/rutile phases showed higher photocatalytic activity than single anatase [245]. Rutile is the most stable form compared to the other two forms that are metastable, brookite, and, anatase so at high temperatures it will turn into rutile [246]. It is common to regulate the calcination temperature to produce TiO2 with a mixture of anatase and rutile phases [247,248,249,250,251]. However, there are also other factors that can be used to control the appearance of these two phases such as the pH and the solvent used [252,253]. Several reports on the formation of mixed phase anatase/rutile structures are summarized in Table 2.
TiO2 with anatase/rutile can be used for the production of hydrogen, water splitting, and the degradation of organic pollutants. The factors that influence the increase in anatase/rutile photocatalyst activity are numerous energy-level staggered interfaces between the anatase and rutile, large specific surface area, and an enhancement in bridged hydroxyl functional groups on the surface [253].
Photocatalysis activity and degradation pathways from TiO2 with mixed phases are dependent on organic substances targeted by photocatalysis [254]. In many cases, increased anatase content tends to result in better TiO2 photocatalysis activity [255]. When combined with rutile, which is large in size and has high crystallinity, it will be an excellent photocatalytic. This is because degradation is a reaction that requires oxygen and anatase is a good oxygen absorber when compared to rutile. In addition, rutiles that have high crystallinity can also increase intrinsic photocatalysis activity [256]. The synergistic interaction between the anatase (011) and (110) rutile fields, which can also enhance the separation of electron holes and suppress electron and hole recombination, resulting in good photocatalysis activity [257]. Table 3 shows the performance of anatase/rutile photocatalytic degradation for the organic pollutants. Photocatalytic activity against the degradation of organic compounds is carried out by adding photocatalysts to the solution of organic compounds; then, concentration reduction was observed with a UV–vis spectrophotometer [247], or by measuring the conductivity of CO2 resulting from degradation [248]. For hydrogen generation, the hydrogen formation is measured with gas chromatography (GC) [249].
According to Ding et al. [63] the photocatalytic activity for the production of hydrogen (H2) and oxygen (O2) from pure water was significantly increased by the anatase/rutile synthesized using the hydrothermal/calcination method. It proves that rutile/anatase performed significantly better than pure rutile and pure anatase. In another study, anatase/rutile nanoparticles were modified with oxygen vacancies (TiO2−x). This heterophase TiO2−x nanoparticles exhibit superior photocatalytic activity for converting CO2 to methane and can accelerate electron–hole separation [260].

4.1.2. Anatase/Brookite

Among the three main crystallographic forms of TiO2, metastable brookite has received the least amount of research due to the difficulties with its synthesis in a pure form [261]. However, the development of several successful methods such as hydrothermal and sol–gel methods, pure brookite now exists. For photocatalytic activity, pure anatase exhibits a higher activity compared to other phases because the indirect band gap can exhibit lower recombination rates due to the longer lifetimes for photo-excited electrons and holes [70]. However, phase mixtures of various polymorphs are known to have synergistic effects and exhibit improved photocatalytic activity [56]. Compared to anatase/rutile, mixed-phase titania containing brookite has received less attention [262]. Table 4 summarizes a few studies on the preparation of mixed-phase TiO2 nanostructures with a tuned ratio of brookite in the titania mixtures. Meanwhile, Table 5 shows the performance of anatase/brookite photocatalytic degradation for the organic pollutants.
Anatase/brookite is successfully produced by hydrothermal method using titanium bis(ammonium lactate) dihydroxide (TALH) in the presence of high concentrations of urea (≥6.0 M) as an in situ OH− source. Lower urea concentrations lead to the formation of anatase/brookite mixtures. According to the results, pure anatase is less effective for hydrogen evolution than mixtures of anatase/brookite or pure brookite. On the other hand, pure brookite and mixtures of anatase and brookite show lower photocatalytic activity compared to pure anatase for the photocatalytic degradation of dichloroacetic acid (DCA) that performed at pH 3 [263]. However, TALH is neither affordable nor environmentally friendly (causing a release of ammonia) [262]. Another research work used titanium trichloride (TiCl3) as the titanium source, which can be easily manipulated, and uses tartaric acid (C4H6O6), which is low-cost and safe [265]. The as-prepared TiO2 that consisted of 78.7% anatase and 21.3% brookite showed the highest photocatalytic activity for Rhodamine B degradation, which was 1.2 times greater than commercial P25. A UV–Vis spectrometer was used to measure the maximum absorbance in order to determine the concentration of Rhodamine B. Meanwhile, anatase/brookite synthesized by an alkalescent hydrothermal treatment of TiCl3 by adjusting the NaCl concentration and NH4OH/H2O volume ratio showed that the product containing 53.4% brookite and 46.6% anatase shows the highest photocatalytic activity for degradation Rhodamine B, which is higher than pure phase [266]. For synthesized via a simple hydrothermal method with titanium sulfide (TiS2) as the precursors in sodium hydroxide solutions showed anatase/brookite TiO2 is 2.2 times more active—when measured by the H2 yield per unit area of the photocatalyst surface—than commercial P25 [267]. Another research work using titanium(III) sulfate (Ti2(SO4)3) in the presence of glycine through hydrothermal treatment shows that a sample containing 38.2% brookite and 61.8% anatase exhibited the highest efficiency for the removal cylindrospermopsin (CYN) and diclofenac than pure anatase, pure brookite, and commercial P25. The concentration of CYN was analyzed by high-performance liquid chromatography with a photodiode-array detector (PDA). Meanwhile, diclofenac concentration is measured by the UV–vis spectrophotometer [262,267].
TiO2 nanoparticles with anatase/brookite were also synthesized using a modified sol–gel method at low temperatures [268]. The mixture obtained at pH 2 and calcined at 200 °C had the highest activity for methylene blue degradation that measured by UV-Vis spectroscopy at 660 nm compared to commercial P25. The modified conventional sol–gel method is using the microwave-assisted sol–gel technique [269]. The phase transformation of TiO2 was investigated by hydrochloric and acetic acid. Three titania polymorphs were found when hydrochloric acid was used as a catalyst. On the other hand, a single anatase phase was obtained when acetic acid was added after only 15 min of reaction time. The ultrasonic-assisted sol–gel method [270], which used weak and strong acids, demonstrated a significant effect on their morphology, size, crystallinity, and photocatalytic performance. The photocatalytic activity showed that anatase and rutile phases with a high proportion of anatase (69:31 and 93:7, respectively), had the highest photoactivity for the degradation of MB compared with anatase/brookite (70:30).
According to Cihlar et al. [271], anatase/brookite is produced using the sol–gel complex method. This method consists of sol–gel, hydrothermal, and solid-state reactions that used titanium tetraisopropoxide (TTIP) as the precursor and the complex-forming (chelate-forming) substances used were glycine, EDTA, acetylacetone, and hydroxycarboxylic acids (lactic acid (LA), citric acid (CA), and tartaric acid (TA)) and their salts. These materials use to produce hydrogen by photocatalysis. The results show that anatase/brookite nanoparticles with 36% brookite, which were made using lactic acid, catalyzed the maximum rate of H2 evolution. The presence of anatase particles made using acetylacetone was found to have the lowest rate of H2 evolution. Another research work investigated the phase composition of TiO2 using lactic acid/Ti molar ratio ranging from 0.02 to 3.0 ratio via sol–gel complex synthesis [272]. Anatase/brookite was formed at low LA/Ti molar ratios, followed by anatase/brookite/rutile particles at average molar ratios, and pure anatase at high molar ratios. Anatase/brookite nanoparticles made at LA/Ti molar ratios between 0.03 and 0.1 showed the highest photocatalytic activity in the production of hydrogen by water splitting. Table 6 shows the performance of heterophase TiO2 for hydrogen production.

4.1.3. Rutile/Brookite

Rutile/brookite as a photocatalyst was reported in 2008 by Di Paola et al. [273]. The sample was synthesized by thermohydrolysis of TiCl4 in HCl or NaCl aqueous solutions. Rutile mixtures were obtained depending on the acidity of the medium. The photocatalytic activity was evaluated by 4-nitrophenol degradation and quantitative determination was performed by measuring its absorption at 315 nm with a spectrophotometer UV. The highest photocatalytic activity for 4-nitrophenol degradation corresponded to the powders consisting of heterophase, compared to the pure phase [274].
Another way to synthesize a controlled rutile/brookite is by the hydrothermal method. This method uses titanium tetrachloride as a titanium source and triethylamine as a “regulating reagent” to adjust the ratio of brookite to rutile. The research found that the TiO2 sample with the highest photocatalytic activity for RhB degradation was obtained in a solution of 3 mL triethylamine and 10 mL water and contained 38% brookite and 62% rutile [275].
In another study, brookite/rutile was obtained using the solvothermal method. The morphology and structure of the samples were greatly changed by varying the ratio of TiCl4 to t-BuOH precursors. The phenol degradation measured by HPLC system showed that the highest photocatalytic activity was obtained from a composition of 72% brookite and 28% rutile. The highest phenol degradation activity was obtained from a composition of 72% brookite and 28% rutile. This indicates that the biphase of brookite/rutile with optimized phase proportions is responsible for the efficient synergy effect [276,277].

4.1.4. Biphase of TiO2 (B)

TiO2 (B) is a less stable phase of TiO2 than anatase and rutile [114]. When compared to anatase, rutile, and brookite, TiO2 (B) is the least dense polymorph of TiO2 [278]. TiO2 (B) is suitable for use as a place of Li intercalation because it has a structure that is relatively open, with free space and continuous channels [115]. The synthetic TiO2 (B) nanowire-based electrode exhibited unique electronic properties, e.g., favourable charge-transfer ability, negative-shifted appeared flat-band potential, the existence of abundant surface states or oxygen vacancies, and high-level donor density [279,280]. There is an alternative synthesis of TiO2 (B), through the formation of a titanium complex obtained from the reaction between titanium metal powder and H2O2, NH3, and glycolic acid which produces a yellowish gel. The resulting gel is then added to H2SO4 to change the pH and put into the autoclave for hydrothermal treatment [173]. According to research by Wang et al. [73], the right quantity of HF could prevent the phase transition of TiO2 (B) to anatase. The mechanism of this phase transformation occurs because the F anion adsorbed on the surface of TiO2 (B) can efficiently reduce the surface energy from 0.63 J/m2 for a clean surface to −0.22 J/m2 for the surface adsorbed on F anion. Several methods for synthesizing TiO2 (B) are shown in Table 7.
However, TiO2 (B), as a photocatalyst composited with another TiO2 polymorph, was performed by Li et al. [67]. The biphase of TiO2 was synthesized from K2Ti2O5 through ion exchange and calcination. The nanofiber of core–shell crystal structure had a thin TiO2 (B) shell and an anatase core. The core–shell anatase/TiO2 (B) nanofiber showed increased photocatalytic activity in iodine oxidation reactions, with a reaction increase of 20–50% compared to single crystal anatase nanofiber or single crystal TiO2 (B). The same m was also conducted by Yang et al. [61], by synthesizing core–shell TiO2 (B)/anatase. A number of characteristics of this unusual structure improve the photocatalytic activity against the degradation of sulforhodamine B when exposed to UV light.
The advantage of having a biphase of TiO2 (B)/anatase is that TiO2 (B) and anatase will form a heterojunction; the photogenerated electrons and holes can be transferred to the anatase phase and TiO2 (B) phase, respectively (Figure 13), thereby reducing recombination [195,284]. In the study of Mikrut et al. [285], TiO2 (B)/anatase was synthesized in the form of nanobelts. Similar to the anatase/rutile composite, a synergistic effect of the presence of the two phases was observed for TiO2 (B)/anatase (2:98). It is well known that TiO2 (B) is regarded as an optimized photocatalyst component for oxidation reactions involving hydroxyl radicals rather than promoting hydrogen evolution reactions.
Zhu et al. [286] reported another TiO2 (B) heterophase. In their research, the disintegration of TiO2 (B)/rutile used hydrothermal and calcination methods. The two phases are connected by an angle division at the phase interface (Figure 14). Electrons and holes are effectively separated across the phase interface as a result of the different conduction band and valence band positions between the two phases. The best results as photocatalysis were obtained at a TiO2 (B)/rutile ratio of 2/1, indicating the highest photocurrent and the best H2 evolution performance.

4.2. Triphase of TiO2

Numerous studies have been conducted to improve the activity of TiO2, but the two most significant findings are increasing the specific surface area and using the mixed phase [287]. The triphase of polymorph TiO2 shows a significant increase in photocatalytic activity compared to the pure and biphase phases [71]. Allen et al. [288] reported the effect of thermal treatment of heterophase polymorph TiO2 prepared by hydrolyzing titanium tetraisopropoxide at room temperature, dried at 382 K, and calcined at different temperatures for 1 h up to 1172 K. The results demonstrate that a mixture of brookite and anatase phases were seen up to 772 K, while a mixture of all three phases (anatase, brookite, and rutile) was present at 872 K, and a rutile-only phase was present at 1097 K and above.
Fischer et al. [70], synthesized a mixture of TiO2 phases (anatase, rutile, and brookite) on a polyethersulfone (PES) membrane via low-temperature dissolution-precipitation. In order to achieve that, the concentration of titanium precursor (titanium(IV) isopropoxide) was held constant while the amounts of hydrochloric acid and reaction temperature were adjusted to range from 0.1 to 1 M and 25 to 130 °C, respectively. The result showed that the best degradation rate of methylene blue were measured via microplate reader at a wavelength of λ = 660 nm, showing the highest activity that was obtained by 79% anatase and 21% brookite. Meanwhile, the recovery test showed that anatase, brookite, and rutile samples (70, 26, and 4%, respectively) did not degrade and completely recovered their photocatalytic abilities after at least nine additional cycles.
In a different study by Kaplan et al. [289], the formation of TiO2 that consists of anatase (43%), rutile (24%), and brookite (33%) was obtained by synthesizing TiO2 mixed phase using the sol–gel method and then hydrothermally treating it at a mild temperature (175 °C, 24 h). It was discovered that the photocatalytic activity of TiO2 nanocomposite successfully converted nearly 60% of the pollutant bisphenol into CO2 in H2O after being irradiated by UV light for 60 min. In contrast, only a lower level of mineralization was attained by the TiO2 P25 Degussa benchmark catalyst. This is due to a significantly reduced resistance to the accumulation of carbonaceous deposits on the catalyst surface.

5. Conclusions

This review summarizes most of the papers on TiO2 polymorphs (anatase, rutile, brookite, and TiO2 (B)), pure-phase TiO2 photocatalytic mechanisms, mixed phase, synthesis methods, and photocatalytic activities. This paper explains how to obtain pure phases from various TiO2 polymorphs and examples of their application in photocatalysis. In many cases, anatase has better activity than rutile and brookite. However, several studies have shown that rutile and brookite may have better activity in certain cases. Even thus, one-phase TiO2 has some drawbacks, such as high recombination and low specific surface area. TiO2 mixed phase shows better activity than the pure phase. This is due to good charge separation due to the formation of junctions around the interface. In addition, the synthesis methods that are often used to prepare mixed-phase TiO2 are sol–gel, hydrothermal, and sonochemical methods. Although the phase controller is usually the annealing temperature for anatase/rutile, in the hydrothermal method, additives are usually used as phase controllers.
Photocatalytic activity shows that anatase/rutile has much better activity than pure anatase or rutile. In addition, development continues to see the activity of other heterophases, such as anatase/brookite, rutile/brookite, and the heterophase TiO2 (B). However, in the case of brookite and TiO2 (B), the synthesis tends to be more difficult, these heterophase results show good photocatalytic activity. In addition, the triphase of polymorph TiO2 showed good photocatalytic activities for organic pollutants degradation. However, further studies to determine the exact number of compositions for each polymorph for the best activity in triphase TiO2 have not yet been investigated. In conclusion, further investigations are recommended to explore an easy method to synthesize TiO2 polymorphs with a high specific surface area and investigate the best composition of each TiO2 polymorph for photocatalytic activity, especially for triphase.

Author Contributions

D.R.E., S., S.H., T.T., N.K. and I.R. conceptualized the article and supervised the overall work; M.D.P., L.K.S. and G.A.N.S. wrote the original article; M.D.P. and L.K.S. visualized the figures and illustrations. All authors contributed to the writing and revising of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The Riset Data Pustaka dan Daring (RDPD) (ID: 2203/UN6.3.1/PT.00/2022) and Academic Leadership Grant (ALG) Prof. Iman Rahayu, Universitas Padjadjaran (ID: 2203/UN6.3.1/PT.00/2022) for supporting financially.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The study did not report any data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Haider, A.J.; Jameel, Z.N.; Al-Hussaini, I.H. Review on: Titanium dioxide applications. Energy Procedia 2019, 157, 17–29. [Google Scholar] [CrossRef]
  2. Eddy, D.R.; Ishmah, S.N.; Permana, M.D.; Firdaus, M.L. Synthesis of titanium dioxide/silicon dioxide from beach sand as photocatalyst for Cr and Pb remediation. Catalysts 2020, 10, 1248. [Google Scholar] [CrossRef]
  3. Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef]
  4. Glassford, K.M.; Chelikowsky, J.R. Structural and electronic properties of titanium dioxide. Phys. Rev. B 1992, 46, 1284–1298. [Google Scholar] [CrossRef]
  5. Grant, F.A. Properties of rutile (titanium dioxide). Rev. Mod. Phys. 1959, 31, 646–674. [Google Scholar] [CrossRef]
  6. Breckenridge, R.G.; Hosler, W.R. Electrical properties of titanium dioxide semiconductors. Phys. Rev. 1953, 91, 793–802. [Google Scholar] [CrossRef]
  7. Chen, D.; Cheng, Y.; Zhou, N.; Chen, P.; Wang, Y.; Li, K.; Huo, S.; Cheng, P.; Peng, P.; Zhang, R.; et al. Photocatalytic degradation of organic pollutants using TiO2-based photocatalysts: A review. J. Clean. Prod. 2020, 268, 121725. [Google Scholar] [CrossRef]
  8. Barbosa, J.S.; Neto, D.M.A.; Freire, R.M.; Rocha, J.S.; Fechine, L.M.U.D.; Denardin, J.C.; Valentini, A.; de Araújo, T.G.; Mazzetto, S.E.; Fechine, P.B.A. Ultrafast sonochemistry-based approach to coat TiO2 commercial particles for sunscreen formulation. Ultrason. Sonochem. 2018, 48, 340–348. [Google Scholar] [CrossRef]
  9. Jafari, S.; Mahyad, B.; Hashemzadeh, H.; Janfaza, S.; Gholikhani, T.; Tayebi, L. Biomedical Applications of TiO2 Nanostructures: Recent Advances. Int. J. Nanomed. 2020, 15, 3447–3470. [Google Scholar] [CrossRef]
  10. Gázquez, M.J.; Bolívar, J.P.; García-Tenorio, R.; Vaca, F. A review of the production cycle of titanium dioxide pigment. Mater. Sci. Appl. 2014, 2, 441–458. [Google Scholar] [CrossRef]
  11. Day, R.E. The role of titanium dioxide pigments in the degradation and tabilization of polymers in the plastics industry. Polym. Degrad. Stab. 1990, 29, 73–92. [Google Scholar] [CrossRef]
  12. Middlemas, S.; Fang, Z.Z.; Fan, P. A new method for production of titanium dioxide pigment. Hydrometallurgy 2013, 131, 107–113. [Google Scholar] [CrossRef]
  13. Gesenhues, U. Calcination of metatitanic acid to titanium dioxide white pigments. Chem. Eng. Technol. 2001, 24, 685–694. [Google Scholar] [CrossRef]
  14. Yang, B.; Wang, M.; Hu, X.; Zhou, T.; Zang, Z. Highly efficient semitransparent CsPbIBr2 perovskite solar cells via low-temperature processed In2S3 as electron-transport-layer. Nano Energy 2019, 57, 718–727. [Google Scholar] [CrossRef]
  15. Tan, X.; Liu, X.; Liu, Z.; Sun, B.; Li, J.; Xi, S.; Shi, T.; Tang, Z.; Liao, G. Enhancing the optical, morphological and electronic properties of the solution-processed CsPbIBr2 films by Li doping for efficient carbon-based perovskite solar cells. Appl. Surf. Sci. 2020, 499, 143990. [Google Scholar] [CrossRef]
  16. Islam Molla, M.A.; Tateishi, I.; Furukawa, M.; Katsumata, H.; Suzuki, T.; Kaneco, S. Evaluation of Reaction Mechanism for Photocatalytic Degradation of Dye with Self-Sensitized TiO2 under Visible Light Irradiation. Open J. Inorg. Non met. Mater. 2017, 7, 1–7. [Google Scholar]
  17. Rabajczyk, A.; Zielecka, M.; Klapsa, W.; Dziechciarz, A. Self-cleaning coatings and surfaces of modern building materials for the removal of some air pollutants. Materials 2021, 14, 2161. [Google Scholar] [CrossRef]
  18. Roy, P.; Kim, D.; Lee, K.; Schmuki, P. TiO2 nanotubes and their application in dye-sensitized solar cells. Nanoscale 2010, 2, 45–59. [Google Scholar] [CrossRef]
  19. Preethi, V.; Kanmani, S. Photocatalytic hydrogen production. Mater. Sci. Semicond. Process. 2013, 16, 561–575. [Google Scholar] [CrossRef]
  20. Teets, T.S.; Nocera, D.G. Photocatalytic hydrogen production. ChemComm. 2011, 47, 9268–9274. [Google Scholar] [CrossRef]
  21. Zhu, J.; Zäch, M. Nanostructured materials for photocatalytic hydrogen production. Curr. Opin. Colloid Interface Sci. 2009, 4, 260–269. [Google Scholar] [CrossRef]
  22. Du, H.; Liu, Y.N.; Shen, C.C.; Xu, A.W. Nanoheterostructured photocatalysts for improving photocatalytic hydrogen production. Chinese J. Catal. 2017, 38, 1295–1306. [Google Scholar] [CrossRef]
  23. Chatterjee, D.; Dasgupta, S. Visible light induced photocatalytic degradation of organic pollutants. J. Photochem. Photobiol. C Photochem. Rev. 2005, 6, 186–205. [Google Scholar] [CrossRef]
  24. Shayegan, Z.; Lee, C.S.; Haghighat, F. TiO2 photocatalyst for removal of volatile organic compounds in gas phase–A review. Chem. Eng. J. 2018, 334, 2408–2439. [Google Scholar] [CrossRef]
  25. Umar, M.; Aziz, H.A. Photocatalytic degradation of organic pollutants in water. In Organic Pollutants-Monitoring, Risk, and Treatment; Rashed, M.N., Ed.; IntechOpen: London, UK, 2013; Volume 8, pp. 196–197. [Google Scholar]
  26. Dong, H.; Zeng, G.; Tang, L.; Fan, C.; Zhang, C.; He, X.; He, Y. An overview on limitations of TiO2-based particles for photocatalytic degradation of organic pollutants and the corresponding countermeasures. Water Res. 2015, 79, 128–146. [Google Scholar] [CrossRef]
  27. Paumo, H.K.; Dalhatou, S.; Katata-Seru, L.M.; Kamdem, B.P.; Tijani, J.O.; Vishwanathan, V.; Kane, A.; Bahadur, I. TiO2 assisted photocatalysts for degradation of emerging organic pollutants in water and wastewater. J. Mol. Liq. 2021, 331, 115458. [Google Scholar] [CrossRef]
  28. Ameta, R.; Benjamin, S.; Ameta, A.; Ameta, S.C. Photocatalytic degradation of organic pollutants: A review. Mater. Sci. Forum 2013, 734, 247–272. [Google Scholar] [CrossRef]
  29. Wang, C.C.; Du, X.D.; Li, J.; Guo, X.X.; Wang, P.; Zhang, J. Photocatalytic Cr (VI) reduction in metal-organic frameworks: A mini-review. Appl. Catal. B 2016, 193, 198–216. [Google Scholar] [CrossRef]
  30. Kumar, A.; Choudhary, P.; Kumar, A.; Camargo, P.H.; Krishnan, V. Recent advances in plasmonic photocatalysis based on TiO2 and noble metal nanoparticles for energy conversion, environmental remediation, and organic synthesis. Small 2022, 18, 2101638. [Google Scholar] [CrossRef]
  31. Lin, Z.; Zheng, Y.; Deng, F.; Luo, X.; Zou, J.; Shao, P.; Zhang, S.; Tang, H. Target-directed design of dual-functional Z-scheme AgIn5S8/SnS2 heterojunction for Pb (II) capture and photocatalytic reduction of Cr (VI): Performance and mechanism insight. Sep. Purif. Technol. 2021, 277, 119430. [Google Scholar] [CrossRef]
  32. Hubbard, J.S.; Hardy, J.P.; Voecks, G.E.; Golub, E.E. Photocatalytic synthesis of organic compounds from CO and water: Involvement of surfaces in the formation and stabilization of products. J. Mol. Evol. 1973, 2, 149–166. [Google Scholar] [CrossRef]
  33. Lima, M.J.; Silva, A.M.; Silva, C.G.; Faria, J.L.; Reis, N.M. Selective photocatalytic synthesis of benzaldehyde in microcapillaries with immobilized carbon nitride. Chem. Eng. J. 2022, 430, 132643. [Google Scholar] [CrossRef]
  34. Luo, J.; Wang, M.; Chen, L.; Shi, J. Efficient benzaldehyde photosynthesis coupling photocatalytic hydrogen evolution. J. Energy Chem. 2022, 66, 52–60. [Google Scholar] [CrossRef]
  35. Lima, M.J.; Tavares, P.B.; Silva, A.M.; Silva, C.G.; Faria, J.L. Selective photocatalytic oxidation of benzyl alcohol to benzaldehyde by using metal-loaded g-C3N4 photocatalysts. Catal. Today 2017, 287, 70–77. [Google Scholar] [CrossRef]
  36. Chen, J.; Qiu, F.; Xu, W.; Cao, S.; Zhu, H. Recent progress in enhancing photocatalytic efficiency of TiO2-based materials. Appl. Catal. A Gen. 2015, 495, 131–140. [Google Scholar] [CrossRef]
  37. Dong, P.; Hou, G.; Xi, X.; Shao, R.; Dong, F. WO3-based photocatalysts: Morphology control, activity enhancement and multifunctional applications. Environ. Sci. Nano 2017, 4, 539–557. [Google Scholar] [CrossRef]
  38. Zhong, Y.; Wang, J.; Zhang, R.; Wei, W.; Wang, H.; Lü, X.; Bai, F.; Wu, H.; Haddad, R.; Fan, H. Morphology-controlled self-assembly and synthesis of photocatalytic nanocrystals. Nano Lett. 2014, 14, 7175–7179. [Google Scholar] [CrossRef]
  39. Li, X.; Zheng, W.; He, G.; Zhao, R.; Liu, D. Morphology control of TiO2 nanoparticle in microemulsion and its photocatalytic property. ACS Sustain. Chem. Eng. 2014, 2, 288–295. [Google Scholar] [CrossRef]
  40. Rauf, M.A.; Meetani, M.A.; Hisaindee, S. An overview on the photocatalytic degradation of azo dyes in the presence of TiO2 doped with selective transition metals. Desalination 2011, 276, 13–27. [Google Scholar] [CrossRef]
  41. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S.C. Photocatalytic hydrogen production using metal doped TiO2: A review of recent advances. Appl. Catal. B 2019, 244, 1021–1064. [Google Scholar] [CrossRef]
  42. Nah, Y.C.; Paramasivam, I.; Schmuki, P. Doped TiO2 and TiO2 nanotubes: Synthesis and applications. ChemPhysChem 2010, 11, 2698–2713. [Google Scholar] [CrossRef]
  43. Woan, K.; Pyrgiotakis, G.; Sigmund, W. Photocatalytic carbon-nanotube–TiO2 composites. Adv. Mater. 2009, 21, 2233–2239. [Google Scholar] [CrossRef]
  44. Tang, B.; Chen, H.; Peng, H.; Wang, Z.; Huang, W. Graphene modified TiO2 composite photocatalysts: Mechanism, progress and perspective. Nanomaterials 2018, 8, 105. [Google Scholar] [CrossRef]
  45. He, J.; Kumar, A.; Khan, M.; Lo, I.M. Critical review of photocatalytic disinfection of bacteria: From noble metals-and carbon nanomaterials-TiO2 composites to challenges of water characteristics and strategic solutions. Sci. Total Environ. 2021, 758, 143953. [Google Scholar] [CrossRef]
  46. Tekin, D.; Birhan, D.; Kiziltas, H. Thermal, photocatalytic, and antibacterial properties of calcinated nano-TiO2/polymer composites. Mat. Chem. Phys. 2020, 251, 123067. [Google Scholar] [CrossRef]
  47. Murad, E. Identification of minor amounts of anatase in kaolins by Raman spectroscopy. Am. Mineral. 1997, 82, 203–206. [Google Scholar] [CrossRef]
  48. Lu, Y.; Jaeckel, B.; Parkinson, B.A. Preparation and characterization of terraced surfaces of low-index faces of anatase, rutile, and brookite. Langmuir 2006, 22, 4472–4475. [Google Scholar] [CrossRef]
  49. Haggerty, J.E.; Schelhas, L.T.; Kitchaev, D.A.; Mangum, J.S.; Garten, L.M.; Sun, W. High-fraction brookite films from amorphous precursors. Sci. Rep. 2017, 7, 15232. [Google Scholar] [CrossRef]
  50. Permana, M.D.; Noviyanti, A.R.; Lestari, P.R.; Kumada, N.; Eddy, D.R.; Rahayu, I. The Effect of Organometallic Precursors on TiO2 Synthesis and Their Photocatalytic Activity on Phenol Degradation. Kuwait J. Sci. 2022, 49, 1–13. [Google Scholar]
  51. Deo, G.; Turek, A.M.; Wachs, I.E.; Machej, T.; Haber, J.; Das, N.; Eckert, H.; Hirt, A.M. Physical and chemical characterization of surface vanadium oxide supported on titania: Influence of the titania phase (anatase, rutile, brookite and B). Appl. Catal. A Gen. 1992, 91, 27–42. [Google Scholar] [CrossRef]
  52. Dachille, F.; Simons, P.Y.; Roy, R. Pressure-temperature studies of anatase, brookite, rutile and TiO2-II. Am. Mineral. 1968, 53, 1929–1939. [Google Scholar]
  53. Akpan, U.G.; Hameed, B.H. Parameters affecting the photocatalytic degradation of dyes using TiO2-based photocatalysts: A review. J. Hazard. Mater. 2009, 170, 520–529. [Google Scholar] [CrossRef]
  54. Sun, J.; Gao, L.; Zhang, Q. Synthesizing and comparing the photocatalytic properties of high surface area rutile and anatase titania nanoparticles. J. Am. Ceram. Soc. 2003, 86, 1677–1682. [Google Scholar] [CrossRef]
  55. Maver, K.; Arčon, I.; Fanetti, M.; Emin, S.; Valant, M.; Štangar, U.L. Improved photocatalytic activity of anatase-rutile nanocomposites induced by low-temperature sol-gel Sn-modification of TiO2. Catal. Today 2021, 361, 124–129. [Google Scholar] [CrossRef]
  56. Luttrell, T.; Halpegamage, S.; Tao, J.; Kramer, A.; Sutter, E.; Batzill, M. Why is anatase a better photocatalyst than rutile?-Model studies on epitaxial TiO2 films. Sci. Rep. 2014, 4, 1–8. [Google Scholar] [CrossRef]
  57. Al-Mamun, M.R.; Kader, S.; Islam, M.S.; Khan, M.Z.H. Photocatalytic activity improvement and application of UV-TiO2 photocatalysis in textile wastewater treatment: A review. J. Environ. Chem. Eng. 2019, 7, 103248. [Google Scholar] [CrossRef]
  58. Lee, M.S.; Hong, S.-S.; Mohseni, M. Synthesis of photocatalytic nanosized TiO2–Ag particles with sol–gel method using reduction agent. J. Mol. Catal. A: Chem. 2005, 242, 135–140. [Google Scholar] [CrossRef]
  59. Roy, P.; Ho, L.; Periasamy, A.P.; Lin, Y.; Huang, M.; Chang, H. Graphene-ZnO-Au nanocomposites based photocatalytic oxidation of benzoic acid. ScienceJet 2015, 4, 120. [Google Scholar]
  60. Khedr, T.M.; El-Sheikh, S.M.; Hakki, A.; Ismail, A.A.; Badawy, W.A.; Bahnemann, D.W. Highly active non-metals doped mixed-phase TiO2 for photocatalytic oxidation of ibuprofen under visible light. J. Photochem. Photobiol. A Chem. 2017, 346, 530–540. [Google Scholar] [CrossRef]
  61. Yang, D.; Liu, H.; Zheng, Z.; Yuan, Y.; Zhao, J.C.; Waclawik, E.R.; Ke, X.; Zhu, H. An efficient photocatalyst structure: TiO2 (B) nanofibers with a shell of anatase nanocrystals. J. Am. Chem. Soc. 2009, 131, 17885–17893. [Google Scholar] [CrossRef]
  62. Kang, X.; Dong, G.; Dong, T. Oxygen Vacancy Defect Engineering of Heterophase Junction TiO2: Interfacial/Surface Oxygen Vacancies Coadjust the Photocatalytic ROS Production. ACS Appl. Energy Mater. 2022, 6, 1025–1036. [Google Scholar] [CrossRef]
  63. Ding, L.; Yang, S.; Liang, Z.; Qian, X.; Chen, X.; Cui, H.; Tian, J. TiO2 nanobelts with anatase/rutile heterophase junctions for highly efficient photocatalytic overall water splitting. J. Colloid Interface Sci. 2020, 567, 181–189. [Google Scholar] [CrossRef]
  64. Berger, T.; Sterrer, M.; Diwald, O.; Knözinger, E.; Panayotov, D.; Thompson, T.L.; Yates, J.T. Light-induced charge separation in anatase TiO2 particles. J. Phys. Chem. B 2005, 109, 6061–6068. [Google Scholar] [CrossRef]
  65. Cao, F.; Xiong, J.; Wu, F.; Liu, Q.; Shi, Z.; Yu, Y.; Wang, X.; Li, L. Enhanced photoelectrochemical performance from rationally designed anatase/rutile TiO2 heterostructures. ACS Appl. Mater. Interfaces 2016, 8, 12239–12245. [Google Scholar] [CrossRef]
  66. Tian, G.; Fu, H.; Jing, L.; Xin, B.; Pan, K. Preparation and characterization of stable biphase TiO2 photocatalyst with high crystallinity, large surface area, and enhanced photoactivity. J. Phys. Chem. C 2008, 112, 3083–3089. [Google Scholar] [CrossRef]
  67. Li, W.; Liu, C.; Zhou, Y.; Bai, Y.; Feng, X.; Yang, Z.; Lu, L.; Lu, X.; Chan, K.Y. Enhanced photocatalytic activity in anatase/TiO2 (B) core−shell nanofiber. J. Phys. Chem. C 2008, 112, 20539–20545. [Google Scholar] [CrossRef]
  68. Mu, R.; Ao, Y.; Wu, T.; Wang, C.; Wang, P. Synthesis of novel ternary heterogeneous anatase-TiO2 (B) biphase nanowires/Bi4O5I2 composite photocatalysts for the highly efficient degradation of acetaminophen under visible light irradiation. J. Hazard. Mater. 2020, 382, 121083. [Google Scholar] [CrossRef]
  69. Chen, J.; Guan, M.; Zhang, X.; Gong, X. Insights into a rutile/brookite homojunction of titanium dioxide: Separated reactive sites and boosted photocatalytic activity. RSC Adv. 2019, 9, 36615–36620. [Google Scholar] [CrossRef]
  70. Fischer, K.; Gawel, A.; Rosen, D.; Krause, M.; Abdul Latif, A.; Griebel, J.; Prager, A.; Schulze, A. Low-temperature synthesis of anatase/rutile/brookite TiO2 nanoparticles on a polymer membrane for photocatalysis. Catalysts 2017, 7, 209. [Google Scholar] [CrossRef]
  71. An, X.; Hu, C.; Liu, H.; Qu, J. Hierarchical nanotubular anatase/rutile/TiO2 (B) heterophase junction with oxygen vacancies for enhanced photocatalytic H2 production. Langmuir 2018, 34, 1883–1889. [Google Scholar] [CrossRef]
  72. Wang, C.; Zhang, X.; Liu, Y. Coexistence of an anatase/TiO2 (B) heterojunction and an exposed (001) facet in TiO2 nanoribbon photocatalysts synthesized via a fluorine-free route and topotactic transformation. Nanoscale 2014, 6, 5329–5337. [Google Scholar] [CrossRef] [PubMed]
  73. Wang, Z.; Wang, Y.; Zhang, W.; Wang, Z.; Ma, Y.; Zhou, X. Fabrication of TiO2 (B)/anatase heterophase junctions at high temperature via stabilizing the surface of TiO2 (B) for enhanced photocatalytic activity. J. Phys. Chem. C 2019, 123, 1779–1789. [Google Scholar] [CrossRef]
  74. Tanemura, S.; Miao, L.; Wunderlich, W.; Tanemura, M.; Mori, Y.; Toh, S.; Kaneko, K. Fabrication and characterization of anatase/rutile–TiO2 thin films by magnetron sputtering: A review. Sci. Technol. Adv. Mater. 2005, 6, 11–17. [Google Scholar] [CrossRef]
  75. Yamakata, A.; Vequizo, J.J.M. Curious behaviors of photogenerated electrons and holes at the defects on anatase, rutile, and brookite TiO2 powders: A review. J. Photochem. Photobiol. C Photochem. Rev. 2019, 40, 234–243. [Google Scholar] [CrossRef]
  76. Zhang, D.; Dong, S. Challenges in band alignment between semiconducting materials: A case of rutile and anatase TiO2. Prog. Nat. Sci. Mater. Int. 2019, 29, 277–284. [Google Scholar] [CrossRef]
  77. Basavarajappa, P.S.; Patil, S.B.; Ganganagappa, N.; Reddy, K.R.; Raghu, A.V.; Reddy, C.V. Recent progress in metal-doped TiO2, non-metal doped/co-doped TiO2 and TiO2 nanostructured hybrids for enhanced photocatalysis. Int. J. Hydrogen Energy 2020, 45, 7764–7778. [Google Scholar] [CrossRef]
  78. Connelly, K.; Wahab, A.K.; Idriss, H. Photoreaction of Au/TiO2 for hydrogen production from renewables: A review on the synergistic effect between anatase and rutile phases of TiO2. Mater. Renew. Sustain. Energy 2012, 1, 1–12. [Google Scholar]
  79. Mazierski, P.; Mikolajczyk, A.; Bajorowicz, B.; Malankowska, A.; Zaleska-Medynska, A.; Nadolna, J. The role of lanthanides in TiO2-based photocatalysis: A review. Appl. Catal. B 2018, 233, 301–317. [Google Scholar] [CrossRef]
  80. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. Dual cocatalysts in TiO2 photocatalysis. Adv. Mater. 2019, 31, 1807660. [Google Scholar] [CrossRef]
  81. Zaleska, A. Doped-TiO2: A review. Recent Pat. Eng. 2008, 2, 157–164. [Google Scholar] [CrossRef]
  82. Adachi, S. Physical Properties of III–V Semiconductor Compounds; John Wiley & Sons: New York, NY, USA, 1992. [Google Scholar]
  83. Ma, Y.; Wang, X.; Jia, Y.; Chen, X.; Han, H.; Li, C. Titanium dioxide-based nanomaterials for photocatalytic fuel generations. Chem. Rev. 2014, 114, 9987–10043. [Google Scholar] [CrossRef] [PubMed]
  84. Smith, A.M.; Nie, S. Semiconductor nanocrystals: Structure, properties, and band gap engineering. Acc. Chem. Res. 2010, 43, 190–200. [Google Scholar] [CrossRef] [PubMed]
  85. Sitt, A.; Hadar, I.; Banin, U. Band-gap engineering, optoelectronic properties and applications of colloidal heterostructured semiconductor nanorods. Nano Today 2013, 8, 494–513. [Google Scholar] [CrossRef]
  86. Edgar, J.H. Prospects for device implementation of wide band gap semiconductors. J. Mater. Res. 1992, 7, 235–252. [Google Scholar] [CrossRef]
  87. Ajmal, A.; Majeed, I.; Malik, R.N.; Idriss, H.; Nadeem, M.A. Principles and mechanisms of photocatalytic dye degradation on TiO2 based photocatalysts: A comparative overview. RSC Adv. 2014, 4, 37003–37026. [Google Scholar] [CrossRef]
  88. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 photocatalysis: Concepts, mechanisms, and challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef]
  89. Linsebigler, A.L.; Lu, G.; Yates Jr, J.T. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  90. Kumar, A.; Khan, M.; He, J.; Lo, I.M. Recent developments and challenges in practical application of visible–light–driven TiO2–based heterojunctions for PPCP degradation: A critical review. Water Res. 2020, 170, 115356. [Google Scholar] [CrossRef]
  91. Liu, L.; Li, Y. Understanding the reaction mechanism of photocatalytic reduction of CO2 with H2O on TiO2-based photocatalysts: A review. Aerosol Air Qual. Res. 2014, 14, 453–469. [Google Scholar] [CrossRef]
  92. Pawar, R.R.; Kim, M.; Kim, J.G.; Hong, S.M.; Sawant, S.Y.; Lee, S.M. Efficient removal of hazardous lead, cadmium, and arsenic from aqueous environment by iron oxide modified clay-activated carbon composite beads. Appl. Clay Sci. 2018, 162, 339–350. [Google Scholar] [CrossRef]
  93. Zhang, J.; Tian, B.; Wang, L.; Xing, M.; Lei, J. Mechanism of photocatalysis. In Photocatalysis: Fundamentals, Materials and Applications; Zhang, J., Tian, B., Wang, L., Xing, M., Lei, J., Eds.; Springer: Singapore, 2018; pp. 1–15. [Google Scholar]
  94. Regmi, C.; Joshi, B.; Ray, S.K.; Gyawali, G.; Pandey, R.P. Understanding mechanism of photocatalytic microbial decontamination of environmental wastewater. Front. Chem. 2018, 6, 33. [Google Scholar] [CrossRef] [PubMed]
  95. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 photocatalysis: Mechanisms and materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  96. Wenderich, K.; Mul, G. Methods, mechanism, and applications of photodeposition in photocatalysis: A review. Chem. Rev. 2016, 116, 14587–14619. [Google Scholar] [CrossRef] [PubMed]
  97. Zhu, T.; Ye, X.; Zhang, Q.; Hui, Z.; Wang, X.; Chen, S. Efficient utilization of photogenerated electrons and holes for photocatalytic redox reactions using visible light-driven Au/ZnIn2S4 hybrid. J. Hazard. Mater. 2019, 367, 277–285. [Google Scholar] [CrossRef]
  98. Gerischer, H. Electron-transfer kinetics of redox reactions at the semiconductor/electrolyte contact. A new approach. J. Phys. Chem. 1991, 95, 1356–1359. [Google Scholar] [CrossRef]
  99. Hagfeldt, A.; Graetzel, M. Light-induced redox reactions in nanocrystalline systems. Chem. Rev. 1995, 95, 49–68. [Google Scholar] [CrossRef]
  100. Ge, M.; Cao, C.; Huang, J.; Li, S.; Chen, Z.; Zhang, K.Q.; Al-Deyab, S.S.; Lai, Y. A review of one-dimensional TiO2 nanostructured materials for environmental and energy applications. J. Mater. Chem. A 2016, 4, 6772–6801. [Google Scholar] [CrossRef]
  101. Ahmed, S.; Rasul, M.G.; Martens, W.N.; Brown, R.; Hashib, M.A. Heterogeneous photocatalytic degradation of phenols in wastewater: A review on current status and developments. Desalination 2010, 261, 3–18. [Google Scholar] [CrossRef]
  102. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  103. Feng, X.; Guo, H.; Patel, K.; Zhou, H.; Lou, X. High performance, recoverable Fe3O4ZnO nanoparticles for enhanced photocatalytic degradation of phenol. Chem. Eng. J. 2014, 244, 327–334. [Google Scholar] [CrossRef]
  104. Zhang, J.; Nosaka, Y. Mechanism of the OH radical generation in photocatalysis with TiO2 of different crystalline types. J. Phys. Chem. C 2014, 118, 10824–10832. [Google Scholar] [CrossRef]
  105. Sobczyński, A.; Duczmal, Ł.; Zmudziński, W. Phenol destruction by photocatalysis on TiO2: An attempt to solve the reaction mechanism. J. Mol. Catal. A Chem. 2004, 213, 225–230. [Google Scholar] [CrossRef]
  106. Gong, M.; Xiao, S.; Yu, X.; Dong, C.; Ji, J.; Zhang, D.; Xing, M. Research progress of photocatalytic sterilization over semiconductors. RSC Adv. 2019, 9, 19278–19284. [Google Scholar] [CrossRef] [PubMed]
  107. Dambournet, D.; Belharouak, I.; Amine, K. Tailored preparation methods of TiO2 anatase, rutile, brookite: Mechanism of formation and electrochemical properties. Chem. Mater. 2010, 22, 1173–1179. [Google Scholar] [CrossRef]
  108. Reyes-Coronado, D.; Rodríguez-Gattorno, G.; Espinosa-Pesqueira, M.E.; Cab, C.; De Coss, R.D.; Oskam, G. Phase-pure TiO2 nanoparticles: Anatase, brookite and rutile. Nanotechnology 2008, 19, 145605. [Google Scholar] [CrossRef]
  109. Tompsett, G.A.; Bowmaker, G.A.; Cooney, R.P.; Metson, J.B.; Rodgers, K.A.; Seakins, J.M. The Raman spectrum of brookite, TiO2 (PBCA, Z = 8). J. Raman Spectrosc. 1995, 26, 57–62. [Google Scholar] [CrossRef]
  110. Beltran, A.; Gracia, L.; Andres, J. Density functional theory study of the brookite surfaces and phase transitions between natural titania polymorphs. J. Phys. Chem. B 2006, 110, 23417–23423. [Google Scholar] [CrossRef]
  111. Monai, M.; Montini, T.; Fornasiero, P. Brookite: Nothing new under the sun? Catalysts 2017, 7, 304. [Google Scholar] [CrossRef]
  112. Li, Z.; Cong, S.; Xu, Y. Brookite vs anatase TiO2 in the photocatalytic activity for organic degradation in water. ACS Catal. 2014, 4, 3273–3280. [Google Scholar] [CrossRef]
  113. Banfield, J.F.; Veblen, D.R.; Smith, D.J. The identification of naturally occurring TiO2 (B) by structure determination using high-resolution electron microscopy, image simulation, and distance-least-squares refinement. Am. Mineral. 1991, 76, 343–353. [Google Scholar]
  114. Xiang, G.; Wang, Y.G.; Li, J.; Zhuang, J.; Wang, X. Surface-specific interaction by structure-match confined pure high-energy facet of unstable TiO2 (B) polymorph. Sci. Rep. 2013, 3, 1411. [Google Scholar] [CrossRef] [PubMed]
  115. Armstrong, A.R.; Armstrong, G.; Canales, J.; Bruce, P.G. TiO2-B nanowires. Angew. Chem. Int. Ed. 2004, 43, 2286–2288. [Google Scholar] [CrossRef] [PubMed]
  116. Nosheen, S.; Galasso, F.S.; Suib, S.L. Role of Ti-O bonds in phase transitions of TiO2. Langmuir 2009, 25, 7623–7630. [Google Scholar] [CrossRef] [PubMed]
  117. Opra, D.P.; Gnedenkov, S.V.; Sinebryukhov, S.L. Recent efforts in design of TiO2 (B) anodes for high-rate lithium-ion batteries: A review. J. Power Sources 2019, 442, 227225. [Google Scholar] [CrossRef]
  118. Kim, W.; Tachikawa, T.; Moon, G.H.; Majima, T.; Choi, W. Molecular-level understanding of the photocatalytic activity difference between anatase and rutile Nanoparticles. Angew. Chem. 2014, 126, 14260–14265. [Google Scholar] [CrossRef]
  119. Sclafani, A.; Herrmann, J.M. Comparison of the photoelectronic and photocatalytic activities of various anatase and rutile forms of titania in pure liquid organic phases and in aqueous solutions. J. Phys. Chem. 1996, 100, 13655–13661. [Google Scholar] [CrossRef]
  120. Zhang, J.; Zhou, P.; Liu, J.; Yu, J. New understanding of the difference of photocatalytic activity among anatase, rutile and brookite TiO2. Phys. Chem. Chem. Phys. 2014, 16, 20382–20386. [Google Scholar] [CrossRef]
  121. Wu, X.; Ng, Y.H.; Wang, L.; Du, Y.; Dou, S.X.; Amal, R.; Scott, J. Improving the photo-oxidative capability of BiOBr via crystal facet engineering. J. Mater. Chem. A 2017, 5, 8117–8124. [Google Scholar] [CrossRef]
  122. Zhang, L.; Ran, J.; Qiao, S.Z.; Jaroniec, M. Characterization of semiconductor photocatalysts. Chem. Soc. Rev. 2019, 48, 5184–5206. [Google Scholar] [CrossRef]
  123. Wang, L.; Zhao, J.; Liu, H.; Huang, J. Design, modification and application of semiconductor photocatalysts. J. Taiwan Inst. Chem. Eng. 2018, 93, 590–602. [Google Scholar] [CrossRef]
  124. Comparelli, R.; Fanizza, E.; Curri, M.L.; Cozzoli, P.D.; Mascolo, G.; Passino, R.; Agostiano, A. Photocatalytic degradation of azo dyes by organic-capped anatase TiO2 nanocrystals immobilized onto substrates. Appl. Catal. B 2005, 55, 81–91. [Google Scholar] [CrossRef]
  125. Katal, R.; Masudy-Panah, S.; Tanhaei, M.; Farahani, M.H.D.A.; Jiangyong, H. A review on the synthesis of the various types of anatase TiO2 facets and their applications for photocatalysis. Chem. Eng. J. 2020, 384, 123384. [Google Scholar] [CrossRef]
  126. Porkodi, K.; Arokiamary, S.D. Synthesis and spectroscopic characterization of nanostructured anatase titania: A photocatalyst. Mater. Charact. 2007, 58, 495–503. [Google Scholar] [CrossRef]
  127. Sivalingam, G.; Nagaveni, K.; Hegde, M.S.; Madras, G. Photocatalytic degradation of various dyes by combustion synthesized nano anatase TiO2. Appl. Catal. B 2003, 45, 23–38. [Google Scholar] [CrossRef]
  128. Periyat, P.; Naufal, B.; Ullattil, S.G. A review on high temperature stable anatase TiO2 photocatalysts. Mater. Sci. Forum 2016, 855, 78–93. [Google Scholar] [CrossRef]
  129. Nosaka, Y.; Nosaka, A.Y. Reconsideration of intrinsic band alignments within anatase and rutile TiO2. J. Phys. Chem. Lett. 2016, 7, 431–434. [Google Scholar] [CrossRef]
  130. Yuan, L.D.; Deng, H.X.; Li, S.S.; Wei, S.H.; Luo, J.W. Unified theory of direct or indirect band-gap nature of conventional semiconductors. Phys. Rev. B 2018, 98, 245203. [Google Scholar] [CrossRef]
  131. Gu, L.; Srot, V.; Sigle, W.; Koch, C.; van Aken, P.; Scholz, F.; Thapa, S.B.; Kirchner, C.; Jetter, M.; Rühle, M. Band-gap measurements of direct and indirect semiconductors using monochromated electrons. Phys. Rev. B 2007, 75, 195214. [Google Scholar] [CrossRef]
  132. López, R.; Gómez, R. Band-gap energy estimation from diffuse reflectance measurements on sol–gel and commercial TiO2: A comparative study. J. Sol-Gel Sci. Technol. 2012, 61, 1–7. [Google Scholar] [CrossRef]
  133. Zhou, W.; Umezawa, N.; Ma, R.; Sakai, N.; Ebina, Y.; Sano, K.; Liu, M.; Ishida, Y.; Aida, T.; Sasaki, T. Spontaneous direct band gap, high hole mobility, and huge exciton energy in atomic-thin TiO2 nanosheet. Chem. Mater. 2018, 30, 6449–6457. [Google Scholar] [CrossRef]
  134. Wetchakun, N.; Incessungvorn, B.; Wetchakun, K.; Phanichphant, S. Influence of calcination temperature on anatase to rutile phase transformation in TiO2 nanoparticles synthesized by the modified sol–gel method. Mater. Lett. 2012, 82, 195–198. [Google Scholar] [CrossRef]
  135. Bagheri, S.; Julkapli, N.M. Mixed-phase TiO2 photocatalysis: Correlation between phase composition and photodecomposition of water pollutants. Rev. Inorg. Chem. 2017, 37, 11–28. [Google Scholar] [CrossRef]
  136. Ou, X.; Liu, X.; Liu, W.; Rong, W.; Li, J.; Lin, Z. Surface defects enhance the adsorption affinity and selectivity of Mg(OH)2 towards As (V) and Cr (VI) oxyanions: A combined theoretical and experimental study. Environ. Sci. Nano 2018, 5, 2570–2578. [Google Scholar] [CrossRef]
  137. Yang, H.G.; Sun, C.H.; Qiao, S.Z.; Zou, J.; Liu, G.; Smith, S.C.; Cheng, H.M.; Lu, G.Q. Anatase TiO2 single crystals with a large percentage of reactive facets. Nature 2008, 453, 638–641. [Google Scholar] [CrossRef] [PubMed]
  138. Fagan, R.; Synnott, D.W.; McCormack, D.E.; Pillai, S.C. An effective method for the preparation of high temperature stable anatase TiO2 photocatalysts. Appl. Surf. Sci. 2016, 371, 447–452. [Google Scholar] [CrossRef]
  139. Bubacz, K.; Choina, J.; Dolat, D.; Morawski, A.W. Methylene Blue and Phenol Photocatalytic Degradation on Nanoparticles of Anatase TiO2. Pol. J. Environ. Stud. 2010, 19, 685–691. [Google Scholar]
  140. Etacheri, V.; Seery, M.K.; Hinder, S.J.; Pillai, S.C. Oxygen rich titania: A dopant free, high temperature stable, and visible-light active anatase photocatalyst. Adv. Funct. Mater. 2011, 21, 3744–3752. [Google Scholar] [CrossRef]
  141. Lv, K.; Cheng, B.; Yu, J.; Liu, G. Fluorine ions-mediated morphology control of anatase TiO2 with enhanced photocatalytic activity. Phys. Chem. Chem. Phys. 2012, 14, 5349–5362. [Google Scholar] [CrossRef]
  142. Wang, X.; Kafizas, A.; Li, X.; Moniz, S.J.; Reardon, P.J.; Tang, J.; Parkin, I.P.; Durrant, J.R. Transient absorption spectroscopy of anatase and rutile: The impact of morphology and phase on photocatalytic activity. J. Phys. Chem. C 2015, 119, 10439–10447. [Google Scholar] [CrossRef]
  143. Bickley, R.I.; Stone, F.S. Photoadsorption and photocatalysis at rutile surfaces: I. Photoadsorption of oxygen. J. Catal. 1973, 31, 389–397. [Google Scholar] [CrossRef]
  144. Bickley, R.I.; Munuera, G.; Stone, F.S. Photoadsorption and photocatalysis at rutile surfaces: II. Photocatalytic oxidation of isopropanol. J. Catal. 1973, 31, 398–407. [Google Scholar] [CrossRef]
  145. Amano, F.; Nakata, M.; Yamamoto, A.; Tanaka, T. Rutile titanium dioxide prepared by hydrogen reduction of Degussa P25 for highly efficient photocatalytic hydrogen evolution. Catal. Sci. Technol. 2016, 6, 5693–5699. [Google Scholar] [CrossRef]
  146. Shi, J.; Chen, J.; Feng, Z.; Chen, T.; Lian, Y.; Wang, X.; Li, C. Photoluminescence characteristics of TiO2 and their relationship to the photoassisted reaction of water/methanol mixture. J. Phys. Chem. C 2007, 111, 693–699. [Google Scholar] [CrossRef]
  147. Park, N.G.; Van de Lagemaat, J.; Frank, A.A. Comparison of dye-sensitized rutile-and anatase-based TiO2 solar cells. J. Phys. Chem. B 2000, 104, 8989–8994. [Google Scholar] [CrossRef]
  148. Zhao, J.; Liu, S.; Zhang, X.; Xu, Y. Different effects of fluoride and phosphate anions on TiO2 photocatalysis (rutile). Catal. Sci. Technol. 2020, 10, 6552–6561. [Google Scholar] [CrossRef]
  149. Wang, Y.; Zhang, L.; Deng, K.; Chen, X.; Zou, Z. Low temperature synthesis and photocatalytic activity of rutile TiO2 nanorod superstructures. J. Phys. Chem. C 2007, 111, 2709–2714. [Google Scholar] [CrossRef]
  150. Miyoshi, A.; Nishioka, S.; Maeda, K. Water splitting on rutile TiO2-based photocatalysts. Chem. Eur. J. 2018, 24, 18204–18219. [Google Scholar] [CrossRef]
  151. Liu, Q.Y.; Wang, H.D.; Tang, R.; Cheng, Q.; Yuan, Y.J. Rutile TiO2 nanoparticles with oxygen vacancy for photocatalytic nitrogen fixation. ACS Appl. Nano Mater. 2021, 4, 8674–8679. [Google Scholar] [CrossRef]
  152. Yurdakal, S.; Palmisano, G.; Loddo, V.; Augugliaro, V.; Palmisano, L. Nanostructured rutile TiO2 for selective photocatalytic oxidation of aromatic alcohols to aldehydes in water. J. Am. Chem. Soc. 2008, 130, 1568–1569. [Google Scholar] [CrossRef]
  153. Jung, H.S.; Kim, H. Origin of low photocatalytic activity of rutile TiO2. Electron. Mater. Lett. 2009, 5, 73–76. [Google Scholar] [CrossRef]
  154. Zhao, Z.; Zhang, X.; Zhang, G.; Liu, Z.; Qu, D.; Miao, X.; Feng, P.; Sun, Z. Effect of defects on photocatalytic activity of rutile TiO2 nanorods. Nano Res. 2015, 8, 4061–4071. [Google Scholar] [CrossRef]
  155. Djokić, V.R.; Marinković, A.D.; Petrović, R.D.; Ersen, O.; Zafeiratos, S.; Mitrić, M.; Ophus, C.; Radmilović, V.R.; Janaćković, D.T. Highly active rutile TiO2 nanocrystalline photocatalysts. ACS Appl. Mater. Interfaces 2020, 12, 33058–33068. [Google Scholar] [CrossRef] [PubMed]
  156. Zhang, J.; Song, Y.; Lu, F.; Fei, W.; Mengqiong, Y.; Genxiang, L.; Qian, X.; Xiang, W.; Can, L. Photocatalytic degradation of rhodamine B on anatase, rutile, and brookite TiO2. Chinese J. Catal. 2011, 32, 983–991. [Google Scholar] [CrossRef]
  157. Pauling, L.; Sturdivant, J.H., XV. The crystal structure of brookite. Z. Kristallogr. Cryst. Mater. 1928, 68, 239–256. [Google Scholar] [CrossRef]
  158. Di Paola, A.; Bellardita, M.; Palmisano, L. Brookite, the least known TiO2 photocatalyst. Catalysts 2013, 3, 36–73. [Google Scholar] [CrossRef]
  159. Zhang, H.; Banfield, J.F. Understanding polymorphic phase transformation behavior during growth of nanocrystalline aggregates: Insights from TiO2. J. Phys. Chem. B 2000, 104, 3481–3487. [Google Scholar] [CrossRef]
  160. Xie, J.; Lü, X.; Liu, J.; Shu, H. Brookite titania photocatalytic nanomaterials: Synthesis, properties, and applications. Pure Appl. Chem. 2009, 81, 2407–2415. [Google Scholar] [CrossRef]
  161. Thuong, H.T.T.; Kim, C.T.T.; Quang, L.N.; Kosslick, H. Highly active brookite TiO2-assisted photocatalytic degradation of dyes under the simulated solar−UVA radiation. Prog. Nat. Sci. Mater. Int. 2019, 29, 641–647. [Google Scholar] [CrossRef]
  162. Kandiel, T.A.; Robben, L.; Alkaima, A.; Bahnemann, D. Brookite versus anatase TiO2 photocatalysts: Phase transformations and photocatalytic activities. Photochem. Photobiol. Sci. 2013, 12, 602–609. [Google Scholar] [CrossRef]
  163. Khan, S.; Je, M.; Kim, D.; Lee, S.; Cho, S.H.; Song, T.; Choi, H. Mapping point defects of brookite TiO2 for photocatalytic activity beyond anatase and P25. J. Phys. Chem. C 2020, 124, 10376–10384. [Google Scholar] [CrossRef]
  164. Zhuang, B.; Shi, H.; Zhang, H.; Zhang, Z. Sodium doping in brookite TiO2 enhances its photocatalytic activity. Beilstein J. Nanotechnol. 2022, 13, 599–609. [Google Scholar] [CrossRef] [PubMed]
  165. Žerjav, G.; Žižek, K.; Zavašnik, J.; Pintar, A. Brookite vs. rutile vs. anatase: Whats behind their various photocatalytic activities? J. Environ. Chem. Eng. 2022, 10, 107722. [Google Scholar] [CrossRef]
  166. Bellardita, M.; Di Paola, A.; Megna, B.; Palmisano, L. Absolute crystallinity and photocatalytic activity of brookite TiO2 samples. Appl. Catal. B 2017, 201, 150–158. [Google Scholar] [CrossRef]
  167. Choi, M.; Lim, J.; Baek, M.; Choi, W.; Kim, W.; Yong, K. Investigating the unrevealed photocatalytic activity and stability of nanostructured brookite TiO2 film as an environmental photocatalyst. ACS Appl. Mater. Interfaces 2017, 9, 16252–16260. [Google Scholar] [CrossRef]
  168. Do, H.H.; Tran, T.K.C.; Ung, T.D.T.; Dao, N.T.; Nguyen, D.D.; Trinh, T.H.; Hoang, T.D.; Le, T.L.; Tran, T.T.H. Controllable fabrication of photocatalytic TiO2 brookite thin film by 3D-printing approach for dyes decomposition. J. Water Process. Eng. 2021, 43, 102319. [Google Scholar] [CrossRef]
  169. Marchand, R.; Brohan, L.; Tournoux, M. TiO2 (B) a new form of titanium dioxide and the potassium octatitanate K2Ti8O17. Mater. Res. Bull. 1980, 15, 1129–1133. [Google Scholar] [CrossRef]
  170. Dylla, A.G.; Henkelman, G.; Stevenson, K.J. Lithium insertion in nanostructured TiO2 (B) architectures. Acc. Chem. Res. 2013, 46, 1104–1112. [Google Scholar] [CrossRef]
  171. Fehse, M.; Ventosa, E. Is TiO2 (B) the future of titanium-based battery materials? ChemPlusChem 2015, 80, 785–795. [Google Scholar] [CrossRef] [PubMed]
  172. Hua, X.; Liu, Z.; Bruce, P.G.; Grey, C.P. The morphology of TiO2 (B) nanoparticles. J. Am. Chem. Soc. 2015, 137, 13612–13623. [Google Scholar] [CrossRef] [PubMed]
  173. Liu, S.; Jia, H.; Han, L.; Wang, J.; Gao, P.; Xu, D.; Yang, J.; Che, S. Nanosheet-constructed porous TiO2–B for advanced lithium-ion batteries. Adv. Mater. 2012, 24, 3201–3204. [Google Scholar] [CrossRef] [PubMed]
  174. Ren, Y.; Liu, Z.; Pourpoint, F.; Armstrong, A.R.; Grey, C.P.; Bruce, P.G. Nanoparticulate TiO2 (B): An anode for lithium-ion batteries. Angew. Chem. 2012, 124, 2206–2209. [Google Scholar] [CrossRef]
  175. Chakraborty, A.K.; Qi, Z.; Chai, S.Y.; Lee, C.; Park, S.Y.; Jang, D.J.; Lee, W.I. Formation of highly crystallized TiO2 (B) and its photocatalytic behavior. Appl. Catal. B 2010, 93, 368–375. [Google Scholar] [CrossRef]
  176. Zhang, Y.; Xing, Z.; Liu, X.; Li, Z.; Wu, X.; Jiang, J.; Li, M.; Zhu, Q.; Zhou, W. Ti3+ self-doped blue TiO2 (B) single-crystalline nanorods for efficient solar-driven photocatalytic performance. ACS Appl. Mater. Interfaces 2016, 8, 26851–26859. [Google Scholar] [CrossRef] [PubMed]
  177. Lee, J.S.; You, K.H.; Park, C.B. Highly photoactive, low bandgap TiO2 nanoparticles wrapped by graphene. Adv. Mater. 2012, 24, 1084–1088. [Google Scholar] [CrossRef] [PubMed]
  178. Peighambardoust, N.S.; Khameneh Asl, S.; Mohammadpour, R.; Asl, S.K. Band-gap narrowing and electrochemical properties in N-doped and reduced anodic TiO2 nanotube arrays. Electrochim. Acta 2018, 270, 245–255. [Google Scholar] [CrossRef]
  179. Sharma, P.K.; Cortes, M.A.L.R.M.; Hamilton, J.W.J.; Han, Y.; Byrne, J.A.; Nolan, M. Surface modification of TiO2 with copper clusters for band gap narrowing. Catal. Today 2017, 321–322, 9–17. [Google Scholar] [CrossRef]
  180. Vargas Hernández, J.; Coste, S.; García Murillo, A.; Carrillo Romo, F.; Kassiba, A. Effects of metal doping (Cu, Ag, Eu) on the electronic and optical behavior of nanostructured TiO2. J. Alloys Compd. 2017, 710, 355–363. [Google Scholar] [CrossRef]
  181. Zhang, J.; Tian, B.; Wang, L.; Xing, M. Photocatalysis; Springer: Singapore, 2018. [Google Scholar]
  182. Huang, Y.; Cao, J.J.; Kang, F.; You, S.J.; Chang, C.W.; Wang, Y.F. High selectivity of visible-light-driven La-doped TiO2 photocatalysts for NO removal. Aerosol Air Qual. Res. 2017, 17, 2555–2565. [Google Scholar] [CrossRef]
  183. Marques, J.; Gomes, T.D.; Forte, M.A.; Silva, R.F.; Tavares, C.J. A new route for the synthesis of highly-active N-doped TiO2 nanoparticles for visible light photocatalysis using urea as nitrogen precursor. Catal. Today 2019, 326, 36–45. [Google Scholar] [CrossRef]
  184. Liu, D.; Tian, R.; Wang, J.; Nie, E.; Piao, X.; Li, X.; Sun, Z. Photoelectrocatalytic degradation of methylene blue using F doped TiO2 photoelectrode under visible light irradiation. Chemosphere 2017, 185, 574–581. [Google Scholar] [CrossRef]
  185. Boningari, T.; Inturi, S.N.R.; Suidan, M.; Smirniotis, P.G. Novel one-step synthesis of sulfur doped-TiO2 by flame spray pyrolysis for visible light photocatalytic degradation of acetaldehyde. Chem. Eng. J. 2018, 339, 249–258. [Google Scholar] [CrossRef]
  186. Payormhorm, J.; Idem, R. Synthesis of C-doped TiO2 by sol-microwave method for photocatalytic conversion of glycerol to value-added chemicals under visible light. Appl. Catal. A-Gen. 2020, 590, 117362. [Google Scholar] [CrossRef]
  187. Nagaraj, G.; Raj, A.D.; Irudayaraj, A.A.; Josephine, R.L. Tuning the optical band Gap of pure TiO2 via photon induced method. Optik 2018, 179, 889–894. [Google Scholar]
  188. Hu, J.; Zhang, S.; Cao, Y.; Wang, H.; Yu, H.; Peng, F. Novel highly active anatase/rutile TiO2 photocatalyst with hydrogenated heterophase interface structures for photoelectrochemical water splitting into hydrogen. ACS Sustain. Chem. Eng. 2018, 6, 10823–10832. [Google Scholar] [CrossRef]
  189. Lu, S.; Yang, S.; Hu, X.; Liang, Z.; Guo, Y.; Xue, Y.; Cui, H.; Tian, J. Fabrication of TiO2 nanoflowers with bronze (TiO2 (B))/anatase heterophase junctions for efficient photocatalytic hydrogen production. Int. J. Hydrogen Energy 2019, 44, 24398–24406. [Google Scholar] [CrossRef]
  190. Jimenez-Relinque, E.; Castellote, M. Hydroxyl radical and free and shallowly trapped electron generation and electron/hole recombination rates in TiO2 photocatalysis using different combinations of anatase and rutile. Appl. Catal. A-Gen. 2018, 565, 20–25. [Google Scholar] [CrossRef]
  191. Xu, X.; Guo, Y.; Liang, Z.; Cui, H.; Tian, J. Remarkable charge separation and photocatalytic efficiency enhancement through TiO2 (B)/anatase heterophase junctions of TiO2 nanobelts. Int. J. Hydrogen Energy 2019, 44, 27311–27318. [Google Scholar] [CrossRef]
  192. Ruan, X.; Cui, X.; Cui, Y.; Fan, X.; Li, Z.; Xie, T.; Ba, K.; Jia, G.; Zhang, H.; Zhang, L.; et al. Favorable energy band alignment of TiO2 anatase/rutile heterophase homojunctions yields photocatalytic hydrogen evolution with quantum efficiency exceeding 45.6%. Adv. Energy Mater. 2022, 12, 2200298. [Google Scholar] [CrossRef]
  193. Foster, H.A.; Ditta, I.B.; Varghese, S.; Steele, A. Photocatalytic disinfection using titanium dioxide: Spectrum and mechanism of antimicrobial activity. Appl. Microbiol. Biotechnol. 2011, 90, 1847–1868. [Google Scholar] [CrossRef]
  194. Zhang, C.; Li, Y.; Shuai, D.; Shen, Y.; Wang, D. Progress and challenges in photocatalytic disinfection of waterborne Viruses: A review to fill current knowledge gaps. Chem. Eng. J. 2019, 355, 399–415. [Google Scholar] [CrossRef]
  195. Luo, Z.; Poyraz, A.S.; Kuo, C.H.; Miao, R.; Meng, Y.; Chen, S.Y.; Jiang, T.; Wenos, C.; Suib, S.L. Crystalline mixed phase (anatase/rutile) mesoporous titanium dioxides for visible light photocatalytic activity. Chem. Mater. 2015, 27, 6–17. [Google Scholar] [CrossRef]
  196. Franciosi, A.; Van de Walle, C.G. Heterojunction band offset engineering. Surf. Sci. Rep. 1996, 25, 1–140. [Google Scholar] [CrossRef]
  197. Tung, R.T.; Kronik, L. Charge density and band offsets at heterovalent semiconductor interfaces. Adv. Theory Simul. 2018, 1, 1700001. [Google Scholar] [CrossRef]
  198. Scanlon, D.O.; Dunnill, C.W.; Buckeridge, J.; Shevlin, S.A.; Logsdail, A.J.; Woodley, S.M.; Catlow, C.R.A.; Powell, M.; Palgrave, R.G.; Parkin, I.P.; et al. Band alignment of rutile and anatase TiO2. Nat. Mater. 2013, 12, 798–801. [Google Scholar] [CrossRef] [PubMed]
  199. Singh, M.K.; Mehata, M.S. Phase-dependent optical and photocatalytic performance of synthesized titanium dioxide (TiO2) nanoparticles. Optik 2019, 193, 163011. [Google Scholar] [CrossRef]
  200. Sato, T.; Taya, M. Enhancement of phage inactivation using photocatalytic titanium dioxide particles with different crystalline structures. Biochem. Eng. J. 2006, 28, 303–308. [Google Scholar] [CrossRef]
  201. Zhu, W.; Kong, L.; Long, W.; Shi, X.; Zhou, D.; Sun, L.; Zhou, L.; Yang, G.; Liu, X.; Liu, H.; et al. Band Structure Engineering of Black Phosphorus/Graphene/MoS2 van der Waals Heterojunctions for Photovoltaic and Op-toelectronic Device Application. J. Phys. Conf. Ser. 2021, 1865, 022021. [Google Scholar] [CrossRef]
  202. Sun, X.; Chang, Y.; Cheng, Y.; Feng, Y.; Zhang, H. Band alignment-driven oxidative injury to the skin by anatase/rutile mixed-phase titanium dioxide nanoparticles under sunlight exposure. Toxicol. Sci. 2018, 164, 300–312. [Google Scholar] [CrossRef]
  203. Zhang, Z.; Yates, J.T., Jr. Band bending in semiconductors: Chemical and physical consequences at surfaces and interfaces. Chem. Rev. 2012, 112, 5520–5551. [Google Scholar] [CrossRef]
  204. Vittadini, A.; Casarin, M.; Selloni, A. Chemistry of and on TiO2-anatase surfaces by DFT calculations: A partial review. Theor. Chem. Acc. 2007, 117, 663–671. [Google Scholar] [CrossRef]
  205. Ju, M.G.; Sun, G.; Wang, J.; Meng, Q.; Liang, W. Origin of high photocatalytic properties in the mixed-phase TiO2: A first-principles theoretical study. ACS Appl. Mater. Interfaces 2014, 6, 12885–12892. [Google Scholar] [CrossRef] [PubMed]
  206. Deskins, N.A.; Kerisit, S.; Rosso, K.M.; Dupuis, M. Molecular dynamics characterization of rutile-anatase interfaces. J. Phys. Chem. C 2007, 111, 9290–9298. [Google Scholar] [CrossRef]
  207. Davlatshoevich, N.D.; Ashur, K.M.; Saidali, B.A.; Kholmirzotagoykulovich, K.; Lyubchyk, A.; Ibrahim, M. Investigation of structural and optoelectronic properties of N-doped hexagonal phases of TiO2 (TiO2-xNx) nanoparticles with DFT realization: Optimization of the band gap and optical properties for visible-light absorption and photovoltaic applications. Biointerface Res. Appl. Chem. 2022, 12, 3836–3848. [Google Scholar]
  208. Liang, B.; Mianxin, S.; Tianliang, Z.; Xiaoyong, Z.; Qingqing, D. Band gap calculation and photo catalytic activity of rare earths doped rutile TiO2. J. Rare Earths 2009, 27, 461–468. [Google Scholar]
  209. Quesada-Cabrera, R.; Sotelo-Vazquez, C.; Bear, J.C.; Darr, J.A.; Parkin, I.P. Photocatalytic Evidence of the Rutile-to-Anatase Electron Transfer in Titania. Adv. Mater. Interfaces 2014, 1, 1400069. [Google Scholar] [CrossRef]
  210. Ullattil, S.G.; Periyat, P. Sol-gel synthesis of titanium dioxide. In Sol-Gel Materials for Energy, Environment and Electronic Applications; Pillai, S.C., Hehir, S., Eds.; Springer Cham: New York, NY, USA, 2017; pp. 271–283. [Google Scholar]
  211. Wang, C.C.; Ying, J.Y. Sol− gel synthesis and hydrothermal processing of anatase and rutile titania nanocrystals. Chem. Mater. 1999, 11, 3113–3120. [Google Scholar] [CrossRef]
  212. Permana, M.D.; Noviyanti, A.R.; Lestari, P.R.; Kumada, N.; Eddy, D.R.; Rahayu, I. Enhancing the photocatalytic activity of TiO2/Na2Ti6O13 composites by gold for the photodegradation of phenol. ChemEngineering 2022, 6, 69. [Google Scholar] [CrossRef]
  213. Arnal, P.; Corriu, R.J.; Leclercq, D.; Mutin, P.H.; Vioux, A. Preparation of anatase, brookite and rutile at low temperature by non-hydrolytic sol–gel methods. J. Mater. Chem. 1996, 6, 1925–1932. [Google Scholar] [CrossRef]
  214. Castrejón-Sánchez, V.H.; López, R.; Ramón-González, M.; Enríquez-Pérez, Á.; Camacho-López, M.; Villa-Sánchez, G. Annealing control on the anatase/rutile ratio of nanostructured titanium dioxide obtained by sol-gel. Crystals 2018, 9, 22. [Google Scholar] [CrossRef]
  215. Noviyanti, A.R.; Asyiah, E.N.; Permana, M.D.; Dwiyanti, D.; Eddy, D.R. Preparation of Hydroxyapatite-Titanium Dioxide Composite from Eggshell by Hydrothermal Method: Characterization and Antibacterial Activity. Crystals 2022, 12, 1599. [Google Scholar] [CrossRef]
  216. Aruna, S.T.; Tirosh, S.; Zaban, A. Nanosize rutile titania particle synthesis via hydrothermal method without mineralizers. J. Mater. Chem. 2000, 10, 2388–2391. [Google Scholar] [CrossRef]
  217. Li, G.; Ciston, S.; Saponjic, Z.V.; Chen, L.; Dimitrijevic, N.M.; Rajh, T.; Gray, K.A. Synthesizing mixed-phase TiO2 nanocomposites using a hydrothermal method for photo-oxidation and photoreduction applications. J. Catal. 2008, 253, 105–110. [Google Scholar] [CrossRef]
  218. Sun, C.; Wang, N.; Zhou, S.; Hu, X.; Zhou, S.; Chen, P. Preparation of self-supporting hierarchical nanostructured anatase/rutile composite TiO2 film. Chem. Comm. 2008, 28, 3293–3295. [Google Scholar] [CrossRef] [PubMed]
  219. Bao, Y.; Wei, P.; Xia, X.; Huang, Z.; Homewood, K.; Gao, Y. Remarkably enhanced H2 response and detection range in Nb doped rutile/anatase heterophase junction TiO2 thin film hydrogen sensors. Sens. Actuators B Chem. 2019, 301, 127143. [Google Scholar] [CrossRef]
  220. He, J.; Du, Y.; Bai, Y.; An, J.; Cai, X.; Chen, Y.; Wang, P.; Yang, X.; Feng, Q. Facile formation of anatase/rutile TiO2 nanocomposites with enhanced photocatalytic activity. Molecules 2019, 24, 2996. [Google Scholar] [CrossRef] [PubMed]
  221. Lin, X.; Sun, M.; Gao, B.; Ding, W.; Zhang, Z.; Anandan, S.; Umar, A. Hydrothermally regulating phase composition of TiO2 nanocrystals toward high photocatalytic activity. J. Alloys Compd. 2021, 850, 156653. [Google Scholar] [CrossRef]
  222. Mohamed, M.A.; Jaafar, J.; Zain, M.F.M.; Minggu, L.J.; Kassim, M.B.; Salehmin, M.N.I.; Rosmi, M.S.; Salleh, W.N.; Othman, M.H.D. Concurrent growth, structural and photocatalytic properties of hybridized C, N co-doped TiO2 mixed phase over g-C3N4 nanostructured. Scr. Mater. 2018, 142, 143–147. [Google Scholar] [CrossRef]
  223. Li, J.; Xu, X.; Liu, X.; Qin, W.; Wang, M.; Pan, L. Metal-organic frameworks derived cake-like anatase/rutile mixed phase TiO2 for highly efficient photocatalysis. J. Alloys Compd. 2017, 690, 640–646. [Google Scholar] [CrossRef]
  224. Peng, F.; Gao, H.; Zhang, G.; Zhu, Z.; Zhang, J.; Liu, Q. Synergistic effects of Sm and C co-doped mixed phase crystalline TiO2 for visible light photocatalytic activity. Materials 2017, 10, 209. [Google Scholar] [CrossRef]
  225. Jacob, K.A.; Peter, P.M.; Jose, P.E.; Balakrishnan, C.J.; Thomas, V.J. A simple method for the synthesis of anatase-rutile mixed phase TiO2 using a convenient precursor and higher visible-light photocatalytic activity of Co–doped TiO2. Mater. Today Proc. 2021, 49, 1408–1417. [Google Scholar] [CrossRef]
  226. Li, H.; Shen, X.; Liu, Y.; Wang, L.; Lei, J.; Zhang, J. Facile phase control for hydrothermal synthesis of anatase-rutile TiO2 with enhanced photocatalytic activity. J. Alloys Compd. 2015, 646, 380–386. [Google Scholar] [CrossRef]
  227. Wang, Q.; Qiao, Z.; Jiang, P.; Kuang, J.; Liu, W.; Cao, W. Hydrothermal synthesis and enhanced photocatalytic activity of mixed-phase TiO2 powders with controllable anatase/rutile ratio. Solid State Sci. 2018, 77, 14–19. [Google Scholar] [CrossRef]
  228. Yang, Z.; Wang, B.; Cui, H.; An, H.; Pan, Y.; Zhai, J. Synthesis of crystal-controlled TiO2 nanorods by a hydrothermal method: Rutile and brookite as highly active photocatalysts. J. Phys. Chem. C 2015, 119, 16905–16912. [Google Scholar] [CrossRef]
  229. Luthfiah, A.; Permana, M.D.; Deawati, Y.; Firdaus, M.L.; Rahayu, I.; Eddy, D.R. Photocatalysis of nanocomposite titania–natural silica as antibacterial against Staphylococcus aureus and Pseudomonas aeruginosa. RSC Adv. 2021, 11, 38528–38536. [Google Scholar] [CrossRef] [PubMed]
  230. Duvarci, Ö.Ç.; Çiftçioğlu, M. Preparation and characterization of nanocrystalline titania powders by sonochemical synthesis. Powder Technol. 2012, 228, 231–240. [Google Scholar] [CrossRef]
  231. Garibay-Febles, V.; Hernández-Pérez, I.; Arceo, L.D.B.; Meza-Espinoza, J.S.; Espinoza-Tapia, J.C.; González-Reyes, L. Microstructural study by electron microscopy of sonochemical synthesized TiO2 nanoparticles. Acta Microsc. 2017, 26, 56–64. [Google Scholar]
  232. Hernández-Perez, I.; Maubert, A.M.; Rendón, L.; Santiago, P.; Herrera-Hernández, H.; Díaz-Barriga Arceo, L.; Garibay Febles, V.; Palacios Gonzalez, E.; González-Reyes, L. Ultrasonic synthesis: Structural, optical and electrical correlation of TiO2 nanoparticles. Int. J. Electrochem. Sci. 2012, 7, 8832–8847. [Google Scholar]
  233. Noman, M.T.; Militky, J.; Wiener, J.; Saskova, J.; Ashraf, M.A.; Jamshaid, H.; Azeem, M. Sonochemical synthesis of highly crystalline photocatalyst for industrial applications. Ultrasonics 2018, 83, 203–213. [Google Scholar] [CrossRef]
  234. Arami, H.; Mazloumi, M.; Khalifehzadeh, R.; Sadrnezhaad, S.K. Sonochemical preparation of TiO2 nanoparticles. Mater. Lett. 2007, 61, 4559–4561. [Google Scholar] [CrossRef]
  235. Ibrahim, A.; Mekprasart, W.; Pecharapa, W. Anatase/Rutile TiO2 composite prepared via sonochemical process and their photocatalytic activity. Mater. Today: Proc. 2017, 4, 6159–6165. [Google Scholar]
  236. Yu, J.C.; Zhang, L.; Yu, J. Direct sonochemical preparation and characterization of highly active mesoporous TiO2 with a bicrystalline framework. Chem. Mater. 2002, 14, 4647–4653. [Google Scholar] [CrossRef]
  237. Ozawa, T.; Iwasaki, M.; Tada, H.; Akita, T.; Tanaka, K.; Ito, S. Low-temperature synthesis of anatase–brookite composite nanocrystals: The junction effect on photocatalytic activity. J. Colloid Interface Sci. 2005, 281, 510–513. [Google Scholar] [CrossRef]
  238. Matthews, A. The crystallization of anatase and rutile from amorphous titanium dioxide under hydrothermal conditions. Am. Mineral. 1976, 61, 419–424. [Google Scholar]
  239. Hanaor, D.A.; Sorrell, C.C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2011, 46, 855–874. [Google Scholar] [CrossRef]
  240. Sikong, L.; Damchan, J.; Kooptarnond, K.; Niyomwas, S. Effect of doped SiO2 and calcinations temperature on phase transformation of TiO2 photocatalyst prepared by sol-gel method. Songklanakarin J. Sci. Technol. 2008, 30, 385–391. [Google Scholar]
  241. Pillai, S.C.; Periyat, P.; George, R.; McCormack, D.E.; Seery, M.K.; Hayden, H.; Colreavy, J.; Corr, D.; Hinder, S.J. Synthesis of High-Temperature Stable Anatase TiO2 Photocatalyst. J. Phys. Chem. C 2007, 111, 1605–1611. [Google Scholar] [CrossRef]
  242. Yuangpho, N.; Le, S.T.T.; Treerujiraphapong, T.; Khanitchaidecha, W.; Nakaruk, A. Enhanced photocatalytic performance of TiO2 particles via effect of anatase–rutile ratio. Physica E Low Dimens. Syst. Nanostruct. 2015, 67, 18–22. [Google Scholar] [CrossRef]
  243. Scarpelli, F.; Mastropietro, T.F.; Poerio, T.; Godbert, N. Mesoporous TiO2 thin films: State of the art. In Titanium Dioxide-Material for a Sustainable Environment; Yang, D., Ed.; IntechOpen: London, UK, 2018; Volume 508, pp. 135–142. [Google Scholar]
  244. Liu, L.; Zhao, H.; Andino, J.M.; Li, Y. Photocatalytic CO2 reduction with H2O on TiO2 nanocrystals: Comparison of anatase, rutile, and brookite polymorphs and exploration of surface chemistry. ACS Catal. 2012, 2, 1817–1828. [Google Scholar] [CrossRef]
  245. Hurum, D.C.; Agrios, A.G.; Gray, K.A.; Rajh, T.; Thurnauer, M.C. Explaining the enhanced photocatalytic activity of Degussa P25 mixed-phase TiO2 using EPR. J. Phys. Chem. B 2003, 107, 4545–4549. [Google Scholar] [CrossRef]
  246. Mancinelli, R.; Botti, A.; Bruni, F.; Ricci, M.A.; Soper, A.K. Hydration of sodium, potassium, and chloride ions in solution and the concept of structure maker/breaker. J. Phys. Chem. B 2007, 111, 13570–13577. [Google Scholar] [CrossRef]
  247. Zaw, Y.Y.; Channei, D.A.D.; Threrujirapapong, T.; Khanitchaidecha, W.; Nakaruk, A. Effect of anatase/rutile phase ratio on the photodegradation of methylene blue under uv irradiation. Mater. Sci. Forum 2020, 998, 78–83. [Google Scholar] [CrossRef]
  248. Wetchakun, N.; Phanichphant, S. Effect of temperature on the degree of anatase–rutile transformation in titanium dioxide nanoparticles synthesized by the modified sol–gel method. Curr. Appl. Phys. 2008, 8, 343–346. [Google Scholar] [CrossRef]
  249. Liu, Y.; Zou, X.; Li, L.; Shen, Z.; Cao, Y.; Wang, Y.; Cui, L.; Cheng, J.; Wang, Y.; Li, X. Engineering of anatase/rutile TiO2 heterophase junction via in-situ phase transformation for enhanced photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2021, 599, 795–804. [Google Scholar] [CrossRef] [PubMed]
  250. Bakardjieva, S.; Šubrt, J.; Štengl, V.; Dianez, M.J.; Sayagues, M.J. Photoactivity of anatase–rutile TiO2 nanocrystalline mixtures obtained by heat treatment of homogeneously precipitated anatase. Appl. Catal. B 2005, 58, 193–202. [Google Scholar] [CrossRef]
  251. Yaemsunthorn, K.; Kobielusz, M.; Macyk, W. TiO2 with tunable anatase-to-rutile nanoparticles ratios: How does the photoactivity depend on the phase composition and the nature of photocatalytic reaction? ACS Appl. Nano Mater. 2021, 4, 633–643. [Google Scholar] [CrossRef]
  252. Xia, X.; Peng, S.; Bao, Y.; Wang, Y.; Lei, B.; Wang, Z.; Huang, Z.; Gao, Y. Control of interface between anatase TiO2 nanoparticles and rutile TiO2 nanorods for efficient photocatalytic H2 generation. J. Power Sources 2018, 376, 11–17. [Google Scholar] [CrossRef]
  253. Lei, Y.; Yang, Y.; Zhang, P.; Zhou, J.; Wu, J.; Li, K.; Wang, W.; Chen, L. Controllable one-step synthesis of mixed-phase TiO2 nanocrystals with equivalent anatase/rutile ratio for enhanced photocatalytic performance. Nanomaterials 2021, 11, 1347. [Google Scholar] [CrossRef]
  254. Bernardini, C.; Cappelletti, G.; Dozzi, M.V.; Selli, E. Photocatalytic degradation of organic molecules in water: Photoactivity and reaction paths in relation to TiO2 particles features. J. Photochem. Photobiol. A: Chem. 2010, 211, 185–192. [Google Scholar] [CrossRef]
  255. Bojinova, A.; Kralchevska, R.; Poulios, I.; Dushkin, C. Anatase/rutile TiO2 composites: Influence of the mixing ratio on the photocatalytic degradation of Malachite Green and Orange II in slurry. Mater. Chem. Phys. 2007, 106, 187–192. [Google Scholar] [CrossRef]
  256. Cong, S.; Xu, Y. Explaining the high photocatalytic activity of a mixed phase TiO2: A combined effect of O2 and crystallinity. J. Phys. Chem. C 2011, 115, 21161–21168. [Google Scholar] [CrossRef]
  257. Ohno, T.; Sarukawa, K.; Matsumura, M. Crystal faces of rutile and anatase TiO2 particles and their roles in photocatalytic reactions. New J. Chem. 2002, 26, 1167–1170. [Google Scholar] [CrossRef]
  258. Zhang, Y.; Chen, J.; Li, X. Preparation and photocatalytic performance of anatase/rutile mixed-phase TiO2 nanotubes. Catal. Lett. 2010, 139, 129–133. [Google Scholar] [CrossRef]
  259. Almashhori, K.; Ali, T.T.; Saeed, A.; Alwafi, R.; Aly, M.; Al-Hazmi, F.E. Antibacterial and photocatalytic activities of controllable (anatase/rutile) mixed phase TiO2 nanophotocatalysts synthesized via a microwave-assisted sol–gel method. New J. Chem. 2020, 44, 562–570. [Google Scholar] [CrossRef]
  260. Xiong, J.; Zhang, M.; Cheng, G. Facile polyol-triggered anatase–rutile heterophase TiO2-x nanoparticles for enhancing photocatalytic CO2 reduction. J. Colloid Interface Sci. 2020, 579, 872–877. [Google Scholar] [CrossRef] [PubMed]
  261. Lin, H.; Li, L.; Zhao, M.; Huang, X.; Chen, X.; Li, G.; Yu, R. Synthesis of high-quality brookite TiO2 single-crystalline nanosheets with specific facets exposed: Tuning catalysts from inert to highly reactive. J. Am. Chem. Soc. 2012, 134, 8328–8331. [Google Scholar] [CrossRef] [PubMed]
  262. Khedr, T.M.; El-Sheikh, S.M.; Kowalska, E.; Abdeldayem, H.M. The synergistic effect of anatase and brookite for photocatalytic generation of hydrogen and diclofenac degradation. J. Environ. Chem. Eng. 2021, 9, 106566. [Google Scholar] [CrossRef]
  263. Kandiel, T.A.; Feldhoff, A.; Robben, L.; Dillert, R.; Bahnemann, D.W. Tailored titanium dioxide nanomaterials: Anatase nanoparticles and brookite nanorods as highly active photocatalysts. Chem. Mater. 2010, 22, 2050–2060. [Google Scholar] [CrossRef]
  264. Shen, X.; Zhang, J.; Tian, B.; Anpo, M. Tartaric acid-assisted preparation and photocatalytic performance of titania nanoparticles with controllable phases of anatase and brookite. J. Mater. Sci. 2012, 47, 5743–5751. [Google Scholar] [CrossRef]
  265. Shen, X.; Tian, B.; Zhang, J. Tailored preparation of titania with controllable phases of anatase and brookite by an alkalescent hydrothermal route. Catal. Today 2013, 201, 151–158. [Google Scholar] [CrossRef]
  266. Tay, Q.; Liu, X.; Tang, Y.; Jiang, Z.; Sum, T.C.; Chen, Z. Enhanced photocatalytic hydrogen production with synergistic two-phase anatase/brookite TiO2 nanostructures. J. Phys. Chem. C 2013, 117, 14973–14982. [Google Scholar] [CrossRef]
  267. El-Sheikh, S.M.; Khedr, T.M.; Zhang, G.; Vogiazi, V.; Ismail, A.A.; O’Shea, K.; Dionysiou, D.D. Tailored synthesis of anatase–brookite heterojunction photocatalysts for degradation of cylindrospermopsin under UV–Vis light. Chem. Eng. J. 2017, 310, 428–436. [Google Scholar] [CrossRef]
  268. Mutuma, B.K.; Shao, G.N.; Kim, W.D.; Kim, H.T. Sol–gel synthesis of mesoporous anatase–brookite and anatase–brookite–rutile TiO2 nanoparticles and their photocatalytic properties. J. Colloid Interface Sci. 2015, 442, 1–7. [Google Scholar] [CrossRef] [PubMed]
  269. Andrade-Guel, M.; Díaz-Jiménez, L.; Cortés-Hernández, D.; Cabello-Alvarado, C.; Ávila-Orta, C.; Bartolo-Pérez, P.; Gamero-Melo, P. Microwave assisted sol–gel synthesis of titanium dioxide using hydrochloric and acetic acid as catalysts. Bol. Soc. Esp. Ceram. Vidr. 2019, 58, 171–177. [Google Scholar] [CrossRef]
  270. Quintero, Y.; Mosquera, E.; Diosa, J.; García, A. Ultrasonic-assisted sol–gel synthesis of TiO2 nanostructures: Influence of synthesis parameters on morphology, crystallinity, and photocatalytic performance. J. Sol-Gel Sci. Technol. 2020, 94, 477–485. [Google Scholar] [CrossRef]
  271. Cihlar, J.; Kasparek, V.; Kralova, M.; Castkova, K. Biphasic anatase-brookite nanoparticles prepared by sol–gel complex synthesis and their photocatalytic activity in hydrogen production. Int. J. Hydrogen Energy 2015, 40, 2950–2962. [Google Scholar] [CrossRef]
  272. Cihlar, J.; Navarro, L.K.T.; Kasparek, V.; Michalicka, J.; Cihlar Jr, J.; Kastyl, J.; Castkova, K.; Celko, L. Influence of LA/Ti molar ratio on the complex synthesis of anatase/brookite nanoparticles and their hydrogen production. Int. J. Hydrogen Energy 2021, 46, 8578–8593. [Google Scholar] [CrossRef]
  273. Di Paola, A.; Cufalo, G.; Addamo, M.; Bellardita, M.; Campostrini, R.; Ischia, M.; Ceccato, R.; Palmisano, L. Photocatalytic activity of nanocrystalline TiO2 (brookite, rutile and brookite-based) powders prepared by thermohydrolysis of TiCl4 in aqueous chloride solutions. Colloids Surf. A Physicochem. Eng. Asp. 2008, 317, 366–376. [Google Scholar] [CrossRef]
  274. Luo, B.; Li, Z.; Xu, Y. The positive effect of anatase and rutile on the brookite-photocatalyzed degradation of phenol. RSC Adv. 2015, 5, 105999–106004. [Google Scholar] [CrossRef]
  275. Xu, H.; Zhang, L. Controllable one-pot synthesis and enhanced photocatalytic activity of mixed-phase TiO2 nanocrystals with tunable brookite/rutile ratios. J. Phys. Chem. C 2009, 113, 1785–1790. [Google Scholar] [CrossRef]
  276. Cao, Y.; Li, X.; Bian, Z.; Fuhr, A.; Zhang, D.; Zhu, J. Highly photocatalytic activity of brookite/rutile TiO2 nanocrystals with semi-embedded structure. Appl. Catal. B 2016, 180, 551–558. [Google Scholar] [CrossRef]
  277. Kim, M.G.; Lee, J.E.; Kim, K.S.; Kang, J.M.; Lee, J.H.; Kim, K.H.; Cho, M.; Lee, S.G. Photocatalytic degradation of methylene blue under UV and visible light by brookite–rutile bi-crystalline phase of TiO2. New J. Chem. 2021, 45, 3485–3497. [Google Scholar] [CrossRef]
  278. Qi, L.; Liu, Y.; Li, C. Controlled synthesis of TiO2-B nanowires and nanoparticles for dye-sensitized solar cells. Appl. Surf. Sci. 2010, 257, 1660–1665. [Google Scholar] [CrossRef]
  279. Wang, G.; Wang, Q.; Lu, W.; Li, J. Photoelectrochemical study on charge transfer properties of TiO2−B nanowires with an application as humidity sensors. J. Phys. Chem. B 2006, 110, 22029–22034. [Google Scholar] [CrossRef] [PubMed]
  280. Feist, T.P.; Davies, P.K. The soft chemical synthesis of TiO2 (B) from layered titanates. J. Solid State Chem. 1992, 101, 275–295. [Google Scholar] [CrossRef]
  281. Yamamoto, K.; Tomita, K.; Fujita, K.; Kobayashi, M.; Petrykin, V.; Kakihana, M. Synthesis of TiO2 (B) using glycolato titanium complex and post-synthetic hydrothermal crystal growth of TiO2 (B). J. Cryst. Growth 2009, 311, 619–622. [Google Scholar] [CrossRef]
  282. Kobayashi, M.; Petrykin, V.V.; Kakihana, M.; Tomita, K.; Yoshimura, M. One-step synthesis of TiO2 (B) nanoparticles from a water-soluble titanium complex. Chem. Mater. 2007, 19, 5373–5376. [Google Scholar] [CrossRef]
  283. Giannuzzi, R.; Manca, M.; De Marco, L.; Belviso, M.R.; Cannavale, A.; Sibillano, T.; Giannini, C.; Cozzoli, P.D.; Gigli, G. Ultrathin TiO2 (B) nanorods with superior lithium-ion storage performance. ACS Appl. Mater. Interfaces 2014, 6, 1933–1943. [Google Scholar] [CrossRef]
  284. Cai, J.; Wang, Y.; Zhu, Y.; Wu, M.; Zhang, H.; Li, X.; Jiang, Z.; Meng, M. In situ formation of disorder-engineered TiO2 (B)-anatase heterophase junction for enhanced photocatalytic hydrogen evolution. ACS Appl. Mater. Interfaces 2015, 7, 24987–24992. [Google Scholar] [CrossRef]
  285. Mikrut, P.; Kobielusz, M.; Indyka, P.; Macyk, W. Photocatalytic activity of TiO2 polymorph B revisited: Physical, redox, spectroscopic, and photochemical properties of TiO2 (B)/anatase series of titanium dioxide materials. Mater. Today Sustain. 2020, 10, 100052. [Google Scholar] [CrossRef]
  286. Zhu, J.; Zhu, S.; Kong, X.; Liang, Y.; Li, Z.; Wu, S.; Luo, S.; Chang, C.; Cui, Z. Rutile-coated B-phase TiO2 heterojunction nanobelts for photocatalytic H2 evolution. ACS Appl. Nano Mater. 2020, 3, 10349–10359. [Google Scholar] [CrossRef]
  287. Chalastara, K.; Guo, F.; Elouatik, S.; Demopoulos, G.P. Tunable composition aqueous-synthesized mixed-phase TiO2 nanocrystals for photo-assisted water decontamination: Comparison of anatase, brookite and rutile photocatalysts. Catalysts 2020, 10, 407. [Google Scholar] [CrossRef]
  288. Allen, N.S.; Mahdjoub, N.; Vishnyakov, V.; Kelly, P.J.; Kriek, R.J. The effect of crystalline phase (anatase, brookite and rutile) and size on the photocatalytic activity of calcined polymorphic titanium dioxide (TiO2). Polym. Degrad. Stab. 2018, 150, 31–36. [Google Scholar] [CrossRef]
  289. Kaplan, R.; Erjavec, B.; Dražić, G.; Grdadolnik, J.; Pintar, A. Simple synthesis of anatase/rutile/brookite TiO2 nanocomposite with superior mineralization potential for photocatalytic degradation of water pollutants. Appl. Catal. B 2016, 181, 465–474. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of TiO2 photocatalytic principle.
Figure 1. Schematic diagram of TiO2 photocatalytic principle.
Nanomaterials 13 00704 g001
Figure 2. Crystal structure of TiO2 anatase, rutile, brookite, and TiO2 (B).
Figure 2. Crystal structure of TiO2 anatase, rutile, brookite, and TiO2 (B).
Nanomaterials 13 00704 g002
Figure 3. Bandgap energies, VB, and CB for anatase, rutile, and brookite on the potential scale (V) versus the normal hydrogen electrode (NHE).
Figure 3. Bandgap energies, VB, and CB for anatase, rutile, and brookite on the potential scale (V) versus the normal hydrogen electrode (NHE).
Nanomaterials 13 00704 g003
Figure 4. Mechanism of band gap narrowing for peroxo-titania (anatase) complex. Reproduced with permission from [140]. Copyright John Wiley and Sons, 2011.
Figure 4. Mechanism of band gap narrowing for peroxo-titania (anatase) complex. Reproduced with permission from [140]. Copyright John Wiley and Sons, 2011.
Nanomaterials 13 00704 g004
Figure 5. Limitations of TiO2 as a photocatalyst material.
Figure 5. Limitations of TiO2 as a photocatalyst material.
Nanomaterials 13 00704 g005
Figure 6. Mechanism of photocatalysis in anatase, rutile, and anatase–rutile heterophase.
Figure 6. Mechanism of photocatalysis in anatase, rutile, and anatase–rutile heterophase.
Nanomaterials 13 00704 g006
Figure 7. Heterojunction of mixed-phase TiO2 nanoparticles: (a) alignment of the conduction band and valence band between the rutile and anatase phases, which causes the separation of electron holes at the heterojunction; and (b) heterojunction density variations during the transformation of the anatase phase to rutile.
Figure 7. Heterojunction of mixed-phase TiO2 nanoparticles: (a) alignment of the conduction band and valence band between the rutile and anatase phases, which causes the separation of electron holes at the heterojunction; and (b) heterojunction density variations during the transformation of the anatase phase to rutile.
Nanomaterials 13 00704 g007
Figure 8. Interface formation between rutile (110) and anatase (101) before and after the cooling steps. Reproduced with permission from [206]. Copyright American Chemical Society, 2007.
Figure 8. Interface formation between rutile (110) and anatase (101) before and after the cooling steps. Reproduced with permission from [206]. Copyright American Chemical Society, 2007.
Nanomaterials 13 00704 g008
Figure 9. Absorption spectra of anatase (A), rutile I, multilayer films before (A/Am/R) and after (A/R) heat treatment at 500 °C for 10 h. Inset: Tauc plots showing approximate bandgap energy values of the A,R crystalline phases. Reproduced with permission from [209]. Copyright John Wiley and Sons, 2014.
Figure 9. Absorption spectra of anatase (A), rutile I, multilayer films before (A/Am/R) and after (A/R) heat treatment at 500 °C for 10 h. Inset: Tauc plots showing approximate bandgap energy values of the A,R crystalline phases. Reproduced with permission from [209]. Copyright John Wiley and Sons, 2014.
Nanomaterials 13 00704 g009
Figure 10. Schematic of sol–gel method to produce anatase/rutile heterophase.
Figure 10. Schematic of sol–gel method to produce anatase/rutile heterophase.
Nanomaterials 13 00704 g010
Figure 11. Schematic of hydrothermal method to produce anatase/rutile heterophase.
Figure 11. Schematic of hydrothermal method to produce anatase/rutile heterophase.
Nanomaterials 13 00704 g011
Figure 12. Schematic of sonochemical method to produce anatase/rutile heterophase.
Figure 12. Schematic of sonochemical method to produce anatase/rutile heterophase.
Nanomaterials 13 00704 g012
Figure 13. Schematic diagram illustrating the charge transfer across the TiO2 (B)/anatase heterophase junction.
Figure 13. Schematic diagram illustrating the charge transfer across the TiO2 (B)/anatase heterophase junction.
Nanomaterials 13 00704 g013
Figure 14. Possible connection model of the mixed phase at the phase interface.
Figure 14. Possible connection model of the mixed phase at the phase interface.
Nanomaterials 13 00704 g014
Table 1. Crystal structure and physical parameters of TiO2 polymorphs.
Table 1. Crystal structure and physical parameters of TiO2 polymorphs.
PhaseCrystal System, Space GroupLattice ParameterDensity (g/cm3)Band Gap Energy (eV)Ref.
BrookiteOrtorombik, Pbcaa = 9.148 Å4.123.14–3.31[117]
b = 5.447 Å
c = 5.145 Å
V = 257.38 Å3
RutileTetragonal, P42/mnma = b = 4.594 Å4.253.02–3.04[117]
c = 2.959 Å
V = 62.45 Å3
AnataseTetragonal, I41/amda = b = 3.784 Å3.893.20–3.23[117]
c = 9.515 Å
V = 136.24 Å3
TiO2 (B)Monoclinic, C2/ma = 12.179 Å3.733.09–3.22[113]
b = 3.741 Å
c = 6.525 Å
β = 107.054
V = 284.22 Å3
Table 2. Various methods for synthesis TiO2 anatase/rutile.
Table 2. Various methods for synthesis TiO2 anatase/rutile.
PrecursorMethodPhase ControllerApplicationRef.
TiO2 P25HydrothermalCalcination temperatureWater splitting[63]
TTIPHydrothermal Calcination temperatureMethylene blue degradation[247]
TTIP Sol–gelCalcination temperatureOxalic acid degradation[248]
TBOTHydrothermalpHHydrogen generation[252]
TiOSO4Urea precipitation methodCalcination temperature4-chlorophenol degradation [250]
TiCl4One step condensing refluxSolvent Rhodamine B degradation[253]
TiCl3HydrothermalCalcination temperatureHydrogen generation [251]
Table 3. Photocatalytic activity of heterophase anatase/rutile.
Table 3. Photocatalytic activity of heterophase anatase/rutile.
PollutantOptimum
Composition (%) *
Dye
Concentration
Light SourceIrradiation Time
(min)
Efficiency
(%)
Ref.
Methylene blueA 78.5
R 21.5
1 × 10−5 MUV lamp 20~94%[248]
Rhodamine BA 48.0
R 52.0
4.8 mg/L300 W Xenon lamp
(365 nm)
6099%[253]
Methyl orangeA 70.0
R 30.0
20 mg/L125 W Mercury lamp (365 nm)40 90.6%[258]
Methylene blueA 30.8
R 69.2
2.92 × 10−5 M175 W UV lamp9085.8%[220]
Crystal violetA 81.6
R 18.4
100 ppmUV lamp (365 nm)30092.65%[259]
Methylene blueA 81.6
R 18.4
100 ppmUV lamp (365 nm)30094.77%[259]
* A: Anatase, R: Rutile.
Table 4. Various methods for synthesis TiO2 anatase/brookite.
Table 4. Various methods for synthesis TiO2 anatase/brookite.
PrecursorMethodAdditiveApplicationRef.
TALHHydrothermalUreaHydrogen production and dichloroacetic acid (DCA) degradation[263]
TiCl3HydrothermalTartaric acidRhodamine B degradation[264]
TiCl3HydrothermalNaCl and
NH4OH
Rhodamine B degradation[265]
TiS2HydrothermalNaOH Hydrogen generation[266]
Ti2(SO4)3HydrothermalGlycine with NH4OH or NaOHCylindrospermopsin
degradation
[267]
TTIPSol–gelHNO3Methylene blue degradation[268]
TTIPMicrowave assisted sol–gelHCl and CH3COOH-[269]
TTIPUltrasonic-assisted sol–gelHNO3 and CH3COOHMethylene blue degradation[270]
TTIPSol–gel complexhydroxycarboxylic acids Hydrogen generation[271]
Ti(Opr)4Sol–gel complexLactic acidHydrogen generation[272]
Ti2(SO4)3HydrothermalGlycine with NH4OH or NaOHDiclofenac degradation[273]
Table 5. Photocatalytic activity of heterophase anatase/brookite.
Table 5. Photocatalytic activity of heterophase anatase/brookite.
PollutantOptimum
Composition (%) *
Photocatalyst LoadingDye
Concentration
Light SourceIrradiation Time
(min)
Efficiency
(%)
Ref.
Rhodamine BA 78.7
B 21.3
-20 mg/L300 W Hg lamp (365 nm)12098[264]
Rhodamine BA 46.6
B 53.4
-20 mg/L300 W Hg lamp (365 nm)10095[265]
Methylene blueA 80.0
B 20.0
0.6 g/L32 mg/L100 W mercury lamp7098[268]
Methylene blueA 79.0
B 21.0
1 g/L5 mg/L100 W UV lamp (365 nm)24060[270]
Cylindrospermopsin (CYN)A 61.8
B 38.2
0.25 g/L1 × 10−6 M15 W florescence lamp (310–720 nm)15100[267]
Diclofenac (DCF)A 61.8
B 38.2
1 g/L10 mg/LUV-A
irradiation
120100[262]
* A: Anatase, B: Brookite.
Table 6. Photocatalytic activity of heterophase TiO2 for hydrogen production.
Table 6. Photocatalytic activity of heterophase TiO2 for hydrogen production.
Optimum
Composition (%) *
Amount of Catalyst Light Source H2 Evolution Rate Ref.
A 73.5
R 26.5
20 mg/100 mL 300 W Xe lamp584 μmol g−1 h−1 [63]
A 12.0
R 88.0
20 mg/80 mL 350 W Xe lamp74,400 μmol g−1 h−1 [252]
A 72.0
B 28.0
37.5 mg/75 mL 1000 W Xe lamp4267 μmol g−1 h−1 [263]
A 88.4
B 11.6
45 mg/100 mL 800 W Xe−Hg lamp~3500 μmol g−1 h−1 [266]
A 74.0
B 36.0
20 mg/100 mL 450 W Xe lamp101.4 μmol g−1 min−1 [271]
A 54.0
B 46.0
20 mg/100 mL 450 W Xe lamp101.4 μmol g−1 min−1 [272]
* A: anatase, R: rutile, B: brookite.
Table 7. Various methods for synthesis TiO2 (B).
Table 7. Various methods for synthesis TiO2 (B).
PrecursorMethodAdditiveApplicationRef.
Titanium metal powderPost-synthetic Hydrothermal crystal growthH2SO4-[281]
Titanium metal powderHydrothermalH2SO4-[282]
TiCl4Solvothermal-Lithium-ion batteries [173]
TiO2 P25Hydrothermal -DSSC[278]
TiO2 P25Hydrothermal -Humidity sensors[279]
TTIPSurfactant-assisted nonaqueous sol–gel routeOleic acidLithium ion-batteries[283]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Eddy, D.R.; Permana, M.D.; Sakti, L.K.; Sheha, G.A.N.; Solihudin; Hidayat, S.; Takei, T.; Kumada, N.; Rahayu, I. Heterophase Polymorph of TiO2 (Anatase, Rutile, Brookite, TiO2 (B)) for Efficient Photocatalyst: Fabrication and Activity. Nanomaterials 2023, 13, 704. https://doi.org/10.3390/nano13040704

AMA Style

Eddy DR, Permana MD, Sakti LK, Sheha GAN, Solihudin, Hidayat S, Takei T, Kumada N, Rahayu I. Heterophase Polymorph of TiO2 (Anatase, Rutile, Brookite, TiO2 (B)) for Efficient Photocatalyst: Fabrication and Activity. Nanomaterials. 2023; 13(4):704. https://doi.org/10.3390/nano13040704

Chicago/Turabian Style

Eddy, Diana Rakhmawaty, Muhamad Diki Permana, Lintang Kumoro Sakti, Geometry Amal Nur Sheha, Solihudin, Sahrul Hidayat, Takahiro Takei, Nobuhiro Kumada, and Iman Rahayu. 2023. "Heterophase Polymorph of TiO2 (Anatase, Rutile, Brookite, TiO2 (B)) for Efficient Photocatalyst: Fabrication and Activity" Nanomaterials 13, no. 4: 704. https://doi.org/10.3390/nano13040704

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