Next Article in Journal
An Eco-Friendly Fluidizable FexOy/CaO-γ-Al2O3 Catalyst for Tar Cracking during Biomass Gasification
Next Article in Special Issue
Treatment of Produced Water with Photocatalysis: Recent Advances, Affecting Factors and Future Research Prospects
Previous Article in Journal
The Influence of a Surface Treatment of Metallic Titanium on the Photocatalytic Properties of TiO2 Nanotubes Grown by Anodic Oxidation
Previous Article in Special Issue
Photocatalytic Degradation of Atenolol by TiO2 Irradiated with an Ultraviolet Light Emitting Diode: Performance, Kinetics, and Mechanism Insights
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Impact of Titanium Dioxide (TiO2) Modification on Its Application to Pollution Treatment—A Review

MOE Key Laboratory of Pollution Processes and Environmental Criteria/Tianjin Key Laboratory of Environmental Remediation and Pollution Control/College of Environmental Science and Engineering, Nankai University, No. 38 Tongyan Road, Jinnan District, Tianjin 300350, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(7), 804; https://doi.org/10.3390/catal10070804
Submission received: 12 July 2020 / Accepted: 17 July 2020 / Published: 20 July 2020
(This article belongs to the Special Issue Advances in Photocatalytic Wastewater Purification)

Abstract

:
A high-efficiency method to deal with pollutants must be found because environmental problems are becoming more serious. Photocatalytic oxidation technology as the environmentally-friendly treatment method can completely oxidate organic pollutants into pollution-free small-molecule inorganic substances without causing secondary pollution. As a widely used photocatalyst, titanium dioxide (TiO2) can greatly improve the degradation efficiency of pollutants, but several problems are noted in its practical application. TiO2 modified by different materials has received extensive attention in the field of photocatalysis because of its excellent physical and chemical properties compared with pure TiO2. In this review, we discuss the use of different materials for TiO2 modification, highlighting recent developments in the synthesis and application of TiO2 composites using different materials. Materials discussed in the article can be divided into nonmetallic and metallic. Mechanisms of how to improve catalytic performance of TiO2 after modification are discussed, and the future development of modified TiO2 is prospected.

1. Introduction

In recent years, photocatalytic technology has attracted extensive attention because it is environmentally friendly, low-cost, and has efficient characteristics. In 1972, Fujishima et al. [1] reported that TiO2 was used as a photoelectrocatalyst to split water into hydrogen. Since then, increasing research has focused on TiO2. In 1976, Carey et al. [2] used photocatalytic technology to treat polychlorinated biphenyls, an organic pollution that is difficult to degrade, and experimental results found that the dechlorination rate of polychlorinated biphenyls was close to 100%. In 1977, Frank et al. [3] found that TiO2 could effectively degrade cyanide (CN), which was the beginning of photocatalytic technology applied to pollution control. The degradation of photocatalytic technology can be summarized into four stages: photoexcitation, carrier capture, formation of radicals, and oxidation reaction. Compared with traditional catalytic technologies, photocatalytic technology has many advantages. First, reaction conditions such as sunlight, room temperature, and normal atmospheric pressure are common and easy to obtain. Second, the degradation processes and products of catalytic decomposition are pollution free, which are in line with the requirements of low-carbon environmental protection. Third, the characteristics of non-toxic, stable, low cost, and recyclable further promote development [4]. The core of photocatalytic technology is the photocatalyst, and many materials can act as a photocatalyst [5]. Table 1 shows the published data of different photocatalysts, including TiO2 [6,7,8], SrTiO3 [9,10,11], ZnO [12,13,14], WO3 [15,16,17], ZrO2 [18,19,20], and g-C3N4 [21,22,23], and their performance. Among these photocatalysts, TiO2 occupies an important position due to its stable physical and chemical properties, strong oxidation capacity, high photocatalytic activity, and excellent biocompatibility [24,25,26].
The TiO2 can be synthesized by many methods and mainly include precipitation method, solvothermal method, sol-gel method, microemulsion method, spray pyrolysis method and electrochemical synthesis method. Zhang et al. [27] prepared the TiO2/BC catalyst material by the sol-gel method and the degradation rates of reactive brilliant blue KN-R in the dyeing wastewater can reach 97%. However, there are still has some problems of pure TiO2 in application. The rapid recombination of photo-generated electron-hole pairs is the biggest obstacle that affect the practical application of TiO2 [28], because recombination of photogenerated charge carriers can reduce the overall quantum efficiency [29]. The poor photosensitivity of TiO2 under visible/solar irradiation is also a problem [30,31]. Generally, the conventional TiO2 has broad intrinsic band gaps wide band gap (3.2 eV for anatase and 3.0 eV for rutile) which makes the TiO2 only absorb UV radiation (wavelength < 400 nm), which accounts for only ∼5% of the sunlight [32,33]. What is more, nano-TiO2 is easy to agglomerate, which extremely limits the application. Therefore, in order to solve these problems and improve the catalytic activity of TiO2 photocatalysts, composites have become mainstream. In addition to the improvement of photocatalysis, composites can yield other benefits. For example, composites can tune the surface properties, i.e., ability to adsorb pollutants. Composites are also beneficial toward the stabilization of nanoparticles against phenomena such as sintering or aggregation [34].
The modification of TiO2 to overcome the problems in the use of pure TiO2 is one of the widely studied topics in the field of photocatalysis. Modified TiO2 can improve its photocatalytic activity from different mechanisms, including the reduction of the band gap of TiO2-based materials, the decrease of the probability of recombination between electron and hole. In recent years, different aspects are applied to improve the photocatalytic efficiency of TiO2. One of the methods is ion-doped TiO2 or coupled with other semiconductor composites to reduce the forbidden band width of TiO2 and increase its absorption wavelength [35,36,37]. Another way is to deposit precious metals or metal oxides on the surface of TiO2 and add electron capture agents or use photo-catalysis to prevent TiO2 photo-generated electron-hole pair recombination thus improve the photocatalytic efficiency of TiO2 [38,39,40]. In addition, dye photosensitization and providing a suitable carrier for TiO2 would be the efficient methods to modify TiO2 [41].
In recent years, many reviews have been conducted on the modification of TiO2. Serpone [42] reviewed the different mechanisms of anion- and cation-doping TiO2. He reported that TiO2 photocatalysts doped with either anions or cations have recently been shown to have their absorption edge red-shifted to lower energies (longer wavelengths), thus enhancing photonic efficiencies of photoassisted surface redox reactions, and author argued that the red-shift of the absorption edge is due to the formation of color centers. Devi and Kavitha [43] reviewed the photocatalytic activity of non-metal doped TiO2 for a wide variety of pollutant degradation under UV/visible light, with special emphasis on nitrogen-doped TiO2. They also discussed the mechanisms of photocatalytic reactions according to the charge carrier generation–separation–transfer–recombination dynamics together with pollutant adsorption and their reactions with reactive oxygenated species in liquid or gaseous regime. Asahi et al. [44] reviewed previous studies on nonmetal-doped TiO2 for visible-light sensitization. Among the enormous number of studies and references on this topic, they focused on N-doped TiO2. The present review will concentrate on the application of modified TiO2 in different media. First, this review summarizes the principles and types of different materials for modifying TiO2. Then, it discusses the application and progress of modified TiO2 in treating different pollutants. Finally, it assesses the critical application challenges and potential future research directions.

2. Non-Metallic Materials Modified TiO2

2.1. Non-Metallic Materials Supported TiO2

Loading TiO2 on the carrier can effectively overcome the problems mentioned above [45]. At present, the non-metallic materials commonly used in loading TiO2 can be divided into non-metallic porous minerals, glasses, carbon materials, and polymer materials.
The treatment of polluted water by porous mineral composites has aroused attention because of the advantages of high specific surface area, strong adsorption characteristics, and the ability of targeted enrichment of pollutants with the development of non-metallic porous minerals. [46]. As shown in Figure 1, TiO2 loads into the pores or surfaces of mineral materials and then forms the non-metallic porous minerals/nano-TiO2 composite system, which can solve the problem of the agglomeration of nano-TiO2 particles. Pollutants can be adsorbed to the surface of nano-TiO2 through the ion exchange and increase the contact probability of catalysts with pollutants to improve the degradation rate. Hence, porous mineral/nano-TiO2 composite systems can improve the photodegradation efficiency of nano-TiO2. Liang et al. [47] used montmorillonite as the matrix to prepare a TiO2/montmorillonite composite photocatalyst. The results showed that the montmorillonite matrix improved the capacity of optical absorption capacity from 70% to 87% because the visible light absorption ability (390–780 nm) of the composites was enhanced compared with pure TiO2. Moreover, the ultraviolet light absorption ability of montmorillonite-supported nano-TiO2 composites was improved. Therefore, adding montmorillonite carrier enhances the absorbance of visible light and ultraviolet light. Zhu et al. [48] prepared Mn-TiO2/sepiolite photocatalytic material using the sol-gel method at different calcining temperatures. They found that the degradation rate of emerald dye could reach 98% when the calcining temperature was 400 °C because sepiolite and Mn broaden the spectral response range of TiO2. Saqib et al. [49] prepared a nano-TiO2 supported on zeolite using the liquid impregnation method. The experimental results showed that the degradation rate of methylene blue (MB) dye by the loaded material was four times that of pure TiO2 because zeolite has a UV-visible radiation transparency, which allowed the excitation of light to penetrate into opaque solid powder and reach the substrate molecules present in intraparticles spaces. Zeolite might have substantially enhanced the proximity of organic molecules to the available active sites where the degradation reaction needs to take place.
Glass has good light transmission properties and is easy to make into photoreactors with different shapes. Therefore, glass-based carriers have also received widespread attention. The form of glass carrier includes glass sheet, glass tube, glass spring, hollow glass bead, and glass fiber. Malakootian et al. [50] fixed nano-TiO2 on a glass plate and the removal rate of ciprofloxacin by the composite material reached 93%. Espino-Estévez et al. [51] attached the self-made high-activity TiO2 to the inner wall of the glass tube reactor by the dipping-lifting method. These results showed that the material had good photocatalytic effect and recycling performance under the ultraviolet light for the degradation of phenol, diclofenac and isoproturon. The degradation rates reached 81% for phenol, 68% for diclofenac and 57% for isoproturon. This was because the decrease of the size of the TiO2 aggregates The SEM images showed that coatings prepared after milling the TiO2 suspension were more homogeneous without surface aggregates, which increased the contact area with pollutants. However, the glass surface is relatively smooth, which makes TiO2 have poor adhesion on it and causes uneven loading of TiO2.
Combining carbon materials with TiO2 can substantially improve the photocatalytic activity of TiO2 mainly because C in the carbon material can effectively promote the separation of photogenerated electrons and holes as an electron trap [52]. Carbon materials include carbon fiber (CF), carbon nitride, activated carbon, carbon nanotubes, and graphene. Chu et al. [53] used the microwave hydrothermal method to load nano-TiO2 on CF. The experimental results showed that TiO2/CF has a good photocatalytic activity. The degradation rate of rhodamine B after 1 h of ultraviolet light irradiation reached 95%. When the catalyst was used for 10 cycles, the degradation rate of the dye still reached 88%. They also reported that compared with pure TiO2 particles, TiO2/CF was easily recycled when used as a photocatalyst. Nitric acid oxidation treatment of CF generated polar functional groups, which improved the bonding properties between TiO2 and CF. Hu et al. [54] prepared 3D flower-like g-C3N4/TiO2 composite spherical materials (FCTCMs) using a simple solvothermal method and studied their photocatalytic performance. The results showed that the photocatalytic performance of FCTCMs under visible light was twice that of FTMs (3D flower-like structure of TiO2 microspheres) because g-C3N4 can broaden the photoresponse range and reduce the recombination rate of photogenerated electron–hole pairs. Cunha et al. [55] prepared TiO2/activated carbon composites (TiO2/AC), and the degradation rate of benzodiazapine drugs reached 98% mainly due to the sorption capacity of activated carbons. Thus, pollutants tend to be adsorbed more efficiently on the surface of the composite close to the TiO2 catalyst. Moreover, TiO2/AC composites enhance the generation of superoxide radicals and hydroxyl radicals.
Titanate nanomaterials (especially titanate nanotubes, TNTs) as the carriers of TiO2 and the adsorbents for heavy metals have attracted great attention from researchers [56]. Because their high specific area, great ion-exchange properties, easy solid–liquid separation and abundant functional groups [57,58]. As the carriers of TiO2, titanate nanomaterials can provide lots of nano-scale reactive sites when reacting with contaminants and improve the separation between catalysts and pollutants for the good sedimentation property of titanate nanomaterials [31]. Liu et al. [59] synthesized the TiO2/titanate nanosheet composite material (TNS) and used of Cr(VI) and 4-cholophenol (4-CP) as the target pollutants to tested its performance. The resulted showed that the Cr(VI) removal efficiency attained 99.7% within 120 min and the removal efficiency of 4-CP was 98% within 60 min. They reported that the excellent performance was mainly because the synergetic photocatalysis and autosynchronous doping. The efficiency of the separation between electron-hole pairs was enhanced due to the combination of photo-reduction of Cr(VI) by electrons and photo-degradation of 4-CP by holes. In addition, the adsorption of the reduced Cr(III) by TNS can narrowed energy band gap and enhanced photocatalytic activity of the materials. Zhao et al. [60] constructed the TiO2 decorated titanate nanotubes composite (TiO2/TNTs) and used for photocatalytic degradation of bisphenol A (BPA). They compared the performance of the TiO2, TNTs, and TiO2/TNTs. The experiment result showed that in the first cycle, the degradation rate of BPA using TiO2, TNTs, and TiO2/TNTs was 100%, 5.8%, and 94% under UV light. Although the removal efficiency of BPA by TiO2 is slightly higher than that of TiO2/TNTs in the first cycle, reusability of TiO2/TNTs was proved in the next cycles. After five reuse cycles, the degradation rate of BPA still reached 91% by using TiO2/TNTs and 95.8% of the material could be separated after 6 h gravity-settling, while the 1.8% BPA was removed and 93% of TiO2 will lose after gravity settling and cannot be reused in pure TiO2 group. Besides, Li et al. [61], Cheng et al. [62], and Ji et al. [63] also used titanate materials modified TiO2 to treated the 4-chlorophenol, phenanthrene, and sulfamethazine, respectively.
In addition, many types of polymer materials have been chosen as catalyst carriers, such as polyethylene (PE), polyvinylidene chloride (PVDF), and polyaniline (PANI) Tu et al. [64] incorporated rectorite (REC) into a porous polycaprolactone (PCL)/TiO2 nanofiber and tested its photocatalytic performance. The results showed that the degradation rate of PCL/TiO2/REC to rhodamine B was 98%. Because the porous PCL mats could provide large contact area for TiO2. Besides, the addition of rectorite (REC) could reduce the diameters of fibers and enlarge the specific surface area, which might improve the photocatalytic efficiency. Ni et al. [65] fixed nano-TiO2 nanometers in a polyvinyl alcohol–ethylene–ethylene nanofiber scaffold and tested its photocatalytic performance by using methylene blue (MB) as a target pollutant. The experimental results showed that the degradation rate of MB was 97% in 150 min. Mohsenzadeh et al. [66] synthesized the PANI-TiO2 nanocomposite using the n-situ deposition oxidative polymerization method. The results showed that the material had good photocatalytic degradation performance for 1,2-dichloroethane wastewater because of the large contact area for TiO2.
Although non-metallic materials loaded TiO2 can improve the photocatalytic activity, there are still many problems. High-quality carriers and immobilization methods to complete the photocatalyst loading is necessary. Furthermore, the interaction between carriers and photocatalyst and its effect on catalytic efficiency need to be investigated.

2.2. Non-Metal Element Doping Modification

The doping of non-metal elements has always been a hot spot in the field of photocatalytic modification. Doping non-metal atoms can broaden the photoresponse range of TiO2. Non-metal ion doping reconstructs TiO2 valence band and moves it upward, which can shorten the gap width (Figure 2).
The lifetime of photogenerated electron-hole pairs and the region of light response to visible light are vital for TiO2 to increase the value in pollutant treatment [67]. In recent years, it has been found that doping non-metallic elements such as S, N, C, B, I, and F to TiO2 successfully extended the optical response range of nano-TiO2 to the visible light region. Non-metal element doping means that oxygen or partial oxygen in the TiO2 was replaced by non-metals. The new energy band was reconstructed and the width of the forbidden band was shortened during doping the non-metal element [68,69,70]. As a result, the light response range was expanded. Doping non-metallic elements could increase the impurity level in the forbidden band of TiO2 and help the band energy level keep higher than the reduction level, where the energy level of TiO2 are overlap [71,72,73]. Asahi et al. [74] first replaced a small amount of lattice oxygen in TiO2 with non-metallic element N doping, and successfully achieved visible light catalytic activity of TiO2. The preparation method of non-metal element doped TiO2 photocatalyst can be divided into post-treatment and process treatment. Process treatment means that non-metal elements was doped during the formation of TiO2, and post-treatment means that non-metal elements was doped after TiO2 formation.
Li et al. [75] prepared N-doped TiO2 using the sol-gel method, and the degradation results of dye methylene orange (MO) showed that the degradation rate of pure TiO2 was less than 5% after 180 min under visible light, whereas that using N-doped TiO2 as the catalyst was over 95% after 90 min. N doping substantially enhanced the ability of TiO2 to degrade MO under visible light because N impurity and Ti3+ acted cooperatively to narrow the band gap of N-doped TiO2. Jyothi et al. [76] prepared N and F co-doped TiO2 using the hydrothermal method to remove bromoethane in the solution. The doping element inhibited the photogenerated electron–hole recombination to generate the hydroxyl radical (•OH). The synergistic effect increased the removal rate of bromoethane for 90 min from 54% of pure TiO2 to 94%. Rahbar et al. [77] prepared S and N co-doped carbon quantum dots (CQDs)/TiO2 composites using the hydrothermal method. The degradation rate of acidic AR88 (azo dyes) under visible light irradiation was 54%, which was higher than that of pure TiO2. They also reported that CQDs allowed the separation of charges due to the electron transport characteristics. In addition, the surface functional group enhanced the photocatalytic activity by providing a higher adsorption capacity on the photocatalyst surface, and pollutant molecules were adsorbed on the photocatalyst surface to promote the photocatalytic reaction. Liu et al. [78] used I-doped TiO2 (I-TiO2) material as the electrode of the photoelectrocatalytic method. The experimental results showed that the removal efficiency of diclofenac through modified photoelectrode reached 60%, whereas only 10% was removed using the TiO2 photoelectrode after 2 h (1.4 V) of visible light. Photoexcited electrons in the conduction band (CB) of TiO2 can be accepted by the carbon structure due to the high electron storage capacity of carbon materials such as carbon nanotubes. Therefore, the doping of non-metal element not only extends the photocatalytic reaction to visible light, but also improves the photon efficiency of TiO2 by promoting the separation of charge carriers.
Although doping non-metallic elements improves the visible light response of TiO2, the band gap width reduces. As a result, the oxidizing ability of the TiO2 nanocrystalline phase is directly reduced, and the adsorbed substances cannot be completely degraded. Therefore, the development of non-metal doping remains a hot and difficult issue in the field of photocatalysis.

3. Metal Materials Modified TiO2

Metal materials such as stainless steel, nickel mesh and nickel foam can be used as carriers for TiO2 to solve the pollutant problems. However, metals are generally expensive and damage the crystal lattice in some respects. Therefore, metals are hardly used as carriers. Since the surface of metal is similar to that of glass, it generally has poor adhesion and is difficult to load. Hence, precious metal deposition and metal ion doping are mostly used to modify TiO2. At present, the metal materials which are used to modify TiO2 include transition metals (Cr, Fe, and Cu) [79,80,81], precious metals (Ag, Au, and Pt) [82,83,84], and rare earth metals (Ce, La, and Nd) [85,86,87].

3.1. Precious Metal Materials Deposition

Precious metal materials with a large radius are easy to deposit on the surface of TiO2 particles and can be used as an effective trap for electrons when a certain amount of precious metals is deposited [88,89,90]. As shown in Figure 3, electrons can transfer from the surface of TiO2 with a higher Fermi level to the surface of the precious metal with a lower Fermi level. When the Fermi levels of the two surfaces are the same, the electrons will stop transferring and form a Schottky barrier, which can effectively separate photogenerated electron–hole pairs and improve the photocatalytic activity of TiO2 [70,91,92]. Moreover, depositing an appropriate amount of precious metals on the surface of TiO2 can broaden the response range of TiO2 to sunlight and improve the utilization of solar energy, that is, the mechanism of depositing precious metals on the surface of TiO2 to improve the photocatalytic efficiency changes the surface properties of TiO2. As a result, the number of electrons on the surface of TiO2 is reduced, the separation of photogenerated electron–hole pair is promoted, [93,94,95] and the photoelectric conversion efficiency is improved. Precious metal deposition can improve photocatalytic performance, but the deposition amount on the metal surface must be controlled within a suitable range. If the deposition amount is very large, the metal may become the center of recombination of electrons and holes, which improves the probability of the electron–hole recombination. Therefore, it is not conducive to photocatalytic degradation [96,97]. Precious metal deposition modification has a high selectivity for photocatalytic degradation of organics [98].
In the 1980s, Sato et al. [99] reported that Pt deposited on the surface of TiO2 enhanced the photocatalytic efficiency of water conversion to H2 and O2. Later, Kennedy et al. [100] reported that the incorporation of thermally oxidizable Pt into TiO2 increased the photocatalytic activity of TiO2. The oxidation activity of Pt/TiO2 was higher than the sum of the oxidizing properties of pure Pt and pure TiO2. The improvement was due to the accumulation of holes at the Pt/TiO2 interface, which led to a decrease in electron-hole recombination in TiO2. What is more, the desorption of photo-oxidized intermediated on the surface of TiO2 and re-adsorption on the surface of Pt with associated thermal oxidation. Ji et al. [101] prepared the Ag-Carbon-TiO2 composite by using polystyrene/AgNO3 composite fibers as a sacrifice template. They found that the degradation rate of Rhodamine B reached 90%, when the reaction ran 6 h. Shan et al. [102] synthesized biochar-coupled Ag and TiO2 composites by mixing, calcination, and photodeposition method. They tested the photocatalytic performance of the material by using methyl orange (MO) as target pollutants. The result shows that TiO2 modified with Ag showed better photocatalytic degradation performance (the highest decolorization efficiency and mineralization efficiency were 97.48% and 85.38%, respectively) than pure TiO2. They attributed the increase in catalytic efficiency to the promotion of the separation of photogenerated electron hole pairs. Jaafar et al. [103] used in-situ electrochemical method to deposit Ag nanoparticles on the surface of TiO2 and degraded chlorophenol to measure its photocatalytic activity. The study found that Ag-TiO2 catalyst degraded chlorophenol to 94% after 6 h, for the electron–hole separation had been enhanced.
The precious metal deposition on the surface of TiO2 achieves a relatively obvious modification effect that significantly increases the degradation rate of some organic compounds. However, the cost of the precious metal deposition method is very expensive and precious metal deposition modified TiO2 has a high selectivity for photocatalytic degradation of organics, which further limit the application of these materials in pollution treatment.

3.2. Metal Ion Doping TiO2

Doping different metal ions in TiO2 photocatalyst is an effective method to improve its catalytic activity [80,104,105]. Metal ions doped with TiO2 can change the corresponding energy level structure of TiO2 because metals are more active, and electrons are more easily excited, resulting in a wider range of the absorption in a TiO2 system [81,106,107]. As shown in Figure 4, metals can capture the electrons generated by TiO2 excitation, and the electrons inside TiO2 are not easy to return to the original state. After metal doping, metal ions can act as a carrier-trapping center, where metal ions higher than tetravalent are more likely to acquire electrons than titanium ions and metal ions lower than tetravalent trap holes. Ion doping can then stop the recombination of electron–hole pairs, which enables TiO2 to generate more electrons and holes. Thus, the photocatalytic efficiency of TiO2 is improved [108,109]. For metals with many valence states, electrons in the d orbitals can transition and enter the TiO2 lattice, which can reduce the band gap and the energy required for the electron transitions [110,111]. This is vital to improve the activity of TiO2 photocatalyst [112]. The transition group metal ions doped with TiO2 can change the crystalline morphology and energy level structure of nano-sized TiO2 to form impurity energy levels [68]. Photons with lower energy can also undergo transitions, thereby expanding their absorption wavelength range and improving the utilization of visible light [113].
As early as 1994, Choi et al. [114] began studying the doping of metal ions with TiO2. In the experiment, 19 kinds of transition metal ions were doped into nano-TiO2 and showed a better catalytic activity with the dope of Fe3+, Mo5+, Os3+, and Rh3+ due to the match of doped ionic potential and radius with TiO2. Du et al. [115] used Ge4+-doped TiO2 to prepare Ge/TiO2 photocatalytic materials and degraded ciprofloxacin. The radius of Ge4+ was 0.054 nm, which was smaller than that of Ti4+ (0.068 nm). Ge entered the TiO2 lattice and replaced the position of Ti, causing lattice defects and delaying the recombination of electrons and holes. Therefore, the formation of •OH on the surface of TiO2 and the photocatalytic efficiency increased. Degradation rate reached 97% when the calcination temperature was 571 °C and the doping amount was 0.26%. Crisan et al. [116] prepared Fe-doped nano-TiO2 using the sol-gel method, and the absorption spectrum was extended to 546 nm. Moreover, the corresponding band width at the wavelength of 410 nm was 3.03 eV when the Fe content (w) was 2%. Compared with pure TiO2, the degradation rate of nitrobenzene with 0.5% Fe nano-TiO2 was increased from 70% to 88%. Gnanasekaran et al. [117] found that the spectral absorption range of Co-doped TiO2 was extended from 382 to 411 nm when the band gap width was reduced to 3.01 eV. The degradation rate of MO after visible light catalytic treatment for 240 min reached 53%, which was beneficial to its photocatalytic performance. Huang et al. [118] used the sol-gel method to prepare Mo-doped nano-TiO2 powders, and the degradation rate of MB reached 98% under outdoor sunlight with the amount of 2% Mo6+(w). This result was mainly due to the reduction of the forbidden band from 3.05 to 2.73 eV with the doping of Mo6+ and the wider excitation absorption wavelength. Bhatia and Dhir [112] made Ni-TiO2 and Bi-TiO2 using the sol-gel method and found that the maximum degradation rate of ibuprofen by Bi-TiO2 and Ni-TiO2 reached 89% and 78%, respectively. This finding may be attributed to the increase in specific surface area and the decrease in the crystallite size. Wang et al. [119] prepared an Fe3+-doped TiO2 nanotube array catalyst using a simple hydrothermal method, which increased the degradation rate of MB by about 20%. The study also found that Fe3+ doping provided traps in the TiO2 lattice, which greatly improved the separation effect of electron–hole pairs.

4. Composite Materials Modified TiO2

4.1. The Construction the Heterojunction

The heterojunction is a contact interface, formed as a result of hybridization between two semiconductors [120,121]. The semiconductors used for the heterojunction need to satisfy the condition that they should exhibit different band gaps and the narrow band gap must lie in the visible region [122]. Combining TiO2 with other semiconductors to construct heterojunction can efficiently improve the photocatalytic performance of TiO2 [123,124]. This method can not only improve the effective utilization rate of the electrons by promoting the photo-generated electrons and holes to transfer in the opposite direction but also expand the spectral response range of the composite to visible light and even near infrared region [125,126,127]. Generally, the most widely researched types of the TiO2-based heterojunction can be categorized into two different types depending on the charge carrier separation mechanism, which are conventional type and direct Z-scheme [128,129,130].
Based on the different band and electronic structures, the conventional type can be divided in to three, namely, type-I (straddling gap), type-II (staggered gap), and type-III (broken gap) heterojunctions [131,132]. For type-I heterojunctions, the level of the CB of semiconductor-I is higher than that of semiconductor-II, while the valence band (VB) of semiconductor-I is lower than that of semiconductor- II. However, due to the difference between the band gaps, the photoinduced charges accumulate on smaller band gap semiconductor, which may cause recombination. In type-II heterojunctions (Figure 5b), the level of CB and the VB of semiconductor-II are higher than those of semiconductor-I [133]. In addition, the migration of charge carriers to the opposite directions can be promoted because the difference between the chemical potentials causes a phenomenon called band bending. The band structure of the type-III heterojunctions (Figure 5c) is similar to that of type-II except that the staggered gap becomes so wide that the bandgaps do not overlap [134]. Among these conventional heterojunctions, type-II heterojunction attracts the attention of more researchers [135,136,137]. Ganguly et al. [138] synthesized type-II heterojunctions of the AgBiS2-TiO2 composite and used doxycycline as the target pollutant to test photocatalytic performance. The results showed that the degradation rate reached 100% in 180 min under a 500 W Xe lamp. The enhanced photocatalytic activity was attributed to the decreased rate of recombination of the photogenerated excitons. Liu et al. [139] used other semiconductors such as Bi2MoO6 and TiO2 to fabricate type-II heterojunctions and tested the photocatalytic performance of Bi2MoO6/TiO2. They reported that the degradation rate of ciprofloxacin, tetracycline, and oxytetracyline reached 88%, 78%, and 78%, within 150 min, respectively, when the 350 W Xe lamp with a 420 nm cutoff filter was used as the light. The CB of TiO2 can serve as the electron transfer platform, which can improve the efficiency of the separation of photocarriers at Bi2MoO6/TiO2 heterojunction interface.
In 2013, the concept of the direct Z-scheme photocatalyst was first proposed [140]. Figure 5d is the band arrangement and electron migration mechanism of Z-scheme heterojunctions. The Z-scheme heterojunctions have the same band arrangement as the type-II heterojunctions, but the electron transfer path between semiconductors is different [141]. The electron transfer path between semiconductors is like the English letter “Z” [142]. In the process of photocatalytic reaction, the photogenerated electrons with lower reduction ability in semiconductor-II recombine with the photogenerated holes in semiconductor-I with lower oxidation ability. Therefore, the photogenerated electrons with high reduction ability in semiconductor-I and the photogenerated holes with high oxidation ability in semiconductor-B can be maintained [143]. In addition, the electrostatic attraction between the photogenerated electron on the CB of the semiconductor-II and the photogenerated holes on the VB of the semiconductor-I will promote the migration of the photogenerated electron from the semiconductor-II to the semiconductor-I, while in the type-II heterojunction, the electrostatic repulsion between the photogenerated electron of the semiconductor-I and the semiconductor-II will inhibit the transfer of electrons from semiconductor-I to the semiconductor-II [144,145]. So far, many photocatalytic composites that have the Z-scheme heterojunctions have been manufactured to degrade the pollutants. Wang et al. [146] fabricated the N-doped carbon quantum dot (NCDs)/TiO2 nanosheet with higher surface energy faceted (NCDs/TNS-001) composites and used diclofenac (DCF) as the target pollutants. The photocatalytic efficiency of the composites reached 92% in 60 min under the 350 W Xe lamp. In contrast, only 15.4% of the DCF was degraded in the presence of TNS-001 after 60 min. They reported that the excellent photocatalytic performance might be attributed to the synergistic effects of the highly active facets, up-converted fluorescent properties of NCDs, and efficient charge separation induced by fabricated Z-scheme heterostructures. Hao et al. [147] used the TiO2@g-C3N4 core-shell photocatalysts with the Z-scheme heterojunctions to remove the Rhodamine B from water. The removal efficiency under the 100 Xe lamp was about 96% within 180 min, while the Rhodamine B (RhB) dye shows almost no degradation in the blank test. They attributed the improvement of photocatalytic performance to the formation of the Z-scheme system, which effectively separated photogenerated electrons and holes. Liao et al. [148] prepared a photocatalytic material g-C3N4–Ti3+/TiO2 nanotube arrays and tested its performance of degrading the phenol. At a reaction period of 7 h, the degradation was only 23.4% using TiO2 nanotube arrays, while the degradation rate increased to 74% using the g-C3N4–Ti3+/TiO2 nanotube arrays. This was mainly because that the self-doping of Ti3+ promoted the visible light absorption behavior of the composite and the Z-scheme heterojunctions with efficient space separation of the photo-generated electron–hole.
Semiconductor heterojunction powders have exhibited the enhanced photocatalytic activities, but their practical applications have been limited due to their poor recycling performance from flowing wastewater.

4.2. Different Elements Co-Doping TiO2

The emphasis of single element doping on the modification of TiO2 is different. In order to improve the migration range of absorption edge, photocatalytic performance and thermal stability of TiO2 at the same time, the co-doping of multiple elements is an ideal solution [149]. Co-doped nanoparticles exhibit higher visible light absorption than single doped TiO2 due to a synergistic effect between the two dopants, which can efficiently increase the photocatalytic performance [150]. Co-doping can be divided into different metal elements co-doping [151,152,153], metal elements and non-metal elements doping [154,155,156] and different non-metal elements doping [157,158,159]. As shown in Table 2, many researchers have used co-doping method to modify TiO2 and tested the photocatalytic performance of the materials.
At present, the physical and chemical properties and the doping mechanism of co-doped TiO2 with two different metals have not been thoroughly investigated. Singh et al. [170] synthesized the mesoporous La-Na co-doped TiO2 nanoparticles (NPs). The removal efficiency of MB was almost 100% by using the Na and La doped TiO2, while 35% MB was degraded by using pure TiO2 mainly because of the substitution of large-sized Na+1 and La+3 at Ti+4 sites which was confirmed by the results of XRD and TEM. The doping of these low-valent metal ions led to the formation of O vacancies, which promoted the adsorption of hydroxyl groups on the surface of NPs. The adsorbed hydroxyl group reduced the pHIEP, which was beneficial to the adsorption of cationic MB dyes. Metal components prefer to substitute for the Ti site in the TiO2 lattice to create the dopant level near the CB. Non-metal components can form new levels closest to the VB that reduce the band gap and cause visible light absorption. Therefore, metal and non-metal ion co-doping enhance photocatalytic activity [163,171]. Garg et al. [172] tested the photocatalytic performance of prepared N and Co-co-doped TiO2 on the removal of Bisphenol-A under visible light. The results showed that the maximum degradation rate (95%) was observed when using 1.5% Co and 0.5% N co-doping TiO2. This result was almost twice that of the group using pure TiO2. TiO2 was enhanced because Co and N disturbed the physical properties of the nano particles, producing alterations in crystal structure and energy band gap as well as elemental composition. N could easily substitute O in the TiO2 lattice owing to its atomic size comparable with that of O, and N had small ionization energy and high stability. In addition, the doping of a range of Co could shift the optical absorption edge from UV to visible light range, and Co could behave as recombination centers for the photoinduced charge carriers, thereby decreasing quantum efficiency. Non-metal co-doped TiO2 have been studied extensively [173,174,175]. Zeng et al. [176] prepared B/N co-doped TiO2 photocatalysts and compared their photocatalytic performance with pure TiO2 under simulated sunlight by using flumequine (FLU) as the target compound. The results showed that the degradation rate of FLU by B/N co-doped TiO2 was nearly 100%, whereas that of pure TiO2 was only about 10%. The photocatalytic performance of TiO2 catalyst was evidently enhanced by B/N co-doping. The relative content of rutile in B/N co-doped TiO2 catalysts increased with the increase of the B content, which produced a synergistic effect between anatase and rutile. This synergistic effect can be explained by the formation of a semiconductor junction between the anatase phase and the rutile phase, which promoted the separation of photogenerated electrons and holes, thus improving photocatalytic activity.
The method of co-doping can effectively improve the removal efficiency of pollutants by TiO2. But some of the elements are not suitable for practical use, so it is necessary to find suitable doping materials. And it is essential to find an optimum amount of dopant to increase the separation of charge carriers and prevent the formation of a recombination center.

4.3. Dye Photosensitization

Dye photosensitization means that the photosensitizer (dyes) binds to TiO2 surface by chemical or physical adsorption, so that the absorption wavelength of visible light shifts to the long wavelength, thus expanding the excitation wavelength response range of TiO2 and greatly improving the utilization of sunlight [177,178,179]. The molecule (dyes) absorbing the photon is called as a photosensitizer and the altered material (TiO2) is the acceptor or substrate [180]. As shown in Figure 6, the mechanism of photosensitization is that once the dyes achieve their excited state by the absorption of photons in the visible range of the solar spectrum, electrons from the dyes’ highest occupied molecular orbital (HOMO) are transferred to their lowest unoccupied molecular orbital (LUMO) and subsequently to the conduction band (CB) of TiO2 [181,182,183]. In addition, the dyes in solution can be excited to a triplet state under visible light and transfer their excess energy to the O2. Thus, the electrons in the LUMO react with dissolved oxygen and produce the superoxide anion radical [184]. Dyes used for photosensitization must meet the following characteristics: strong absorption of visible light even the part of the near infrared (NIR) region, photo stability (unless the self-sensitized degradation is required), the existence of some anchoring groups (-SO3H, -COOH, -H2PO3, etc.) and the higher excited state energy than the conduction band (CB) edge of TiO2 [185,186]. According to the composition, dyes can be divided into two categories: organometallic dyes and organic dyes. Organometallic dyes contain a transition metal in the structure and the organic dyes are composed of organic chromophores [187].
Ahmad and Kan [188] used a phthalocyanine-based reactive dye, C.I. reactive blue 25 (RB-25), as a dye photosensitizer for anatase (TiO2) and tested photocatalytic performance by degradation of Rhodamine B (RhB). The result showed that RhB was relatively stable under visible light in the presence of only TiO2. Degradation rate reached more than 90% when the RB-25 dye was adsorbed on the TiO2 mainly because the electrons from the dye increased the electron density in the CB of the TiO2, which enhanced photocatalytic activity under visible light. Mucria et al. [189] modified TiO2 by dye sensitization. The photosensitizers applied were quinizarin and zinc protoporphyrin. The result showed that the removal efficiency of both exceeded 80%. The improvement of their activity could be ascribed to the presence of sensitizing molecules within the nanotubes, whose electronic properties were promoted by the electron confinement effect in semiconductors. Moreover, electron–hole recombination rate for this material in comparison to the higher surface materials because the electrons can initially reach the dye before the CB. Sensitized materials can also be used for desulphurization. Guo et al. [190] loaded TiO2 onto SBA-15 molecular sieves and sensitized with organic dyes (2,9dichloroquinacridone, DCQ) to extend its spectral response range from ultraviolet light to visible light. The material was then applied for the photocatalytic oxidation desulfurization of gasoline. Experimental results showed that DCQ-TiO2@SBA-15 performed better than unsensitized TiO2@SBA-15, and desulfurization rate can reach 96.1% in a reaction time of 90 min.
The modification method of photosensitization can greatly improve the photocatalytic performance of titanium dioxide. However, there are still some problems need to be solved. For example, the organic dye molecule will gradually degrade due to the photocatalytic. So, it is necessary to replace the catalyst continuously. Additionally, the absorption of most photosensitizers is weak in the near-infrared region and there is adsorption competition with pollutants, which limits the development of photosensitization. Therefore, further research is needed to solve these problems.

5. Application of Modified TiO2 Composite Photocatalytic Materials

Photocatalytic treatment technology is the most representative advanced oxidation technology for environmental pollution treatment. It uses the hydroxyl radical (•OH) as a strong oxidant to deeply oxidize and decompose organic pollutants into non-toxic inorganic small molecules [191]. At the same time, the photocatalytic reduction reaction can effectively remove heavy metal ions. Table 3 is the performance of the modified TiO2 in treating pollutants.

5.1. The Application in Water Pollution

Wastewater treatment plants can remove a large majority of the pollution. However, several trace organic compounds or refractory compounds cannot be degraded by the conventional treatment [216]. These pollutants mostly result from domestic and industrial use of pharmaceutical preparations, printing and hygiene products, and pesticides. In recent years, photocatalytic technology has broadened its application in the treatment of organic wastewater. Under the conditions of sufficient O and light, TiO2 uses the photogenerated electrons and holes to degrade almost all organic pollutants in water and convert them into CO2, H2O, and other inorganic substances. Saif et al. [217] prepared lanthanide (Nd3+, Sm3+, Eu3+, Gd3+, Dy3+, and Er3+)-doped TiO2 using the sol-gel method and evaluated their photocatalytic activity in the treatment of actual sewage. In the actual sewage treatment plant application, the mineralization efficiency of Gd3+-TiO2 and Eu3+ TiO2 on chemical oxygen demand (COD) reached 67% and 50%, respectively, after 6 h of light. Lima et al. [218] used Ag/TiO2 photocatalytic material to degrade the hormones which would exist in the sewage treatment plant. The results showed that the degradation rate of the hormone by Ag/TiO2 reached 95% after 3.5 h. In addition to organic pollutants, photocatalytic technology can achieve better treatment of inorganic substances in water. Peng et al. [219] prepared TiO2-CuO/HSC composites. Degradation rates of ammonia nitrogen in water reached 61% and 100% when experiments were operated under ordinary light and ultraviolet light, respectively.
Dyeing wastewater discharged from printing and dyeing factories contains a large number of dye molecules. These dye molecules usually contain a mine groups, aromatic rings, azo groups, etc. Therefore, the chromaticity of dye wastewater is difficult to meet standards for discharge. Traditional biological and physical treatment methods for removing organic pollutants, which include precipitation, adsorption, flocculation, reverse osmosis, and ultrafiltration, are inefficient and unsuitable for industrial applications. So, people choose photocatalytic degradation as an alternative technology to solve the pollutants [220,221,222]. TiO2 can not only effectively remove the color of wastewater, but also decomposes the pollutants into small molecules such as CO2 and H2O. Xu et al. [223] successfully prepared Ag/TiO2 layered structure and reported the degradation rate of the dye under the sunlight reached 99%, while the degradation rate of pure TiO2 was 43%. Ji et al. [224] used C/TiO2 microsphere to degrade the rhoda mine B. The results showed the degradation rate of rhoda mine B reached 96% within 140 min, and the degradation rate still maintained above 80% after the material was used for three times. Fu et al. [225] use graphene oxide/TiO2 (GO/TiO2) composites to degrade the dyes and reported the degradation rate of the dyes reached 96% within 2.5 h.
Tannery wastewater is also a major problem in industrial wastewater because of the containing of large amounts of poorly biodegradable organic chemicals. The general biological treatment method can make the effluent reach the standard, but it still needs multiple treatments to remove most of the COD, color and some organic recalcitrant compounds. So, the photocatalytic treatment using TiO2 become a new method to treat tannery wastewater. He et al. [226] used Mn-doped TiO2 material to degrade the tannery wastewater. The degradation rate of organic pollutants under sunlight was nearly 90%. Bordes et al. [227] deposited the fine-structured photocatalytic TiO2 coatings on austenitic stainless steel coupons by atmospheric plasma spraying (APS) and the total organic carbon (TOC) removal reached 49%, while the decolorization rate reached 75%. Therefore, they reported that the decreased TOC and color removal of the resulting effluent evidenced the effectiveness of the developed coatings for photocatalytic treatment of industrial tannery wastewater.
The large-scale use of pesticides such as herbicides and insecticides have a significant impact on aquatic environment. The harmful effects of these compounds are due to their toxicity and high mobility and persistence in aqueous media [228]. In fact, only a small part of the applied pesticides can protect agricultural products, while most pesticides are lost to the environment through volatilization, hydrolysis, photolysis, or degradation by microbial action [229]. The characteristics of pesticide wastewater lead to ineffective treatment by physical and biological methods, so photocatalytic oxidation is used to treat this type of wastewater, for example organophosphorus pesticides can be mineralized and decomposed by TiO2 and converted into non-toxic CO2, H2O, and PO4−. Abdennouri et al. [230] prepared a nano-TiO2 supported on pillared clay and found that the material can efficiently degrade 2,4-dichlorophenoxyacetic acid, 2,4-dichlorophenoxypropionic acid, and other pesticides in the environment. Both of the degradation reached about 80%.
As an important component of energy, petroleum plays a decisive role in the sustained and rapid development of the national economy. Due to river convergence and marine accidents, a large amount of low-density and water-insoluble oil flows into the ocean every year, causing marine oil pollution and threating to marine life. Traditional processing is performed by mechanical methods, and biological processes are often used as auxiliary processing steps, followed by advanced processes such as adsorption, membrane filtration, and reverse osmosis. However, due to the presence of high concentrations of toxic aromatic and aliphatic hydrocarbons in addition to the presence of phenols and refractory compounds, the biological processes cannot meet the standard of reuse [231,232]. Studies have shown that TiO2 photocatalysts can float on the water surface and are efficient in the degradation of toxic and recalcitrant pollutants [233]. Shivaraju et al. [234] prepared N-doped TiO2. They reported the degradation rate of oil and grease and other organic pollutants in wastewater can reach about 90%.
The pollution of pharmaceuticals in industrial wastewater is serious problem. Traditional wastewater treatment removes most of the pollutants through sedimentation, filtration, adsorption, or biological processes, however, bio-toxic and non-degradable organics usually remain in the water at concentrations above the ppb discharge or reuse limit [235]. Using TiO2 for photocatalytic reaction would be a good method to solve this problem. Solís-Casados et al. [236] used Sn-doped TiO2 to treat diclofenac, paracetamol, and ibuprofen under visible light. The results showed the maximum removal rate of diclofenac was 25%, the maximum removal rate of paracetamol was 25%, and the removal rate of ibuprofen was 18%. All of the three drugs were effectively reduced. Malakootian et al. [237] used Fe3+ doped TiO2 materials to degrade pharmaceutical wastewater and antibiotic-added synthetic solutions. The results showed that the degradation rates of antibiotics can reach 70% and 97%, respectively. Besides, Lcerda et al. [238] and Hou et al. [239] also used the TiO2 to degrade the pharmaceutical wastewater and obtained satisfied result.

5.2. The Application in Air Pollution

Cars provide convenience for people’s travel, but the automobile exhaust gas seriously affects air quality and endangers people’s health. NOx in automobile exhaust not only stimulates the human respiratory system, but also causes problems such as acid rain and photochemical smog. So, the removal of NOx has attracted people’s attention. However, the traditional technologies of NOx removal, including physical adsorption and selective catalytic reduction, cannot remove the NOx effectively at ppb levels [240,241]. TiO2 photocatalyst provides an effective way to solve these problems by mixing in paint, concrete and brick or fixing on the surface of roads and walls. Under sunlight, the TiO2 photocatalyst can oxidize NOx to form nitric acid through a series of reactions. Then the nitric acid reacts with the components fixed the photocatalyst to obtain nitrate. Under the action of rainwater, nitrate ions are formed and washed away. Zhang et al. [242] found that nano-TiO2/diatomite composites efficiently degraded formaldehyde in the air, showing a good application prospect. Qin et al. [243] found that loading nano-TiO2 in concrete during road construction could absorb NO2 in locomotive exhaust, thereby reducing air pollution.
Besides, most volatile organic compounds (VOCs) in air (aldehydes, ketones and alcohols) are oxidizable, it is feasible to remove them by oxidation method. Most of the heterogeneous catalytic oxidation methods commonly used to remove pollutants in the air at high temperatures, which limits the application. Therefore, the photocatalysis method has become a potential method to remove air pollutants by using water vapor and O2 in air at room temperature with low energy consumption [244,245]. Lai et al. [246] prepared Bi-TiO2 to degraded toluene and the degradation rate increased by 77% in terms of CO2 production, as compared to the pure TiO2. Rao et al. [247] used Er3+ doped TiO2 and reported the modified TiO2 exhibited higher photoactivity in comparison with the pure TiO2. The highest removal efficiency of acetaldehyde and o-xylene within 100 min was 99% and 85%, respectively, and ethylene degradation efficiency reached 22% within 180 min.

5.3. The Application in Soil Pollution

There are many types of soil pollutants, which are characterized by the coexistence of emerging and old pollutants. These pollutants include heavy metals, pesticides, antibiotics, and persistent organic compounds, which makes it difficult to get an efficient repair result in soil. Heavy metal pollution and organic pollution are regarded as the main types of soil pollution because of the large amount among the pollutants [248]. The surface of the contaminated soil with a high concentration of pollutants can easily enter the atmosphere or water under the action of wind and water, respectively, leading to other secondary environmental problems such as air pollution, surface water and groundwater pollution [249,250,251]. Therefore, soil remediation is imminent.
The pollutants’ treatments in soil mainly include physical–chemical remediation technology, biological remediation technology, and phytoremediation technology, but all of these have some shortcomings. The photocatalytic technology can completely mineralize the organic pollutants and remove the heavy metals in the soil. This technology also has the advantages of fast decomposition rate, no secondary pollution, and easy operation [252,253,254]. In recent years, photocatalysis has been widely used in organic soil remediation studies, including organic pesticides, aromatic organics, petroleum hydrocarbons, and heavy metals [255,256].
TiO2 is the most widely used catalyst in photocatalyst technology, and it also acts as a key role on pollutants in soil. Kuang et al. [257] reported that the Cd(II) removal efficiency of biological soil crusts increased by 27% than that of pure biological soil crusts after the addition of nano-TiO2. They reported that in the first 30 min, the adsorption rate of BSC + TiO2 composite was faster than that of pure TiO2, which may be due to the high adsorption rate of nano-TiO2. Petroleum-contaminated soil is highly toxic, and photocatalytic degradation using TiO2 can also get the ideal results. Yang et al. [258] pretreated the soil with ultraviolet radiation C (UVC) activated TiO2 under varying moisture conditions to enhance biodegradation of heavy hydrocarbons (HCCs). They reported that total petroleum hydrocarbon (TPH) removal after 24 h exposure to UVC was about 20% in slurries with 300% water holding capacity. In a 10 d bioremediation test, TPH removal in treated soil increased to 27%, compared to 15% for controls without photocatalytic pre-treatment. The improvement mainly because the recalcitrant hydrocarbons were transformed into more bioavailable and biodegradable products so that the pollutants were more readily consumed by soil microorganisms.
Soil remediation of modified TiO2 has achieved some effects in heavy metal pollution and organic pollution, but there are still some problems, including the insufficient light penetration and difficulty in recycling. Therefore, it is necessary to find TiO2 composite materials that can make fuller use of sunlight in the soil or improve the recycling rate.

6. Conclusions and Perspective

Despite the substantial progress in TiO2, considerable opportunities and challenges remain. The synthesis and improvement of TiO2 has become a hot topic to improve the efficiency of environmental treatment. This review comprehensively discusses several synthesis and doping technologies of TiO2, and the effect of each improvement method. Moreover, it elaborates and prospects the application of TiO2-modified materials in the environmental field, especially for water, air, and soil pollution.
Developing a pollutant treatment with visible light-responsive photocatalysts is very urgent and necessary. The photocatalytic performance of TiO2 can be greatly improved through the modification of TiO2. However, many problems in application remain to be solved: (i) in many studies, the system has only one kind of pollutant, which does not match the complex multiple components of the actual pollutants. Gaps exist between material research and application studies for practical application. Whether modified TiO2 can perform well is unknown. Although modified TiO2 shows potential in the treatment of pollutants, most of the works considered in the scope of this review were carried out on a laboratory scale. (ii) The recycling or natural degradation of the modified materials remains an issue. Few studies were devoted to separation, recovery, and reuse of photocatalytic materials for the treatment of the real pollutants. (iii) Introduction of other materials into TiO2 will cause the preparation complexity and cost to increase. Materials used for modification may pollute the environment, such as in modification with heavy metal ions or harmful organic.
Future research should focus on the following aspects to improve the applicability and feasibility of the modified TiO2: (i) more pilot experiments using modified TiO2 should be performed for photocatalytic degradation of real pollutants in water, air, and soil. An understanding of inherent charge transfer dynamics and photocatalytic mechanisms at the nanometer and atomic level will be highly useful in designing effective approaches for enhancing the photocatalytic performance of TiO2. Thus, researchers should understand the mechanism of dealing with actual pollutants. (ii) The efficiency and photostability of the modified TiO2 must be improved. The performance of modified TiO2 is currently limited by the physicochemical properties of these materials. (iii) Materials used to modify TiO2 that cause low harm to the environment and can be used in large patterns must be found or synthesized. Devising an appropriate photocatalyst immobilization strategy to provide a cost-effective solid–liquid separation can save cost and avoid secondary pollution. (iv) A good reactor can improve the utilization rate of light and reduce the electricity costs. Thus, a good design of the reactor is necessary before the experiment. (v) In several cases, toxicity assessment may be even more sensitive than chemical analysis by using modified TiO2. (vi) Although modified TiO2 can have a good degradation effect on pollutants in the laboratory, the durability and recyclability of the catalysts must be considered in actual application.
Facing the problems of complex types of pollutants and tight treatment time, the comprehensive application of multiple treatments for pollutants is the development direction of the current environmental field. The combined use of photocatalysis and other technologies will broaden the application of photocatalysis technology.

Author Contributions

R.L. is mainly responsible for the writing of article content and pictures. T.L. modified the article content and pictures. Q.Z. determined the direction of the article. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by National Natural Science Foundation of China (21876090), the Tianjin Research Program of Application Foundation and Advanced Technology (18JCZDJC39400 and 19YFZCSF00920), the Postdoctoral Science Foundation of China (2019M660985) and the 111 program of the Ministry of Education of China (T2017002).

Acknowledgments

We acknowledge the use of facilities within the School of Environmental Science and Engineering at Nankai University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  2. Carey, J.H.; Lawrence, J.; Tosine, H.M. Photodechlorination of PCB’s in the presence of titanium dioxide in aqueous suspensions. Bull. Environ. Contam. Toxicol. 1976, 16, 697–701. [Google Scholar] [CrossRef]
  3. Frank, S.N.; Bard, A.J. Heterogeneous photocatalytic oxidation of cyanide and sulfite in aqueous solutions at semiconductor powders. J. Phys. Chem. 1977, 81, 1484–1488. [Google Scholar] [CrossRef]
  4. Xu, M.X.; Wang, Y.H.; Geng, J.F.; Jing, D.W. Photodecomposition of NOx on Ag/TiO2 composite catalysts in a gas phase reactor. Chem. Eng. J. 2017, 307, 181–188. [Google Scholar] [CrossRef]
  5. Karthikeyan, C.; Arunachalam, P.; Ramachandran, K.; Al-Mayouf, A.M.; Karuppuchamy, S. Recent advances in semiconductor metal oxides with enhanced methods for solar photocatalytic applications. J. Alloys Compd. 2020, 828, 154281. [Google Scholar] [CrossRef]
  6. Li, Z.Y.; Cao, F.; Wang, L.; Chen, Z.W.; Ji, X.H. A novel ternary MoS2/MoO3/TiO2 composite for fast photocatalytic degradation of rhodamine B under visible-light irradiation. New J. Chem. 2020, 44, 537–542. [Google Scholar] [CrossRef]
  7. Paeng, D.S.; Huy, B.T.; Phuong, N.T.K.; Dao, V.D.; Lee, Y.I. Photocatalytic activity of Yb, Er, Ce-codoped TiO2 for degradation of Rhodamine B and 4-chlorophenol. J. Chem. Technol. Biot. 2019. [Google Scholar] [CrossRef]
  8. Chen, Y.F.; Tang, X.N.; Zhang, B.; Luo, Y.; Li, Y. TiO2@SiO2 Composites: Preparation and Photocatalytic Antimicrobial Performance. J. Inorg. Mater. 2019, 34, 1325–1333. [Google Scholar] [CrossRef] [Green Version]
  9. Yang, D.; Zou, X.Y.; Sun, Y.Y.; Tong, Z.W.; Jiang, Z.Y. Fabrication of three-dimensional porous La-doped SrTiO3 microspheres with enhanced visible light catalytic activity for Cr(VI) reduction. Front. Chem. Sci. Eng. 2018, 12, 440–449. [Google Scholar] [CrossRef]
  10. Yu, X.; Lin, Y.; Liu, H.; Yang, C.; Peng, Y.; Du, C.; Wu, S.; Li, X.; Zhong, Y. Photocatalytic performances of heterojunction catalysts of silver phosphate modified by PANI and Cr-doped SrTiO3 for organic pollutant removal from high salinity wastewater. J. Colloid Interface Sci. 2020, 561, 379–395. [Google Scholar] [CrossRef] [PubMed]
  11. Zhao, Z.H.; Zhang, X.; Lei, Y.Y.; Yang, P.F.; Fan, J.M.; Zhang, B.; Yin, S. Exceptional photocatalytic activity for Ag,Cr- SrTiO3 activated by H2O2 for removal of organic pollutants. Mater. Res. Express 2020, 7, 015034. [Google Scholar] [CrossRef] [Green Version]
  12. Saidani, T.; Zaabat, M.; Aida, M.S.; Boudine, B. Effect of copper doping on the photocatalytic activity of ZnO thin films prepared by sol-gel method. Superlattices Microstruct. 2015, 88, 315–322. [Google Scholar] [CrossRef]
  13. Das, A.; Nair, R.G. Effect of aspect ratio on photocatalytic performance of hexagonal ZnO nanorods. J. Alloys Compd. 2020, 817, 153277. [Google Scholar] [CrossRef]
  14. Shkir, M.; Al-Shehri, B.M.; Pachamuthu, M.P.; Khan, A.; Chandekar, K.V.; AlFaify, S.; Hamdy, M.S. A remarkable improvement in photocatalytic activity of ZnO nanoparticles through Sr doping synthesized by one pot flash combustion technique for water treatments. Colloid Surf. A Physicochem. Eng. Asp. 2020, 587, 124340. [Google Scholar] [CrossRef]
  15. Mena, E.; Rey, A.; Rodriguez, E.M.; Beltran, F.J. Reaction mechanism and kinetics of DEET visible light assisted photocatalytic ozonation with WO3 catalyst. Appl. Catal. B 2017, 202, 460–472. [Google Scholar] [CrossRef]
  16. Zeng, W.G.; Cai, T.; Liu, Y.T.; Wang, L.L.; Dong, W.Y.; Chen, H.; Xia, X.N. An artificial organic-inorganic Z-scheme photocatalyst WO3@Cu@PDI supramolecular with excellent visible light absorption and photocatalytic activity. Chem. Eng. J. 2020, 381, 122691. [Google Scholar] [CrossRef]
  17. Rosaline, D.R.; Inbanathan, S.S.R.; Suganthi, A.; Rajarajan, M.; Kavitha, G.; Srinivasan, R.; Hegazy, H.H.; Umar, A.; Algarni, H.; Manikandan, E. Visible-Light Driven Photocatalytic Degradation of Eosin Yellow (EY) Dye Based on NiO-WO3 Nanoparticles. J. Nanosci. Nanotechnol. 2020, 20, 924–933. [Google Scholar] [CrossRef]
  18. Gionco, C.; Hernandez, S.; Castellino, M.; Gadhi, T.A.; Munoz-Tabares, J.A.; Cerrato, E.; Tagliaferro, A.; Russo, N.; Paganini, M.C. Synthesis and characterization of Ce and Er doped ZrO2 nanoparticles as solar light driven photocatalysts. J. Alloys Compd. 2019, 775, 896–904. [Google Scholar] [CrossRef]
  19. Al-Namshah, K.S.; Mohamed, R.M. Co3O4-ZrO2 nanocomposites: Simple preparation and enhanced photocatalytic performance for cyanide degradation under visible light. Desalin. Water Treat. 2019, 145, 318–325. [Google Scholar] [CrossRef]
  20. Reddy, C.V.; Reddy, I.N.; Ravindranadh, K.; Reddy, K.R.; Shetti, N.P.; Kim, D.; Shim, J.; Aminabhavi, T.M. Copper-doped ZrO2 nanoparticles as high-performance catalysts for efficient removal of toxic organic pollutants and stable solar water oxidation. J. Environ. Manag. 2020, 260, 110088. [Google Scholar] [CrossRef]
  21. Nguyen, T.B.; Huang, C.P.; Doong, R.A.; Chen, C.W.; Dong, C.D. Visible-light photodegradation of sulfamethoxazole (SMX) over Ag-P-codoped g-C3N4 (Ag-P@UCN) photocatalyst in water. Chem. Eng. J. 2020, 384, 123383. [Google Scholar] [CrossRef]
  22. Zhang, J.J.; Zhu, Z.R.; Jiang, J.C.; Li, H. Fabrication of a novel AgI/LaFeO3/g-C3N4 dual Z-scheme photocatalyst with enhanced photocatalytic performance. Mater. Lett. 2020, 262, 127029. [Google Scholar] [CrossRef]
  23. Liu, X.M.; Liu, Y.; Zhang, W.K.; Zhong, Q.Y.; Ma, X.Y. In situ self-assembly of 3D hierarchical 2D/2D CdS/g-C3N4 hereojunction with excellent photocatalytic performance. Mater. Sci. Semicond. Proc. 2020, 105, 104734. [Google Scholar] [CrossRef]
  24. Nalid, N.R.; Majid, A.; Tahir, M.B.; Niaz, N.A.; Khalid, S. Carbonaceous-TiO2 nanomaterials for photocatalytic degradation of pollutants: A review. Ceram. Int. 2017, 43, 14552–14571. [Google Scholar]
  25. Hafeez, H.Y.; Lakhera, S.K.; Karthik, P.; Anpo, M.; Neppolian, B. Facile construction of ternary CuFe2O4-TiO2 nanocomposite supported reduced graphene oxide (rGO) photocatalysts for the efficient hydrogen production. Appl. Surf. Sci. 2018, 449, 772–779. [Google Scholar] [CrossRef]
  26. Khedr, T.M.; El-Sheikh, S.M.; Ismail, A.A.; Kowalska, E.; Bahnemann, D.W. Photodegradation of Microcystin-LR Using Visible Light-Activated C/N-co-Modified Mesoporous TiO2 Photocatalyst. Materials 2019, 12, 1027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Zhang, S.C.; Lu, X.J. Treatment of wastewater containing Reactive Brilliant Blue KN-R using TiO2/BC composite as heterogeneous photocatalyst and adsorbent. Chemosphere 2018, 206, 777–783. [Google Scholar] [CrossRef]
  28. Li, Y.K.; Zhang, P.; Wan, D.Y.; Xue, C.; Zhao, J.T.; Shao, G.S. Direct evidence of 2D/1D heterojunction enhancement on photocatalytic activity through assembling MoS2 nanosheets onto super-long TiO2 nanofibers. Appl. Surf. Sci. 2020, 504, 144361. [Google Scholar] [CrossRef]
  29. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O’Shea, K.; et al. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Appl. Catal. B Environ. 2012, 125, 331–349. [Google Scholar] [CrossRef] [Green Version]
  30. Liu, W.; Zhao, X.; Borthwick, A.G.L.; Wang, Y.Q.; Ni, J.R. Dual-Enhanced Photocatalytic Activity of Fe-Deposited Titanate Nanotubes Used for Simultaneous Removal of As(III) and As(V). ACS Appl. Mater. Interfaces 2015, 7, 19726–19735. [Google Scholar] [CrossRef] [Green Version]
  31. Cai, Z.; Zhao, X.; Wang, T.; Liu, W.; Zhao, D. Reusable Platinum-Deposited Anatase/Hexa-Titanate Nanotubes: Roles of Reduced and Oxidized Platinum on Enhanced Solar-Light-Driven Photocatalytic Activity. ACS Sustain. Chem. Eng. 2017, 5, 547–555. [Google Scholar] [CrossRef]
  32. Shankar, R.; Shim, W.J.; An, J.G.; Yim, U.H. A practical review on photooxidation of crude oil: Laboratory lamp setup and factors affecting it. Water Res. 2015, 68, 304–315. [Google Scholar] [CrossRef] [PubMed]
  33. Tian, J.; Zhao, Z.H.; Kumar, A.; Boughton, R.I.; Liu, H. Recent progress in design, synthesis, and applications of one-dimensional TiO2 nanostructured surface heterostructures: A review. Chem. Soc. Rev. 2014, 43, 6920–6937. [Google Scholar] [CrossRef] [PubMed]
  34. Dahl, M.; Liu, Y.; Yin, Y. Composite Titanium Dioxide Nanomaterials. Chem. Rev. 2014, 114, 9853–9889. [Google Scholar] [CrossRef]
  35. Hlekelele, L.; Durbach, S.H.; Chauke, V.P.; Dziike, F.; Franklyn, P.J. Resin-gel incorporation of high concentrations of W6+ and Zn2+ into TiO2-anatase crystal to form quaternary mixed-metal oxides: Effect on the a lattice parameter and photodegradation efficiency. RSC Adv. 2019, 9, 36875–36883. [Google Scholar] [CrossRef] [Green Version]
  36. Nzaba, S.K.M.; Nyoni, H.H.; Mamba, B.B.; Kuvarega, A.T. Comparative Study of Dendrimer-Templated Nitrogen-Platinum Co-Doped TiO2 for the Photocatalytic Degradation of Azo Dyes in Contaminated Water. Chemistryselect 2019, 4, 12156–12163. [Google Scholar] [CrossRef]
  37. Sun, D.W.; Li, Y.J.; Cao, T.P.; Zhao, Y.H.; Yang, D.K. Preparation of Dy3+-doped YVO4/TiO2 Composite Nanofibers with Three-dimensional Net-like Structure and Enhanced Photocatalytic Activity for Hydrogen Evolution. Chem. J. Chin. Univ. Chin. 2019, 40, 2348–2353. [Google Scholar] [CrossRef]
  38. Deng, H.Y.; He, H.; Sun, S.J.; Zhu, X.T.; Zhou, D.X.; Han, F.X.; Huang, B.; Pan, X.J. Photocatalytic degradation of dye by Ag/TiO2 nanoparticles prepared with different sol-gel crystallization in the presence of effluent organic matter. Environ. Sci. Pollut. Res. 2019, 26, 35900–35912. [Google Scholar] [CrossRef] [PubMed]
  39. Yoshida, T.; Misu, Y.; Yamamoto, M.; Tanabe, T.; Kumagai, J.; Ogawa, S.; Yagi, S. Effects of the amount of Au nanoparticles on the visible light response of TiO2 photocatalysts. Catal. Today 2020, 352, 34–38. [Google Scholar] [CrossRef]
  40. Yang, L.X.; Chen, Z.; Zhang, J.; Wang, C.A. SrTiO3/TiO2 heterostructure nanowires with enhanced electron-hole separation for efficient photocatalytic activity. Front. Mater. Sci. 2019, 13, 342–351. [Google Scholar] [CrossRef]
  41. Al Jitan, S.; Palmisano, G.; Garlisi, C. Synthesis and Surface Modification of TiO2-Based Photocatalysts for the Conversion of CO2. Catalysts 2020, 10, 227. [Google Scholar] [CrossRef] [Green Version]
  42. Serpone, N. Is the band gap of pristine TiO2 narrowed by anion- and cation-doping of titanium dioxide in second-generation photocatalysts? J. Phys. Chem. B 2006, 110, 24287–24293. [Google Scholar] [CrossRef]
  43. Devi, L.G.; Kavitha, R. A review on non metal ion doped titania for the photocatalytic degradation of organic pollutants under UV/solar light: Role of photogenerated charge carrier dynamics in enhancing the activity. Appl. Catal. B Environ. 2013, 140, 559–587. [Google Scholar] [CrossRef]
  44. Asahi, R.; Morikawa, T.; Irie, H.; Ohwaki, T. Nitrogen-Doped Titanium Dioxide as Visible-Light-Sensitive Photocatalyst: Designs, Developments, and Prospects. Chem. Rev. 2014, 114, 9824–9852. [Google Scholar] [CrossRef] [PubMed]
  45. Liao, F.; Chu, L.F.; Guo, C.X.; Guo, Y.J.; Ke, Q.F.; Guo, Y.P. Ytterbium Doped TiO2 Nanofibers on Activated Carbon Fibers Enhances Adsorption and Photocatalytic Activities for Toluene Removal. Chemistryselect 2019, 4, 9222–9231. [Google Scholar] [CrossRef]
  46. Suarez, S.; Jansson, I.; Ohtani, B.; Sanchez, B. From titania nanoparticles to decahedral anatase particles: Photocatalytic activity of TiO2/zeolite hybrids for VOCs oxidation. Catal. Today 2019, 326, 2–7. [Google Scholar] [CrossRef]
  47. Liang, H.; Wang, Z.Q.; Liao, L.M.; Chen, L.; Li, Z.; Feng, J. High performance photocatalysts: Montmorillonite supported-nano TiO2 composites. Optik 2017, 136, 44–51. [Google Scholar] [CrossRef]
  48. Zhu, P.F.; Ren, Z.H.; Wang, R.X.; Duan, M.; Xie, L.S.; Xu, J.; Tian, Y.J. Preparation and visible photocatalytic dye degradation of Mn-TiO2/sepiolite photocatalysts. Front. Mater. Sci. 2020, 14, 33–42. [Google Scholar] [CrossRef]
  49. Saqib, N.U.; Adnan, R.; Shah, I. Zeolite supported TiO2 with enhanced degradation efficiency for organic dye under household compact fluorescent light. Mater. Res. Express 2019, 6, 095506. [Google Scholar] [CrossRef]
  50. Malakootian, M.; Nasiri, A.; Gharaghani, M.A. Photocatalytic degradation of ciprofloxacin antibiotic by TiO2 nanoparticles immobilized on a glass plate. Chem. Eng. Commun. 2020, 207, 56–72. [Google Scholar] [CrossRef]
  51. Espino-Estevez, M.R.; Fernandez-Rodriguez, C.; Gonzalez-Diaz, O.M.; Navio, J.A.; Fernandez-Hevia, D.; Dona-Rodriguez, J.M. Enhancement of stability and photoactivity of TiO2 coatings on annular glass reactors to remove emerging pollutants from waters. Chem. Eng. J. 2015, 279, 488–497. [Google Scholar] [CrossRef]
  52. Pierpaoli, M.; Lewkowicz, A.; Rycewi’cz, M.; Szczodrowski, K.; Ruello, M.L.; Bogdanowicz, R. Enhanced photocatalytic activity of transparent carbon nanowall/TiO2 heterostructures. Mater. Lett. 2020, 262, 127155. [Google Scholar] [CrossRef]
  53. Chu, Z.D.; Qiu, L.L.; Chen, Y.; Zhuang, Z.S.; Du, P.F.; Xiong, J. TiO2-loaded carbon fiber: Microwave hydrothermal synthesis and photocatalytic activity under UV light irradiation. J. Phys. Chem. Solids 2020, 136, 109138. [Google Scholar] [CrossRef]
  54. Hu, C.Y.; Lei, E.; Hu, K.K.; Lai, L.Y.; Zhao, D.; Zhao, W.; Rong, H. Simple synthesis of 3D flower-like g-C3N4/TiO2 composite microspheres for enhanced visible-light photocatalytic activity. J. Mater. Sci. 2020, 55, 151–162. [Google Scholar] [CrossRef]
  55. Cunha, D.L.; Kuznetsov, A.; Araujo, J.R.; Neves, R.S.; Archanjo, B.S.; Canela, M.C.; Marques, M. Optimization of Benzodiazepine Drugs Removal from Water by Heterogeneous Photocatalysis Using TiO2/Activated Carbon Composite. Water Air Soil Pollut. 2019, 230, 141. [Google Scholar] [CrossRef]
  56. Zhang, Y.Y.; Jiang, Z.L.; Huang, J.Y.; Lim, L.Y.; Li, W.L.; Deng, J.Y.; Gong, D.G.; Tang, Y.X.; Lai, Y.K.; Chen, Z. Titanate and titania nanostructured materials for environmental and energy applications: A review. RSC Adv. 2015, 5, 79479–79510. [Google Scholar] [CrossRef]
  57. Zhao, X.; Cai, Z.Q.; Wang, T.; O’Reilly, S.E.; Liu, W.; Zhao, D.Y. A new type of cobalt-deposited titanate nanotubes for enhanced photocatalytic degradation of phenanthrene. Appl. Catal. B Environ. 2016, 187, 134–143. [Google Scholar] [CrossRef]
  58. Doong, R.-A.; Liao, C.-Y. Enhanced visible-light-responsive photodegradation of bisphenol A by Cu, N-codoped titanate nanotubes prepared by microwave-assisted hydrothermal method. J. Hazard. Mater. 2017, 322, 254–262. [Google Scholar] [CrossRef] [PubMed]
  59. Liu, W.; Sun, W.L.; Borthwick, A.G.L.; Wang, T.; Li, F.; Guan, Y.D. Simultaneous removal of Cr(VI) and 4-chlorophenol through photocatalysis by a novel anatase/titanate nanosheet composite: Synergetic promotion effect and autosynchronous doping. J. Hazard. Mater. 2016, 317, 385–393. [Google Scholar] [CrossRef] [Green Version]
  60. Zhao, X.; Du, P.H.; Cai, Z.Q.; Wang, T.; Fu, J.; Liu, W. Photocatalysis of bisphenol A by an easy-settling titania/titanate composite: Effects of water chemistry factors, degradation pathway and theoretical calculation. Environ. Pollut. 2018, 232, 580–590. [Google Scholar] [CrossRef]
  61. Li, F.; Du, P.H.; Liu, W.; Li, X.S.; Ji, H.D.; Duan, J.; Zhao, D.Y. Hydrothermal synthesis of graphene grafted titania/titanate nanosheets for photocatalytic degradation of 4-chlorophenol: Solar-light-driven photocatalytic activity and computational chemistry analysis. Chem. Eng. J. 2018, 331, 685–694. [Google Scholar] [CrossRef]
  62. Cheng, K.Y.; Cai, Z.Q.; Fu, J.; Sun, X.B.; Sun, W.L.; Chen, L.; Zhang, D.D.; Liu, W. Synergistic adsorption of Cu(II) and photocatalytic degradation of phenanthrene by a jaboticaba-like TiO2/titanate nanotube composite: An experimental and theoretical study. Chem. Eng. J. 2019, 358, 1155–1165. [Google Scholar] [CrossRef]
  63. Ji, H.D.; Du, P.H.; Zhao, D.Y.; Li, S.; Sun, F.B.; Duin, E.C.; Liu, W. 2D/1D graphitic carbon nitride/titanate nanotubes heterostructure for efficient photocatalysis of sulfamethazine under solar light: Catalytic “hot spots” at the rutile-anatase-titanate interfaces. Appl. Catal. B 2020, 263, 118357. [Google Scholar] [CrossRef]
  64. Tu, H.; Li, D.; Yi, Y.; Liu, R.; Wu, Y.; Dong, X.Y.; Shi, X.W.; Deng, H.B. Incorporation of rectorite into porous polycaprolactone/TiO2 nanofibrous mats for enhancing photocatalysis properties towards organic dye pollution. Compos. Commun. 2019, 15, 58–63. [Google Scholar] [CrossRef]
  65. Ni, Y.H.; Yan, K.; Xu, F.Y.; Zhong, W.B.; Zhao, Q.H.; Liu, K.; Yan, K.L.; Wang, D. Synergistic effect on TiO2 doped poly (vinyl alcohol-co-ethylene) nanofibrous film for filtration and photocatalytic degradation of methylene blue. Compos. Commun. 2019, 12, 112–116. [Google Scholar] [CrossRef]
  66. Mohsenzadeh, M.; Mirbagheri, S.A.; Sabbaghi, S. Photocatalytic degradation of 1,2-dichloroethane using immobilized PANI-TiO2 nanocomposite in a pilot-scale packed bed reactor. Desalin. Water Treat. 2019, 155, 72–83. [Google Scholar] [CrossRef] [Green Version]
  67. Qin, Y.Y.; Li, H.; Lu, J.; Meng, F.Y.; Ma, C.C.; Yan, Y.S.; Meng, M.J. Nitrogen-doped hydrogenated TiO2 modified with CdS nanorods with enhanced optical absorption, charge separation and photocatalytic hydrogen evolution. Chem. Eng. J. 2020, 384, 123275. [Google Scholar] [CrossRef]
  68. Ratova, M.; West, G.T.; Kelly, P.J.; Xia, X.; Gao, Y. Synergistic effect of doping with nitrogen and molybdenum on the photocatalytic properties of thin titania films. Vacuum 2015, 114, 205–212. [Google Scholar] [CrossRef]
  69. Kovalevskiy, N.; Selishchev, D.; Svintsitskiy, D.; Selishcheva, S.; Berezin, A.; Kozlov, D. Synergistic effect of polychromatic radiation on visible light activity of N-doped TiO2 photocatalyst. Catal. Commun. 2020, 134, 105841. [Google Scholar] [CrossRef]
  70. Fang, W.Z.; Xing, M.Y.; Zhang, J.L. Modifications on reduced titanium dioxide photocatalysts: A review. J. Photochem. Photobio. C 2017, 32, 21–39. [Google Scholar] [CrossRef]
  71. Reda, S.M.; Khairy, M.; Mousa, M.A. Photocatalytic activity of nitrogen and copper doped TiO2 nanoparticles prepared by microwave-assisted sol-gel process. Arab. J. Chem. 2020, 13, 86–95. [Google Scholar] [CrossRef]
  72. Jin, X.D.; Zhou, X.Q.; Sun, P.; Lin, S.Y.; Cao, W.B.; Li, Z.F.; Liu, W.X. Photocatalytic degradation of norfloxacin using N-doped TiO2: Optimization, mechanism, identification of intermediates and toxicity evaluation. Chemosphere 2019, 237, 124433. [Google Scholar] [CrossRef] [PubMed]
  73. Ji, L.J.; Zhou, S.; Liu, X.; Gong, M.D.; Xu, T. Synthesis of carbon- and nitrogen-doped TiO2/carbon composite fibers by a surface-hydrolyzed PAN fiber and their photocatalytic property. J. Mater. Sci. 2020, 55, 2471–2481. [Google Scholar] [CrossRef]
  74. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  75. Li, H.; Hao, Y.B.; Lu, H.Q.; Liang, L.P.; Wang, Y.Y.; Qiu, J.H.; Shi, X.C.; Wang, Y.; Yao, J.F. systematic study on visible-light N-doped TiO2 photocatalyst obtained from ethylenediamine by sol-gel method. Appl. Surf. Sci. 2015, 344, 112–118. [Google Scholar] [CrossRef]
  76. Jyothi, M.S.; Laveena, P.D.; Shwetharani, R.; Balakrishna, G.R. Novel hydrothermal method for effective doping of N and F into nano Titania for both, energy and environmental applications. Mater. Res. Bull. 2016, 74, 478–484. [Google Scholar] [CrossRef]
  77. Rahbar, M.; Mehrzad, M.; Behpour, M.; Mohammadi-Aghdam, S.; Ashrafi, M. S, N co-doped carbon quantum dots/TiO2 nanocomposite as highly efficient visible light photocatalyst. Nanotechnology 2019, 30, 505702. [Google Scholar] [CrossRef]
  78. Liu, D.; Wang, J.Q.; Zhou, J.; Xi, Q.H.; Li, X.; Nie, E.; Piao, X.Q.; Sun, Z. Fabricating I doped TiO2 photoelectrode for the degradation of diclofenac: Performance and mechanism study. Chem. Eng. J. 2019, 369, 968–978. [Google Scholar] [CrossRef]
  79. Ichihara, F.; Sieland, F.; Pang, H.; Philo, D.; Duong, A.T.; Chang, K.; Kako, T.; Bahnemann, D.W.; Ye, J.H. Photogenerated Charge Carriers Dynamics on La- and/or Cr-Doped SrTiO3 Nanoparticles Studied by Transient Absorption Spectroscopy. J. Phys. Chem. C 2020, 124, 1292–1302. [Google Scholar] [CrossRef]
  80. Tian, L.J.; Xing, L.; Shen, X.L.; Li, Q.H.; Ge, S.J.; Liu, B.K.; Jie, L. Visible light enhanced Fe-I-TiO2 photocatalysts for the degradation of gaseous benzene. Atmos. Pollut. Res. 2020, 11, 179–185. [Google Scholar] [CrossRef]
  81. Razali, M.H.; Noor, A.F.M.; Yusoff, M. Physicochemical Properties of a Highly Efficient Cu-Ion-Doped TiO2 Nanotube Photocatalyst for the Degradation of Methyl Orange Under Sunlight. J. Nanosci. Nanotechnol. 2020, 20, 965–972. [Google Scholar] [CrossRef] [PubMed]
  82. Li, Z.; Tian, B.; Zhen, W.L.; Wu, Y.Q.; Lu, G.X. Inhibition of hydrogen and oxygen recombination using oxygen transfer reagent hemin chloride in Pt/TiO2 dispersion for photocatalytic hydrogen generation. Appl. Catal. B 2017, 203, 408–415. [Google Scholar] [CrossRef]
  83. Yu, Y.Q.; Zhu, X.R.; Wang, L.R.; Wu, F.S.; Liu, S.L.; Chang, C.Y.; Luo, X.G. A simple strategy to design 3-layered Au-TiO2 dual nanoparticles immobilized cellulose membranes with enhanced photocatalytic activity. Carbohydr. Polym. 2020, 231, 115694. [Google Scholar] [CrossRef] [PubMed]
  84. Surya, C.; John, N.A.A.; Pandiyan, V.; Ravikumar, S.; Amutha, P.; Sobral, A.J.F.N.; Krishnakumar, B. Costus speciosus leaf extract assisted CS-Pt-TiO2 composites: Synthesis, characterization and their bio and photocatalytic applications. J. Mol. Struct. 2019, 1195, 787–795. [Google Scholar] [CrossRef]
  85. Maarisetty, D.; Baral, S.S. Defect-induced enhanced dissociative adsorption, optoelectronic properties and interfacial contact in Ce doped TiO2: Solar photocatalytic degradation of Rhodamine B. Ceram Int. 2019, 45, 22253–22263. [Google Scholar] [CrossRef]
  86. Zhao, J.; Chen, X.Y.; Zhou, Y.H.; Tian, H.J.; Guo, Q.J.; Hu, X.D. Efficient removal of oil pollutant via simultaneous adsorption and photocatalysis using La-N-TiO2-cellulose/SiO2 difunctional aerogel composite. Res. Chem. Intermed. 2020, 46, 1805–1822. [Google Scholar] [CrossRef]
  87. Liang, J.C.; Wang, J.Y.; Song, K.X.; Wang, X.F.; Yu, K.F.; Liang, C. Enhanced photocatalytic activities of Nd-doped TiO2 under visible light using a facile sol-gel method. J. Rare Earth 2020, 38, 148–156. [Google Scholar] [CrossRef]
  88. Hu, Z.; Yang, C.; Lv, K.; Li, X.; Li, Q.; Fan, J. Single atomic Au induced dramatic promotion of the photocatalytic activity of TiO2 hollow microspheres. Chem. Commun. Camb. Engl. 2020, 56, 1745–1748. [Google Scholar] [CrossRef]
  89. Hayashi, T.; Nakamura, K.; Suzuki, T.; Saito, N.; Murakami, Y. OH radical formation by the photocatalytic reduction reactions of H2O2 on the surface of plasmonic excited Au-TiO2 photocatalysts. Chem. Phys. Lett. 2020, 739, 136958. [Google Scholar] [CrossRef]
  90. Jeantelot, G.; Qureshi, M.; Harb, M.; Ould-Chikh, S.; Anjum, D.H.; Abou-Hamad, E.; Aguilar-Tapia, A.; Hazemann, J.L.; Takanabe, K.; Basset, J.M. TiO2-supported Pt single atoms by surface organometallic chemistry for photocatalytic hydrogen evolution. Phys. Chem. Chem. Phys. 2019, 21, 24429–24440. [Google Scholar] [CrossRef] [Green Version]
  91. Singhal, N.; Kumar, U. Noble metal modified TiO2: Selective photoreduction of CO2 to hydrocarbons. Mol. Catal. 2017, 439, 91–99. [Google Scholar] [CrossRef]
  92. Gao, P.; Yang, L.B.; Xiao, S.T.; Wang, L.Y.; Guo, W.; Lu, J.H. Effect of Ru, Rh, Mo, and Pd Adsorption on the Electronic and Optical Properties of Anatase TiO2(101): A DFT Investigation. Materials 2019, 12, 814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Liu, Y.; Xiao, Z.Z.; Cao, S.; Li, J.H.; Piao, L.Y. Controllable synthesis of Au-TiO2 nanodumbbell photocatalysts with spatial redox region. Chin. J. Catal. 2020, 41, 219–226. [Google Scholar] [CrossRef]
  94. Khatun, F.; Abd Aziz, A.; Sim, L.C.; Monir, M.U. Plasmonic enhanced Au decorated TiO2 nanotube arrays as a visible light active catalyst towards photocatalytic CO2 conversion to CH4. J. Environ. Chem. Eng. 2019, 7. [Google Scholar] [CrossRef]
  95. Sacco, O.; Murcia, J.J.; Lara, A.E.; Hernandez-Laverde, M.; Rojas, H.; Navio, J.A.; Hidalgo, M.C.; Vaiano, V. Pt-TiO2-Nb2O5 heterojunction as effective photocatalyst for the degradation of diclofenac and ketoprofen. Mater. Sci. Semicond. Proc. 2020, 107, 104839. [Google Scholar] [CrossRef]
  96. Qu, J.F.; Chen, D.Y.; Li, N.J.; Xu, Q.F.; Li, H.; He, J.H.; Lu, J.M. Ternary photocatalyst of atomic-scale Pt coupled with MoS2 co-loaded on TiO2 surface for highly efficient degradation of gaseous toluene. Appl. Catal. B 2019, 256. [Google Scholar] [CrossRef]
  97. Ouyang, W.Y.; Munoz-Batista, M.J.; Kubacka, A.; Luque, R.; Fernandez-Garcia, M. Enhancing photocatalytic performance of TiO2 in H2 evolution via Ru co-catalyst deposition. Appl. Catal B Environ. 2018, 238, 434–443. [Google Scholar] [CrossRef]
  98. Li, F.; Huang, H.B.; Li, G.S.; Leung, D.Y.C. TiO2 nanotube arrays modified with nanoparticles of platinum group metals (Pt, Pd, Ru): Enhancement on photoelectrochemical performance. J. Nanopart Res. 2019, 21, 29. [Google Scholar] [CrossRef]
  99. Sato, S.; White, J.M. Photodecomposition of water over Pt/TiO2 catalysts. Chem. Phys. Lett. 1980, 72, 83–86. [Google Scholar] [CrossRef]
  100. Kennedy, J.C.; Datye, A.K. Photothermal heterogeneous oxidation of ethanol over Pt/TiO2. J. Catal. 1998, 179, 375–389. [Google Scholar] [CrossRef]
  101. Ji, L.J.; Qin, X.; Zheng, J.J.; Zhou, S.; Xu, T.; Shi, G.J. Synthesis of Ag-Carbon-TiO2 composite tubes and their antibacterial and organic degradation properties. J. Sol Gel Sci. Technol. 2020, 93, 291–301. [Google Scholar] [CrossRef]
  102. Shan, R.; Lu, L.L.; Gu, J.; Zhang, Y.Y.; Yuan, H.R.; Chen, Y.; Luo, B. Photocatalytic degradation of methyl orange by Ag/TiO2/biochar composite catalysts in aqueous solutions. Mater. Sci. Semicond. Process. 2020, 114, 105088. [Google Scholar] [CrossRef]
  103. Jaafar, F.; Jalil, A.A.; Triwahyono, S.; Efendi, J.; Mukti, R.R.; Jusoh, R.; Jusoh, N.W.C.; Karim, A.H.; Salleh, N.F.M.; Suendo, V. Direct in situ activation of Ag0 nanoparticles in synthesis of Ag/TiO2 and its photoactivity. Appl. Surf. Sci. 2015, 338, 75–84. [Google Scholar] [CrossRef]
  104. Zhu, D.H.; Long, L.; Sun, J.Y.; Wan, H.Q.; Zheng, S.R. Highly active and selective catalytic hydrogenation of p-chloronitrobenzene to p-chloroaniline on Pt@Cu/TiO2. Appl. Surf. Sci. 2020, 504, 144329. [Google Scholar] [CrossRef]
  105. Nyankson, E.; Agyei-Tuffour, B.; Adjasoo, J.; Ebenezer, A.; Dodoo-Arhin, D.; Yaya, A.; Mensah, B.; Efavi, J.K. Synthesis and Application of Fe-Doped TiO2-Halloysite Nanotubes Composite and Their Potential Application in Water Treatment. Adv. Mater. Sci. Eng. 2019, 2019, 4270310. [Google Scholar] [CrossRef] [Green Version]
  106. Komaraiah, D.; Radha, E.; Sivakumar, J.; Reddy, M.V.R.; Sayanna, R. Structural, optical properties and photocatalytic activity of Fe3+ doped TiO2 thin films deposited by sol-gel spin coating. Surf. Interfaces 2019, 17, 100368. [Google Scholar] [CrossRef]
  107. Bhardwaj, S.; Dogra, D.; Pal, B.; Singh, S. Photodeposition time dependant growth, size and photoactivity of Ag and Cu deposited TiO2 nanocatalyst under solar irradiation. Sol. Energy 2019, 194, 618–627. [Google Scholar] [CrossRef]
  108. Ghanbari, S.; Givianrad, M.H.; Azar, P.A. Synthesis and characterization of visible light driven N-Fe-codoped TiO2/SiO2 for simultaneous photoremoval of Cr (VI) and azo dyes in a novel fixed bed continuous flow photoreactor. Can. J. Chem. Eng. 2020, 98, 705–716. [Google Scholar] [CrossRef]
  109. Ahadi, S.; Moalej, N.S.; Sheibani, S. Characteristics and photocatalytic behavior of Fe and Cu doped TiO2 prepared by combined sol-gel and mechanical alloying. Solid State Sci. 2019, 96, 105975. [Google Scholar] [CrossRef]
  110. Zhang, J.H.; Fu, D.; Wang, S.Q.; Hao, R.L.; Xie, Y.X. Photocatalytic removal of chromium(VI) and sulfite using transition metal (Cu, Fe, Zn) doped TiO2 driven by visible light: Feasibility, mechanism and kinetics. J. Ind. Eng. Chem. 2019, 80, 23–32. [Google Scholar] [CrossRef]
  111. Khan, M.A.M.; Siwach, R.; Kumar, S.; Alhazaa, A.N. Role of Fe doping in tuning photocatalytic and photoelectrochemical properties of TiO2 for photodegradation of methylene blue. Opt. Laser Technol. 2019, 118, 170–178. [Google Scholar] [CrossRef]
  112. Bhatia, V.; Dhir, A. Transition metal doped TiO2 mediated photocatalytic degradation of anti-inflammatory drug under solar irradiations. J. Environ. Chem. Eng. 2016, 4, 1267–1273. [Google Scholar] [CrossRef]
  113. Stucchi, M.; Boffito, D.C.; Pargoletti, E.; Cerrato, G.; Bianchi, C.L.; Cappelletti, G. Nano-MnO2 Decoration of TiO2 Microparticles to Promote Gaseous Ethanol Visible Photoremoval. Nanomaterials 2018, 8, 686. [Google Scholar] [CrossRef] [Green Version]
  114. Choi, W.; Termin, A.; Hoffmann, M.R. The Role of Metal Ion Dopants in Quantum-Sized TiO2: Correlation between Photoreactivity and Charge Carrier Recombination Dynamics. J. Phys. Chem. 1994, 98, 13669–13679. [Google Scholar] [CrossRef]
  115. Du, X.-Q.; Ma, X.-L.; Wang, Y.-J.; Gu, F.-N.; Zhang, J. Optimization of the Ge/TiO2 catalyst for the degradation of ciprofloxacin by the response surface methodology. Zhongguo Kangshengsu Zazhi 2019, 44, 750–757. [Google Scholar]
  116. Crisan, M.; Mardare, D.; Ianculescu, A.; Dragan, N.; Nitoi, I.; Crisan, D.; Voicescu, M.; Todan, L.; Oancea, P.; Adomnitei, C.; et al. Iron doped TiO2 films and their photoactivity in nitrobenzene removal from water. Appl. Surf. Sci. 2018, 455, 201–215. [Google Scholar] [CrossRef]
  117. Gnanasekaran, L.; Hemamalini, R.; Saravanan, R.; Ravichandran, K.; Gracia, F.; Gupta, V.K. Intermediate state created by dopant ions (Mn, Co and Zr) into TiO2 nanoparticles for degradation of dyes under visible light. J. Mol. Liq. 2016, 223, 652–659. [Google Scholar] [CrossRef]
  118. Huang, J.G.; Guo, X.T.; Wang, B.; Li, L.Y.; Zhao, M.X.; Dong, L.L.; Liu, X.J.; Huang, Y.T. Synthesis and Photocatalytic Activity of Mo-Doped TiO2 Nanoparticles. J. Spectrosc. 2015, 2015, 681850. [Google Scholar] [CrossRef] [Green Version]
  119. Wang, Q.; Jin, R.; Zhang, M.; Gao, S. Solvothermal preparation of Fe-doped TiO2 nanotube arrays for enhancement in visible light induced photoelectrochemical performance. J. Alloys Compd. 2017, 690, 139–144. [Google Scholar] [CrossRef]
  120. Vinodgopal, K.; Kamat, P.V. Enhanced Rates of Photocatalytic Degradation of an Azo Dye Using SnO2/TiO2 Coupled Semiconductor Thin Films. Environ. Sci. Technol. 1995, 29, 841–845. [Google Scholar] [CrossRef]
  121. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction Photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef] [PubMed]
  122. Kaur, P.K.; Badru, R.; Singh, P.P.; Kaushal, S. Photodegradation of organic pollutants using heterojunctions: A review. J. Environ. Chem. Eng. 2020, 8, 103666. [Google Scholar] [CrossRef]
  123. Li, H.J.; Zhou, Y.; Tu, W.G.; Ye, J.H.; Zou, Z.G. State-of-the-Art Progress in Diverse Heterostructured Photocatalysts toward Promoting Photocatalytic Performance. Adv. Funct. Mater. 2015, 25, 998–1013. [Google Scholar] [CrossRef]
  124. Dursun, S.; Kaya, I.C.; Kalem, V.; Akyildiz, H. UV/visible light active CuCrO2 nanoparticle-SnO2 nanofiber p-n heterostructured photocatalysts for photocatalytic applications. Dalton Trans. 2018, 47, 14662–14678. [Google Scholar] [CrossRef] [PubMed]
  125. Zhang, W.P.; Xiao, X.Y.; Li, Y.; Zeng, X.Y.; Zheng, L.L.; Wan, C.X. Liquid-exfoliation of layered MoS2 for enhancing photocatalytic activity of TiO2/g-C3N4 photocatalyst and DFT study. Appl. Surf. Sci. 2016, 389, 496–506. [Google Scholar] [CrossRef]
  126. Yang, L.; Gao, M.G.; Dai, B.; Guo, X.H.; Liu, Z.Y.; Peng, B.H. Synthesis of spindle-shaped AgI/TiO2 nanoparticles with enhanced photocatalytic performance. Appl. Surf. Sci. 2016, 386, 337–344. [Google Scholar] [CrossRef]
  127. Demirci, S.; Dikici, T.; Yurddaskal, M.; Gultekin, S.; Toparli, M.; Celik, E. Synthesis and characterization of Ag doped TiO2 heterojunction films and their photocatalytic performances. Appl. Surf. Sci. 2016, 390, 591–601. [Google Scholar] [CrossRef]
  128. Sarkar, A.; Gracia-Espino, E.; Wagberg, T.; Shchukarev, A.; Mohl, M.; Rautio, A.R.; Pitkanen, O.; Sharifi, T.; Kordas, K.; Mikkola, J.P. Photocatalytic reduction of CO2 with H2O over modified TiO2 nanofibers: Understanding the reduction pathway. Nano Res. 2016, 9, 1956–1968. [Google Scholar] [CrossRef]
  129. Low, J.X.; Cheng, B.; Yu, J.G. Surface modification and enhanced photocatalytic CO2 reduction performance of TiO2: A review. Appl. Surf. Sci. 2017, 392, 658–686. [Google Scholar] [CrossRef]
  130. Moniz, S.J.A.; Shevlin, S.A.; Martin, D.J.; Guo, Z.X.; Tang, J.W. Visible-light driven heterojunction photocatalysts for water splitting—A critical review. Energy Environ. Sci. 2015, 8, 731–759. [Google Scholar] [CrossRef]
  131. Kumar, A.; Khan, M.; He, J.H.; Lo, I.M.C. 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] [PubMed]
  132. Ong, W.J.; Tan, L.L.; Ng, Y.H.; Yong, S.T.; Chai, S.P. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer To Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef] [PubMed]
  133. Ahmed, S.N.; Haider, W. Heterogeneous photocatalysis and its potential applications in water and wastewater treatment: A review. Nanotechnology 2018, 29, 342001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. McDaniel, H.; Heil, P.E.; Tsai, C.L.; Kim, K.; Shim, M. Integration of Type II Nanorod Heterostructures into Photovoltaics. ACS Nano 2011, 5, 7677–7683. [Google Scholar] [CrossRef]
  135. Gou, J.F.; Ma, Q.L.; Deng, X.Y.; Cui, Y.Q.; Zhang, H.X.; Cheng, X.W.; Li, X.L.; Xie, M.Z.; Cheng, Q.F. Fabrication of Ag2O/TiO2-Zeolite composite and its enhanced solar light photocatalytic performance and mechanism for degradation of norfloxacin. Chem. Eng. J. 2017, 308, 818–826. [Google Scholar] [CrossRef]
  136. Wang, W.; Fang, J.J.; Shao, S.F.; Lai, M.; Lu, C.H. Compact and uniform TiO2@g-C3N4 core-shell quantum heterojunction for photocatalytic degradation of tetracycline antibiotics. Appl. Catal. B 2017, 217, 57–64. [Google Scholar] [CrossRef]
  137. Hou, H.J.; Zhang, X.H.; Huang, D.K.; Ding, X.; Wang, S.Y.; Yang, X.L.; Li, S.Q.; Xiang, Y.G.; Chen, H. Conjugated microporous poly(benzothiadiazole)/TiO2 heterojunction for visible-light-driven H2 production and pollutant removal. Appl. Catal. B 2017, 203, 563–571. [Google Scholar] [CrossRef]
  138. Ganguly, P.; Mathew, S.; Clarizia, L.; Kumar, R.S.; Akande, A.; Hinder, S.; Breen, A.; Pillai, S.C. Theoretical and experimental investigation of visible light responsive AgBiS2-TiO2 heterojunctions for enhanced photocatalytic applications. Appl. Catal. B 2019, 253, 401–418. [Google Scholar] [CrossRef]
  139. Liu, Z.; Tian, J.; Zeng, D.B.; Yu, C.L.; Huang, W.Y.; Yang, K.; Liu, X.Q.; Liu, H. Binary-phase TiO2 modified Bi2MoO6 crystal for effective removal of antibiotics under visible light illumination. Mater. Res. Bull. 2019, 112, 336–345. [Google Scholar] [CrossRef]
  140. Yu, J.G.; Wang, S.H.; Low, J.X.; Xiao, W. Enhanced photocatalytic performance of direct Z-scheme g-C3N4-TiO2 photocatalysts for the decomposition of formaldehyde in air. Phys. Chem. Chem. Phys. 2013, 15, 16883–16890. [Google Scholar] [CrossRef]
  141. Mei, Q.F.; Zhang, F.Y.; Wang, N.; Lu, W.S.; Su, X.T.; Wang, W.; Wu, R.L. Photocatalysts: Z-Scheme Heterojunction Constructed with Titanium Dioxide. Chin. J. Inorg. Chem. 2019, 35, 1321–1339. [Google Scholar] [CrossRef]
  142. Li, H.J.; Tu, W.G.; Zhou, Y.; Zou, Z.G. Z-Scheme Photocatalytic Systems for Promoting Photocatalytic Performance: Recent Progress and Future Challenges. Adv. Sci. 2016, 3, 1500389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Qi, K.Z.; Cheng, B.; Yu, J.G.; Ho, W.K. A review on TiO2-based Z-scheme photocatalysts. Chin. J. Catal. 2017, 38, 1936–1955. [Google Scholar] [CrossRef]
  144. Low, J.X.; Jiang, C.; Cheng, B.; Wageh, S.; Al-Ghamdi, A.A.; Yu, J.G. A Review of Direct Z-Scheme Photocatalysts. Small Methods 2017, 1, 1700080. [Google Scholar] [CrossRef]
  145. Meng, S.G.; Zhang, J.F.; Chen, S.F.; Zhang, S.J.; Huang, W.X. Perspective on construction of heterojunction photocatalysts and the complete utilization of photogenerated charge carriers. Appl. Surf. Sci. 2019, 476, 982–992. [Google Scholar] [CrossRef]
  146. Wang, F.L.; Wu, Y.L.; Wang, Y.F.; Li, J.H.; Jin, X.Y.; Zhang, Q.X.; Li, R.B.; Yan, S.C.; Liu, H.J.; Feng, Y.P.; et al. Construction of novel Z-scheme nitrogen-doped carbon dots/{001} TiO2 nanosheet photocatalysts for broad-spectrum-driven diclofenac degradation: Mechanism insight, products and effects of natural water matrices. Chem. Eng. J. 2019, 356, 857–868. [Google Scholar] [CrossRef]
  147. Hao, J.G.; Zhang, S.F.; Ren, F.; Wang, Z.W.; Lei, J.F.; Wang, X.N.; Cheng, T.; Li, L.B. Synthesis of TiO2@g-C3N4 core-shell nanorod arrays with Z-scheme enhanced photocatalytic activity under visible light. J. Colloid Interface Sci. 2017, 508, 419–425. [Google Scholar] [CrossRef]
  148. Liao, W.J.; Murugananthan, M.; Zhang, Y.R. Synthesis of Z-scheme g-C3N4-Ti3+/TiO2 material: An efficient visible light photoelectrocatalyst for degradation of phenol. Phys. Chem. Chem. Phys. 2015, 17, 8877–8884. [Google Scholar] [CrossRef]
  149. 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]
  150. 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/codoped TiO2 and TiO2 nanostructured hybrids for enhanced photocatalysis. Int. J. Hydrog. Energy 2020, 45, 7764–7778. [Google Scholar] [CrossRef]
  151. Li, S.Y.; Yang, Y.L.; Su, Q.; Liu, X.Y.; Zhao, H.P.; Zhao, Z.X.; Li, J.; Jin, C. Synthesis and photocatalytic activity of transition metal and rare earth element co-doped TiO2 nano particles. Mater. Lett. 2019, 252, 123–125. [Google Scholar] [CrossRef]
  152. Oladipo, G.O.; Akinlabi, A.K.; Alayande, S.O.; Msagati, T.A.M.; Nyoni, H.H.; Ogunyinka, O.O. Synthesis, characterization, and photocatalytic activity of silver and zinc co-doped TiO2 nanoparticle for photodegradation of methyl orange dye in aqueous solution. Can. J. Chem. 2019, 97, 642–650. [Google Scholar] [CrossRef]
  153. Shahan, M.; Ahmed, A.M.; Shehata, N.; Betiha, M.A.; Rabie, A.M. Ni-doped and Ni/Cr co-doped TiO2 nanotubes for enhancement of photocatalytic degradation of methylene blue. J. Colloid Interface Sci. 2019, 555, 31–41. [Google Scholar] [CrossRef]
  154. Xia, H.Y.; Liu, G.Y.; Zhang, R.; Song, L.F.; Chen, H.X. The Photocatalytic Degradation of Vehicle Exhausts by an Fe/N/Co-TiO2 Waterborne Coating under Visible Light. Materials 2019, 12, 3378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Sun, M.X.; Yao, Y.; Ding, W.; Anandan, S. N/Ti3+ co-doping biphasic TiO2/Bi2WO6 heterojunctions: Hydrothermal fabrication and sonophotocatalytic degradation of organic pollutants. J. Alloys Compd. 2020, 820, 153172. [Google Scholar] [CrossRef]
  156. Bayan, E.M.; Lupeiko, T.G.; Pustovaya, L.E.; Volkova, M.G.; Butova, V.V.; Guda, A.A. Zn-F co-doped TiO2 nanomaterials: Synthesis, structure and photocatalytic activity. J. Alloys Compd. 2020, 822, 153662. [Google Scholar] [CrossRef]
  157. Zhao, L.; Xie, Y.Z.; Lin, Q.Y.; Zheng, R.Z.; Diao, Y. Preparation of C, N and P co-doped TiO2 and its photocatalytic activity under visible light. Funct. Mater. Lett. 2019, 12, 1950045. [Google Scholar] [CrossRef]
  158. Wang, H.X.; Zhu, L.N.; Guo, F.Q. Photoelectrocatalytic degradation of atrazine by boron-fluorine co-doped TiO2 nanotube arrays. Environ. Sci. Pollut. Res. 2019, 26, 33847–33855. [Google Scholar] [CrossRef]
  159. Yan, J.C.; Zhao, J.; Hao, L.; Hu, Y.F.; Liu, T.Y.; Guan, S.J.; Zhao, Q.; Zhu, Z.; Lu, Y. Low-temperature S-doping on N-doped TiO2 films and remarkable enhancement on visible-light performance. Mater. Res. Bull. 2019, 120, 110594. [Google Scholar] [CrossRef]
  160. Mohamed, R.M.; Bahnemann, D.W.; Basaleh, A.S.; Qadah, R.H. Photo-catalytic destruction of acetaldehyde using cobalt, copper co-doped titania dioxide nanoparticles beneath Visible light. Appl. Nanosci. 2020, 10, 931–939. [Google Scholar] [CrossRef]
  161. Pham, T.D.; Lee, B.K. Selective removal of polar VOCs by novel photocatalytic activity of metals co-doped TiO2/PU under visible light. Chem. Eng. J. 2017, 307, 63–73. [Google Scholar] [CrossRef]
  162. Isari, A.A.; Hayati, F.; Kakavandi, B.; Rostami, M.; Motevassel, M.; Dehghanifard, E. N, Cu co-doped TiO2@functionalized SWCNT photocatalyst coupled with ultrasound and visible-light: An effective sono-photocatalysis process for pharmaceutical wastewaters treatment. Chem. Eng. J. 2020, 392, 123685. [Google Scholar] [CrossRef]
  163. Sharotri, N.; Sharma, D.; Sud, D. Experimental and theoretical investigations of Mn-N-co-doped TiO2 photocatalyst for visible light induced degradation of organic pollutants. J. Mater. Res. Technol. 2019, 8, 3995–4009. [Google Scholar] [CrossRef]
  164. Sirivallop, A.; Areerob, T.; Chiarakorn, S. Enhanced Visible Light Photocatalytic Activity of N and Ag Doped and Co-Doped TiO2 Synthesized by Using an In-Situ Solvothermal Method for Gas Phase Ammonia Removal. Catalysts 2020, 10, 251. [Google Scholar] [CrossRef] [Green Version]
  165. Chauhan, N.; Singh, V.; Kumar, S.; Sirohi, K. Preparation of palladium, silver, and nitrogen co-doped mesoporous titanium dioxide nanoparticles to investigate their photocatalytic action. Mater. Res. Express 2019, 6, 13. [Google Scholar] [CrossRef]
  166. Osin, O.A.; Yu, T.Y.; Cai, X.M.; Jiang, Y.; Peng, G.T.; Cheng, X.M.; Li, R.B.; Qin, Y.; Lin, S.J. Photocatalytic Degradation of 4-Nitrophenol by C, N-TiO2: Degradation Efficiency vs. Embryonic Toxicity of the Resulting Compounds. Front. Chem. 2018, 6, 192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Zhao, Z.Y.; Feng, M.C.; Peng, Z.J.; Huang, H.W.; Guo, Z.H.; Li, Z.H. Molten-salt fabrication of (N,F)-codoped single-crystal-like titania with high exposure of (001) crystal facet for highly efficient degradation of methylene blue under visible light irradiation. J. Mater. Res. 2018, 33, 1411–1421. [Google Scholar] [CrossRef] [Green Version]
  168. Wang, P.S.; Qi, C.X.; Wen, P.C.; Hao, L.Y.; Xu, X.; Agathopoulos, S. Synthesis of Si, N co-Doped Nano-Sized TiO2 with High Thermal Stability and Photocatalytic Activity by Mechanochemical Method. Nanomaterials 2018, 8, 294. [Google Scholar] [CrossRef] [Green Version]
  169. Zhang, J.Q.; Xing, Z.P.; Cui, J.Y.; Li, Z.Z.; Tan, S.Y.; Yin, J.W.; Zou, J.L.; Zhu, Q.; Zhou, W. C,N co-doped porous TiO2 hollow sphere visible light photocatalysts for efficient removal of highly toxic phenolic pollutants. Dalton Trans. 2018, 47, 4877–4884. [Google Scholar] [CrossRef] [PubMed]
  170. Singh, I.; Birajdar, B. Effective La-Na Co-Doped TiO2 Nano-Particles for Dye Adsorption: Synthesis, Characterization and Study on Adsorption Kinetics. Nanomaterials 2019, 9, 400. [Google Scholar] [CrossRef] [Green Version]
  171. Li, Y.F.; Xu, D.H.; Oh, J.I.; Shen, W.; Li, X.; Yu, Y. Mechanistic Study of Codoped Titania with Nonmetal and Metal Ions: A Case of C plus Mo Codoped TiO2. ACS Catal. 2012, 2, 391–398. [Google Scholar] [CrossRef]
  172. Garg, A.; Singhania, T.; Singh, A.; Sharma, S.; Rani, S.; Neogy, A.; Yadav, S.R.; Sangal, V.K.; Garg, N. Photocatalytic Degradation of Bisphenol-A using N, Co Codoped TiO2 Catalyst under Solar Light. Sci. Rep. 2019, 9, 765. [Google Scholar] [CrossRef]
  173. Wang, F.; Ma, Z.Z.; Ban, P.P.; Xu, X.H. C, N and S codoped rutile TiO2 nanorods for enhanced visible-light photocatalytic activity. Mater. Lett. 2017, 195, 143–146. [Google Scholar] [CrossRef]
  174. Wang, M.G.; Han, J.; Hu, Y.M.; Guo, R. Mesoporous C, N-codoped TiO2 hybrid shells with enhanced visible light photocatalytic performance. RSC Adv. 2017, 7, 15513–15520. [Google Scholar] [CrossRef] [Green Version]
  175. Modanlu, S.; Shafiekhani, A. Synthesis of pure and C/S/N co-doped Titania on Al mesh and their photocatalytic usage in Benzene degradation. Sci. Rep. 2019, 9, 16648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Zeng, X.L.; Sun, X.Z.; Yu, Y.S.; Wang, H.Y.; Wang, Y. Photocatalytic degradation of flumequine with B/N codoped TiO2 catalyst: Kinetics, main active species, intermediates and pathways. Chem. Eng. J. 2019, 378, 122226. [Google Scholar] [CrossRef]
  177. Humayun, M.; Raziq, F.; Khan, A.; Luo, W. Modification strategies of TiO2 for potential applications in photocatalysis: Acritical review. Green Chem. Lett. Rev. 2018, 11, 86–102. [Google Scholar] [CrossRef] [Green Version]
  178. Zhu, K.; Neale, N.R.; Miedaner, A.; Frank, A.J. Enhanced charge-collection efficiencies and light scattering in dye-sensitized solar cells using oriented TiO2 nanotubes arrays. Nano Lett. 2007, 7, 69–74. [Google Scholar] [CrossRef]
  179. Zangeneh, H.; Zinatizadeh, A.A.L.; Habibi, M.; Akia, M.; Isa, M.H. Photocatalytic oxidation of organic dyes and pollutants in wastewater using different modified titanium dioxides: A comparative review. J. Ind. Eng. Chem. 2015, 26, 1–36. [Google Scholar] [CrossRef]
  180. Toumazatou, A.; Arfanis, M.K.; Pantazopoulos, P.A.; Kontos, A.G.; Falaras, P.; Stefanou, N.; Likodimos, V. Slow-photon enhancement of dye sensitized TiO2 photocatalysis. Mater. Lett. 2017, 197, 123–126. [Google Scholar] [CrossRef]
  181. Diaz-Angulo, J.; Lara-Ramos, J.; Mueses, M.; Hernandez-Ramirez, A.; Li Puma, G.; Machuca-Martinez, F. Enhancement of the oxidative removal of diclofenac and of the TiO2 rate of photon absorption in dye-sensitized solar pilot scale CPC photocatalytic reactors. Chem. Eng. J. 2020, 381, 122520. [Google Scholar] [CrossRef]
  182. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface modification of TiO2 photocatalyst for environmental applications. J. Photochem. Photobiol. C Photochem. Rev. 2013, 15, 1–20. [Google Scholar] [CrossRef]
  183. Choi, S.K.; Yang, H.S.; Kim, J.H.; Park, H. Organic dye-sensitized TiO2 as a versatile photocatalyst for solar hydrogen and environmental remediation. Appl. Catal. B 2012, 121, 206–213. [Google Scholar] [CrossRef]
  184. Dong, S.Y.; Feng, J.L.; Fan, M.H.; Pi, Y.Q.; Hu, L.M.; Han, X.; Liu, M.L.; Sun, J.Y.; Sun, J.H. Recent developments in heterogeneous photocatalytic water treatment using visible light-responsive photocatalysts: A review. RSC Adv. 2015, 5, 14610–14630. [Google Scholar] [CrossRef]
  185. Diaz-Angulo, J.; Gomez-Bonilla, I.; Jimenez-Tohapanta, C.; Mueses, M.; Pinzon, M.; Machuca-Martinez, F. Visible-light activation of TiO2 by dye-sensitization for degradation of pharmaceutical compounds. Photochem. Photobiol. Sci. 2019, 18, 897–904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Yun, E.T.; Yoo, H.Y.; Kim, W.; Kim, H.E.; Kang, G.; Lee, H.; Lee, S.; Park, T.; Lee, C.; Kim, J.H.; et al. Visible-light-induced activation of periodate that mimics dye-sensitization of TiO2: Simultaneous decolorization of dyes and production of oxidizing radicals. Appl. Catal. B 2017, 203, 475–484. [Google Scholar] [CrossRef]
  187. Sengupta, D.; Das, P.; Mondal, B.; Mukherjee, K. Effects of doping, morphology and film-thickness of photo-anode materials for dye sensitized solar cell application—A review. Renew. Sust. Energ. Rev. 2016, 60, 356–376. [Google Scholar] [CrossRef]
  188. Ahmad, I.; Kan, C.W. Visible-Light-Driven, Dye-Sensitized TiO2 Photo-Catalyst for Self-Cleaning Cotton Fabrics. Coatings 2017, 7, 192. [Google Scholar] [CrossRef] [Green Version]
  189. Murcia, J.J.; Avila-Martinez, E.G.; Rojas, H.; Cubillos, J.; Ivanova, S.; Penkova, A.; Laguna, O.H. Powder and Nanotubes Titania Modified by Dye Sensitization as Photocatalysts for the Organic Pollutants Elimination. Nanomaterials 2019, 9, 517. [Google Scholar] [CrossRef] [Green Version]
  190. Guo, G.Q.; Guo, H.Y.; Wang, F.; France, L.J.; Yang, W.X.; Mei, Z.H.; Yu, Y.H. Dye-sensitized TiO2@SBA-15 composites: Preparation and their application in photocatalytic desulfurization. Green Energy Environ. 2020, 5, 114–120. [Google Scholar] [CrossRef]
  191. Xu, T.Z.; Zhao, H.C.; Zheng, H.; Zhang, P.Y. Atomically Pt implanted nanoporous TiO2 film for photocatalytic degradation of trace organic pollutants in water. Chem. Eng. J. 2020, 385, 123832. [Google Scholar] [CrossRef]
  192. Lu, L.L.; Shan, R.; Shi, Y.Y.; Wang, S.X.; Yuan, H.R. A novel TiO2/biochar composite catalysts for photocatalytic degradation of methyl orange. Chemosphere 2019, 222, 391–398. [Google Scholar] [CrossRef]
  193. Ismael, M. Enhanced photocatalytic hydrogen production and degradation of organic pollutants from Fe (III) doped TiO2 nanoparticles. J. Environ. Chem. Eng. 2020, 8, 103676. [Google Scholar] [CrossRef]
  194. Xu, J.J.; Chen, Y.F.; Dong, Z.Y.; Peng, Y.N.; Yue, S.T.; Huang, H. Strong effect of multi-electron oxygen reduction reaction on photocatalysis through the promotion of interfacial charge transfer. Appl. Catal. B 2019, 252, 41–46. [Google Scholar] [CrossRef]
  195. Singh, K.; Harish, S.; Archana, J.; Navaneethan, M.; Shimomura, M.; Hayakawa, Y. Investigation of Gd-doped mesoporous TiO2 spheres for environmental remediation and energy applications. Appl. Surf. Sci. 2019, 489, 883–892. [Google Scholar] [CrossRef]
  196. Zhang, J.L.; Pang, Z.Y.; Sun, Q.; Chen, X.; Zhu, Y.A.; Li, M.J.; Wang, J.D.; Qiu, H.; Li, X.Q.; Li, Y.G.; et al. TiO2 nanotube array modified with polypyrrole for efficient photoelectrocatalytic decolorization of methylene blue. J. Alloys Compd. 2020, 820, 153128. [Google Scholar] [CrossRef]
  197. Singaram, B.; Jeyaram, J.; Rajendran, R.; Arumugam, P.; Varadharajan, K. Visible light photocatalytic activity of tungsten and fluorine codoped TiO2 nanoparticle for an efficient dye degradation. Ionics 2019, 25, 773–784. [Google Scholar] [CrossRef]
  198. Janbandhu, S.Y.; Joshi, A.; Munishwar, S.R.; Gedam, R.S. CdS/TiO2 heterojunction in glass matrix: Synthesis, characterization, and application as an improved photocatalyst. Appl. Surf. Sci. 2019, 497, 143758. [Google Scholar] [CrossRef]
  199. Liu, C.Q.; Li, X.; Wu, Y.T.; Zhang, L.Y.; Chang, X.J.; Yuan, X.X.; Wang, X.F. Fabrication of multilayer porous structured TiO2-ZrTiO4-SiO2 heterostructure towards enhanced photo-degradation activities. Ceram. Int. 2020, 46, 476–486. [Google Scholar] [CrossRef]
  200. Wang, R.; Shen, J.; Zhang, W.J.; Liu, Q.Q.; Zhang, M.Y.; Zulfiqar; Tang, H. Build-in electric field induced step-scheme TiO2/W18O49 heterojunction for enhanced photocatalytic activity under visible-light irradiation. Ceram. Int. 2020, 46, 23–30. [Google Scholar] [CrossRef]
  201. Thanh, T.L.T.; Thi, L.N.; Dinh, T.T.; Van, N.N. Enhanced Photocatalytic Degradation of Rhodamine B Using C/Fe Co-Doped Titanium Dioxide Coated on Activated Carbon. J. Chem. 2019, 2019, 2949316. [Google Scholar] [CrossRef] [Green Version]
  202. Kumar, A.; Khan, M.; He, J.H.; Lo, I.M.C. Visible-light-driven magnetically recyclable terephthalic acid functionalized g-C3N4/TiO2 heterojunction nanophotocatalyst for enhanced degradation of PPCPs. Appl. Catal. B 2020, 270, 118898. [Google Scholar] [CrossRef]
  203. Chen, M.; Wang, H.H.; Chen, X.Y.; Wang, F.; Qin, X.X.; Zhang, C.B.; He, H. High-performance of Cu-TiO2 for photocatalytic oxidation of formaldehyde under visible light and the mechanism study. Chem. Eng. J. 2020, 390, 124481. [Google Scholar] [CrossRef]
  204. He, L.; Dong, Y.N.; Zheng, Y.N.; Jia, Q.M.; Shan, S.Y.; Zhang, Y.Q. A novel magnetic MIL-101(Fe)/TiO2 composite for photo degradation of tetracycline under solar light. J. Hazard. Mater. 2019, 361, 85–94. [Google Scholar] [CrossRef] [PubMed]
  205. Namshah, K.S.; Mohamed, R.M. WO3-TiO2 nanocomposites for paracetamol degradation under visible light. Appl. Nanosci. 2018, 8, 2021–2030. [Google Scholar] [CrossRef]
  206. Shi, Z.; Xu, P.F.; Shen, X.F.; Zhang, Y.; Luo, L.; Duoerkun, G.; Zhang, L.S. TiO2/MoS2 heterojunctions-decorated carbon fibers with broad-spectrum response as weaveable photocatalyst/photoelectrode. Mater. Res. Bull. 2019, 112, 354–362. [Google Scholar] [CrossRef]
  207. Djouadi, L.; Khalaf, H.; Boukhatem, H.; Boutoumi, H.; Kezzime, A.; Santaballa, J.A.; Canle, M. Degradation of aqueous ketoprofen by heterogeneous photocatalysis using Bi2S3/TiO2-Montmorillonite nanocomposites under simulated solar irradiation. Appl. Clay Sci. 2018, 166, 27–37. [Google Scholar] [CrossRef]
  208. Leal, J.F.; Cruz, S.M.A.; Almeida, B.T.A.; Esteves, V.I.; Marques, P.; Santos, E.B.H. TiO2-rGO nanocomposite as an efficient catalyst to photodegrade formalin in aquaculture’s waters, under solar light. Environ. Sci. Water Res. Technol. 2020, 6, 1018–1027. [Google Scholar] [CrossRef]
  209. Morones-Esquivel, M.M.; Nunez-Nunez, C.M.; Gonzalez-Burciaga, L.A.; Hernandez-Mendoza, J.L.; Osorio-Revilla, G.I.; Proal-Najera, J.B. Kinetics and statistical approach for 2,5-dichlorophenol degradation in short reaction times by solar TiO2/glass photocatalysis. Rev. Mex. Ing. Quim. 2020, 19, 555–568. [Google Scholar] [CrossRef] [Green Version]
  210. Borzyszkowska, A.F.; Pieczynska, A.; Ofiarska, A.; Lisowski, W.; Nikiforow, K.; Siedlecka, E.M. Photocatalytic degradation of 5-fluorouracil in an aqueous environment via Bi-B co-doped TiO2 under artificial sunlight. Int. J. Environ. Sci. Technol. 2020, 17, 2163–2176. [Google Scholar] [CrossRef]
  211. Tbessi, I.; Benito, M.; Molins, E.; Llorca, J.; Touati, A.; Sayadi, S.; Najjar, W. Effect of Ce and Mn co-doping on photocatalytic performance of sol-gel TiO2. Solid State Sci. 2019, 88, 20–28. [Google Scholar] [CrossRef]
  212. Saqlain, S.; Cha, B.J.; Kim, S.Y.; Ahn, T.K.; Park, C.; Oh, J.M.; Jeong, E.C.; Seo, H.O.; Kim, Y.D. Visible light-responsive Fe-loaded TiO2 photocatalysts for total oxidation of acetaldehyde: Fundamental studies towards large-scale production and applications. Appl. Surf. Sci. 2020, 505, 144160. [Google Scholar] [CrossRef]
  213. Cherni, D.; Moussa, N.; Nsib, M.F.; Olivo, A.; Signoretto, M.; Prati, L.; Villa, A. Photocatalytic degradation of ethylbenzene in gas phase over N or NF doped TiO2 catalysts. J. Mater. Sci. Mater. Electron. 2019, 30, 18919–18926. [Google Scholar] [CrossRef]
  214. Jaleh, B.; Rouzbahani, M.G.; Abedi, K.; Azizian, S.; Ebrahimi, H.; Nasrollahzadeh, M.; Varma, R.S. Photocatalytic decomposition of VOCs by AC-TiO2 and EG-TiO2 nanocomposites. Clean Technol. Environ. Policy 2019, 21, 1259–1268. [Google Scholar] [CrossRef]
  215. Kim, J.; Lee, B.K. Enhanced photocatalytic decomposition of VOCs by visible-driven photocatalyst combined Cu-TiO2 and activated carbon fiber. Process Saf. Environ. Prot. 2018, 119, 164–171. [Google Scholar] [CrossRef]
  216. Belet, A.; Wolfs, C.; Mahy, J.G.; Poelman, D.; Vreuls, C.; Gillard, N.; Lambert, S.D. Sol-gel Syntheses of Photocatalysts for the Removal of Pharmaceutical Products in Water. Nanomaterials 2019, 9, 126. [Google Scholar] [CrossRef] [Green Version]
  217. Saif, M.; Aboul-Fotouh, S.M.K.; El-Molla, S.A.; Ibrahim, M.M.; Ismail, L.F.M. Evaluation of the photocatalytic activity of Ln(3+)-TiO2 nanomaterial using fluorescence technique for real wastewater treatment. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 128, 153–162. [Google Scholar] [CrossRef]
  218. Lima, K.V.; Emídio, E.S.; Pupo Nogueira, R.F.; do Vasconcelos, N.S.L.; Araújo, A.B. Application of a stable Ag/TiO2 film in the simultaneous photodegradation of hormones. J. Chem. Technol. Biotechnol. 2020, 8. [Google Scholar] [CrossRef]
  219. Peng, X.M.; Wang, M.; Hu, F.P.; Qiu, F.X.; Dai, H.L.; Cao, Z. Facile fabrication of hollow biochar carbon-doped TiO2/CuO composites for the photocatalytic degradation of ammonia nitrogen from aqueous solution. J. Alloys Compd. 2019, 770, 1055–1063. [Google Scholar] [CrossRef]
  220. Li, S.M.; Tan, J.; Jiang, Z.J.; Wang, J.; Li, Z.Q. MOF-derived bimetallic Fe-Ni-P nanotubes with tunable compositions for dye-sensitized photocatalytic H2 and O2 production. Chem. Eng. J. 2020, 384, 123354. [Google Scholar] [CrossRef]
  221. Saad, A.M.; Abukhadra, M.R.; Abdel-Kader Ahmed, S.; Elzanaty, A.M.; Mady, A.H.; Betiha, M.A.; Shim, J.-J.; Rabie, A.M. Photocatalytic degradation of malachite green dye using chitosan supported ZnO and Ce-ZnO nano-flowers under visible light. J. Environ. Manag. 2020, 258, 110043. [Google Scholar] [CrossRef] [PubMed]
  222. Espino-Estevez, M.R.; Fernandez-Rodriguez, C.; Gonzalez-Diaz, O.M.; Arana, J.; Espinos, J.P.; Ortega-Mendez, J.A.; Dona-Rodriguez, J.M. Effect of TiO2-Pd and TiO2-Ag on the photocatalytic oxidation of diclofenac, isoproturon and phenol. Chem. Eng. J. 2016, 298, 82–95. [Google Scholar] [CrossRef]
  223. Xu, H.F.; Li, G.; Liu, N.; Zhu, K.R.; Zhu, G.; Jin, S.W. Ag @ hierarchical TiO2 core-shell nanostructures for enhanced photocatalysis. Mater. Lett. 2015, 142, 324–327. [Google Scholar] [CrossRef]
  224. Ji, L.J.; Liu, X.; Xu, T.; Gong, M.D.; Zhou, S. Preparation and photocatalytic properties of carbon/carbon-doped TiO2 double-layer hollow microspheres. J. Sol-Gel Sci. Technol. 2020, 93, 380–390. [Google Scholar] [CrossRef]
  225. Fu, Z.T.; Zhang, S.; Fu, Z.X. Preparation of Multicycle GO/TiO2 Composite Photocatalyst and Study on Degradation of Methylene Blue Synthetic Wastewater. Appl. Sci. (Basel) 2019, 9, 3282. [Google Scholar] [CrossRef] [Green Version]
  226. He, X.; Wang, X.C.; Cheng, B.Z.; Cao, S.; Zhou, J.F. Application of Mn-doped Mesoporous TiO2 in Tannery Wastewater Treatment. J. Soc. Leather Technol. Chem. 2019, 103, 318–322. [Google Scholar]
  227. Bordes, M.C.; Vicent, M.; Moreno, R.; Garcia-Montano, J.; Serra, A.; Sanchez, E. Application of plasma-sprayed TiO2 coatings for industrial (tannery) wastewater treatment. Ceram Int. 2015, 41, 14468–14474. [Google Scholar] [CrossRef] [Green Version]
  228. Kim, K.H.; Kabir, E.; Jahan, S.A. Exposure to pesticides and the associated human health effects. Sci. Total Environ. 2017, 575, 525–535. [Google Scholar] [CrossRef]
  229. Fiorenza, R.; Di Mauro, A.; Cantarella, M.; Iaria, C.; Scalisi, E.M.; Brundo, M.V.; Gulino, A.; Spitaleri, L.; Nicotra, G.; Dattilo, S.; et al. Preferential removal of pesticides from water by molecular imprinting on TiO2 photocatalysts. Chem. Eng. J. 2020, 379, 122309. [Google Scholar] [CrossRef]
  230. Abdennouri, M.; Baalala, M.; Galadi, A.; El Makhfouk, M.; Bensitel, M.; Nohair, K.; Sadiq, M.; Boussaoud, A.; Barka, N. Photocatalytic degradation of pesticides by titanium dioxide and titanium pillared purified clays. Arab. J. Chem. 2016, 9, S313–S318. [Google Scholar] [CrossRef] [Green Version]
  231. Nasseh, N.; Taghavi, L.; Barikbin, B.; Nasseri, M.A. Synthesis and characterizations of a novel FeNi3/SiO2/CuS magnetic nanocomposite for photocatalytic degradation of tetracycline in simulated wastewater. J. Clean. Prod. 2018, 179, 42–54. [Google Scholar] [CrossRef]
  232. de Oliveira, C.P.M.; Viana, M.M.; Amaral, M.C.S. Coupling photocatalytic degradation using a green TiO2 catalyst to membrane bioreactor for petroleum refinery wastewater reclamation. Comp. Immunol. Microbiol. Infect. Dis. 2019, 68, 101403. [Google Scholar] [CrossRef]
  233. Ani, I.J.; Akpan, U.G.; Olutoye, M.A.; Hameed, B.H. Photocatalytic degradation of pollutants in petroleum refinery wastewater by TiO2- and ZnO-based photocatalysts: Recent development. J. Clean. Prod. 2018, 205, 930–954. [Google Scholar] [CrossRef]
  234. Shivaraju, H.P.; Muzakkira, N.; Shahmoradi, B. Photocatalytic treatment of oil and grease spills in wastewater using coated N-doped TiO2 polyscales under sunlight as an alternative driving energy. Int. J. Environ. Sci. Technol. 2016, 13, 2293–2302. [Google Scholar] [CrossRef] [Green Version]
  235. Andronic, L.; Isac, L.; Miralles-Cuevas, S.; Visa, M.; Oller, I.; Duta, A.; Malato, S. Pilot-plant evaluation of TiO2 and TiO2-based hybrid photocatalysts for solar treatment of polluted water. J. Hazard. Mater. 2016, 320, 469–478. [Google Scholar] [CrossRef]
  236. Solis-Casados, D.A.; Escobar-Alarcon, L.; Gomez-Olivan, L.M.; Haro-Poniatowski, E.; Klimova, T. Photodegradation of pharmaceutical drugs using Sn-modified TiO2 powders under visible light irradiation. Fuel 2017, 198, 3–10. [Google Scholar] [CrossRef] [Green Version]
  237. Malakootian, M.; Olama, N.; Malakootian, M.; Nasiri, A. Photocatalytic degradation of metronidazole from aquatic solution by TiO2-doped Fe3+ nano-photocatalyst. Int. J. Environ. Sci. Technol. 2019, 16, 4275–4284. [Google Scholar] [CrossRef]
  238. Lacerda, J.A.S.; Macedo, A.M.; Teixeira, R.I.; Simoes, G.; Ribeiro, E.S.; Forero, J.S.B.; Correa, R.J. TiO2 Decorated Sand Grains for Photodegradation of Pollutants: Methylene Blue and Ciprofloxacin Study. J. Braz. Chem. Soc. 2020, 31, 201–210. [Google Scholar] [CrossRef]
  239. Hou, C.T.; Xie, J.Q.; Yang, H.L.; Chen, S.M.; Liu, H.L. Preparation of Cu2O@TiOF2/TiO2 and its photocatalytic degradation of tetracycline hydrochloride wastewater. RSC Adv. 2019, 9, 37911–37918. [Google Scholar] [CrossRef] [Green Version]
  240. Li, S.P.; Chang, L.B.; Peng, J.H.; Gao, J.Z.; Lu, J.B.; Zhang, F.C.; Zhu, G.Q.; Hojamberdiev, M. Bi-0 nanoparticle loaded on Bi3+-doped ZnWO4 nanorods with oxygen vacancies for enhanced photocatalytic NO removal. J. Alloys Compd. 2020, 818, 152837. [Google Scholar] [CrossRef]
  241. Rao, F.; Zhu, G.Q.; Hojamberdiev, M.; Zhang, W.B.; Li, S.P.; Gao, J.Z.; Zhang, F.C.; Huang, Y.H.; Huang, Y. Uniform Zn2+-Doped BiOI Microspheres Assembled by Ultrathin Nanosheets with Tunable Oxygen Vacancies for Super-Stable Removal of NO. J. Phys. Chem. C 2019, 123, 16268–16280. [Google Scholar] [CrossRef]
  242. Zhang, G.X.; Sun, Z.M.; Duan, Y.W.; Ma, R.X.; Zheng, S.L. Synthesis of nano-TiO2/diatomite composite and its photocatalytic degradation of gaseous formaldehyde. Appl. Surf. Sci. 2017, 412, 105–112. [Google Scholar] [CrossRef]
  243. Qin, Z.B.; Zhang, W.T.; Qian, G.P.; Wu, X.L.; Li, Y.J.M.R.I. The effects of different ways of adding nano-TiO2 to concrete on the degradation performance of NO2. Mater. Res. Innov. 2015, 19, S10–S148. [Google Scholar] [CrossRef]
  244. Zhang, J.H.; Hu, Y.; Qin, J.X.; Yang, Z.X.; Fu, M.L. TiO2-UiO-66-NH2 nanocomposites as efficient photocatalysts for the oxidation of VOCs. Chem. Eng. J. 2020, 385, 123814. [Google Scholar] [CrossRef]
  245. Wang, X.G.; Sun, M.H.; Murugananthan, M.; Zhang, Y.R.; Zhang, L.Z. Electrochemically self-doped WO3/TiO2 nanotubes for photocatalytic degradation of volatile organic compounds. Appl. Catal. B 2020, 260, 118205. [Google Scholar] [CrossRef]
  246. Lai, M.; Zhao, J.; Chen, Q.C.; Feng, S.J.; Bai, Y.J.; Li, Y.X.; Wang, C.Y. Photocatalytic toluene degradation over Bi-decorated TiO2: Promoted O2 supply to catalyst’s surface by metallic Bi. Catal. Today 2019, 335, 372–380. [Google Scholar] [CrossRef]
  247. Rao, Z.P.; Xie, X.F.; Wang, X.; Mahmood, A.; Tong, S.R.; Ge, M.F.; Sun, J. Defect Chemistry of Er3+-Doped TiO2 and Its Photocatalytic Activity for the Degradation of Flowing Gas-Phase VOCs. J. Phys. Chem. C 2019, 123, 12321–12334. [Google Scholar] [CrossRef]
  248. Ali, N.; Bilal, M.; Khan, A.; Ali, F.; Iqbal, H.M.N. Effective exploitation of anionic, nonionic, and nanoparticle-stabilized surfactant foams for petroleum hydrocarbon contaminated soil remediation. Sci. Total Environ. 2020, 704, 135391. [Google Scholar] [CrossRef]
  249. Peikam, E.N.; Jalali, M. Application of three nanoparticles (Al2O3, SiO2 and TiO2) for metal-contaminated soil remediation (measuring and modeling). Int. J. Environ. Sci. Technol. 2019, 16, 7207–7220. [Google Scholar] [CrossRef]
  250. Wang, L.; Li, X.; Tsang, D.C.W.; Jin, F.; Hou, D. Green remediation of Cd and Hg contaminated soil using humic acid modified montmorillonite: Immobilization performance under accelerated ageing conditions. J. Hazard. Mater. 2020, 387, 122005. [Google Scholar] [CrossRef]
  251. Khalid, S.; Shahid, M.; Murtaza, B.; Bibi, I.; Natasha; Asif Naeem, M.; Niazi, N.K. A critical review of different factors governing the fate of pesticides in soil under biochar application. Sci. Total Environ. 2020, 711, 134645. [Google Scholar] [CrossRef]
  252. Eker, G.; Hatipoglu, M. Effect of UV wavelength, temperature and photocatalyst on the removal of PAHs from industrial soil with photodegradation applications. Environ. Technol. 2019, 40, 3793–3803. [Google Scholar] [CrossRef]
  253. Rachna; Rani, M.; Shanker, U. Degradation of tricyclic polyaromatic hydrocarbons in water, soil and river sediment with a novel TiO2 based heterogeneous nanocomposite. J. Environ. Manag. 2019, 248, 109340. [Google Scholar] [CrossRef]
  254. Zyoud, A.; Ateeq, M.; Helal, M.H.; Zyoud, S.H.; Hilal, H.S. Photocatalytic degradation of phenazopyridine contaminant in soil with direct solar light. Environ. Technol. 2019, 40, 2928–2939. [Google Scholar] [CrossRef]
  255. Xu, Q.; Huang, Z.; Ji, S.T.; Zhou, J.; Shi, R.Z.; Shi, W.Y. Cu2O nanoparticles grafting onto PLA fibers via electron beam irradiation: Bifunctional composite fibers with enhanced photocatalytic of organic pollutants in aqueous and soil systems. J. Radioanal. Nucl. Chem. 2020, 323, 253–261. [Google Scholar] [CrossRef]
  256. Theerakarunwong, C.D.; Phanichphant, S. Visible-Light-Induced Photocatalytic Degradation of PAH-Contaminated Soil and Their Pathways by Fe-Doped TiO2 Nanocatalyst. Water Air Soil Poll. 2018, 229, 291. [Google Scholar] [CrossRef]
  257. Kuang, X.; Shao, J.; Peng, L.; Song, H.; Wei, X.; Luo, S.; Gu, J.-D. Nano-TiO2 enhances the adsorption of Cd(II) on biological soil crusts under mildly acidic conditions. J. Contam. Hydrol. 2020, 229, 103583. [Google Scholar] [CrossRef]
  258. Yang, Y.; Javed, H.; Zhang, D.N.; Li, D.Y.; Kamath, R.; McVey, K.; Sra, K.; Alvarez, P.J.J. Merits and limitations of TiO2-based photocatalytic pretreatment of soils impacted by crude oil for expediting bioremediation. Front. Chem. Sci. Eng. 2017, 11, 387–394. [Google Scholar] [CrossRef]
Figure 1. The schematic diagram of non-metallic materials loading TiO2.
Figure 1. The schematic diagram of non-metallic materials loading TiO2.
Catalysts 10 00804 g001
Figure 2. The mechanism of non-metallic element doping modification system.
Figure 2. The mechanism of non-metallic element doping modification system.
Catalysts 10 00804 g002
Figure 3. The mechanism of precious metal materials deposition.
Figure 3. The mechanism of precious metal materials deposition.
Catalysts 10 00804 g003
Figure 4. The mechanism of metal ion doping TiO2.
Figure 4. The mechanism of metal ion doping TiO2.
Catalysts 10 00804 g004
Figure 5. The band arrangement and electron migration mechanism of different heterojunctions. (a) Type-I heterojunction; (b) Type-II heterojunction; (c) Type-III heterojunction; (d) Z-scheme heterojunction.
Figure 5. The band arrangement and electron migration mechanism of different heterojunctions. (a) Type-I heterojunction; (b) Type-II heterojunction; (c) Type-III heterojunction; (d) Z-scheme heterojunction.
Catalysts 10 00804 g005
Figure 6. The mechanism of the dye sensitization process.
Figure 6. The mechanism of the dye sensitization process.
Catalysts 10 00804 g006
Table 1. The published data of different photocatalysts and their performance.
Table 1. The published data of different photocatalysts and their performance.
TypePhotocatalystsLight SourceTarget PollutantDegradation RateRef.
TiO2MoS2/MoO3/TiO2300 W Xe lamprhodamine B95%[6]
Yb, Er, Ce-codoped TiO2Xe lamp4-chlorophenol95%[7]
TiO2@SiO2 composites500 W mercury lampmethyl orange99%[8]
SrTiO3La- SrTiO3500 W Xe lampCr6+84%[9]
Ag3PO4/PANI/Cr: SrTiO3300 W Xe lampphenol99%[10]
Ag, Cr-SrTiO3500 W Xe lampmethyl orange98%[11]
ZnOCu-ZnOblue light lampOrange II70%[12]
ZnOsunlightmethylene blue90%[13]
Sr-ZnOblack lightmethylene blue99%[14]
WO3WO31500 W Xe lampN, N-diethyl-meta- toluamide60%[15]
WO3@Cu@PDI300 W Xe lamptetracycline hydrochloride85%[16]
NiO-WO3150 W tungsten lampeosin yellow95%[17]
ZrO2Ce, Er-codoped ZrO2halogen lamprhodamine B 92%[18]
Co3O4-ZrO2visible lightcyanide100%[19]
Cu- ZrO2Visible lightmethyl orange98%[20]
g-C3N4Ag-P-codoped g-C3N48 W visible lampssulfamethoxazole99%[21]
AgI/LaFeO3/g-C3N4500 W Xe lampnorfloxacin95%[22]
CdS/g-C3N42500 W Xe lamprhodamine B96%[23]
Table 2. The photocatalytic performance of the co-doped TiO2 in treating pollutants.
Table 2. The photocatalytic performance of the co-doped TiO2 in treating pollutants.
Doping ElementsCrystal Phases of TiO2Light Source and Reaction TimeTarget PollutantDegradation RateRef.
Ni, CranataseSunlight 90 minmethylene blue96%[153]
Cu, CoanataseLED 300 minacetaldehyde99%[160]
Ag, V-40 W white light bulbs 180 minhexane gas butyl acetate gas94% 96%[161]
N, Cuanatase200 W Xe lamp 60 minsulfamethoxazole99%[162]
Fe, Ianatasevisible light 60 mingaseous benzene59%[80]
Mn, Nanatase, rutile, wurtziteLED 40 minQuinalphos 2-chlorophenol 92% 88%[163]
N, AganataseLED 360 minmethylene blue99%[164]
Ag, Pd, Nanatasemercury vapor lamp 120 minmalachite green methylene blue mongo red75% 92% 62%[165]
C, Nanatasesimulated sunlight 420 min4-nitrophenol87%[166]
N, Fanatase500 W Xe lamp 150 minmethylene blue89%[167]
Si, Nanatase500 W Xe lamp 180 minRhodamine B86%[168]
C, Nanatase and rutile300 W Xe lamp 150 minphenol92%[169]
Table 3. The performance of the modified TiO2 in treating pollutants.
Table 3. The performance of the modified TiO2 in treating pollutants.
PhotocatalystsCrystal Phases of TiO2Light Source and Reaction TimeTarget PollutantDegradation RateRef.
TiO2/biocharanatase500 W mercury lamp 150 minmethyl orange97%[192]
Fe3+-TiO2 nanoparticlesanatase150 W Xe lamp 240 min4-chlorophenol ethyl orange65% 95%[193]
TiO2-Fe-porphyrin-conjugated microporous polymersanataseXe lamp 90 minmethyl orange96%[194]
Gd-TiO2anatasevisible light 93 minmethylene blue28%[195]
polypyrrole@TiO2anatase and rutile250 W mercury lamp 60 minmethylene blue25%[196]
W, F-TiO2anatase500 W halogen lamp 180 minmethylene blue96%[197]
sodium borosilicate glass SiO2-B2O3-Na2O- ZnO with CdS and TiO2anataseSunlight 300 minindigo carmine dye92%[198]
TiO2-ZrTiO4-SiO2anatase300 W Xe lamp 90 minrhodamine B95%[199]
TiO2-W18O49anatasevisible light 60 minrhodamine B82%[200]
C/Fe-TiO2 coated on activated carbonanatase36 W compact light 140 minrhoda mine B99%[201]
terephthalic acid functionalized g-C3N4/TiO2/Fe3O4@SiO2anatase8 W compact fluorescent lamps 120 minibuprofen97%[202]
Cu-TiO2anatase500 W Xe lamp 140 minformaldehyde100%[203]
MIL-101(Fe)/TiO2anataseSunlight 30 mintetracycline93%[204]
WO3/TiO2anatase500 W Xe lamp 60 minparacetamol100%[205]
MoS2/TiO2/Carbon Fiberrutilevisible light 60 mintetracycline93%[206]
Bi2S3/TiO2/Montmorilloniteanatasemercury vapor lamp 120 minketoprofen90%[207]
TiO2-reduced graphene oxide (TiO2-rGO)anatasesimulated sunlight 90 minformalin98%[208]
TiO2/glassanataseSunlight 30 min2,5-dichlorophenol95%[209]
Bi, B-TiO2anataseXe lamp 90 min5-fluorouracil100%[210]
Ce, Mn- TiO2anatase30 W ultraviolet lamp 240 mindiclofenac94%[211]
Fe-TiO2anatase and rutile visible light 1050 minacetaldehyde65%[212]
N, F-TiO2anatasemercury vapor lamp 180 minethylbenzene33%[213]
activated carbon-TiO2anatase and rutileUV light 20 mintoluene99%[214]
Eosin Y- TiO2anatasevisible light 180 minacetaminophen diclofenac71% 83%[185]
Cu-TiO2 combine with activated carbon fiberanatase and rutilefluorescent lamp 180 minBenzene toluene81% 98%[215]

Share and Cite

MDPI and ACS Style

Li, R.; Li, T.; Zhou, Q. Impact of Titanium Dioxide (TiO2) Modification on Its Application to Pollution Treatment—A Review. Catalysts 2020, 10, 804. https://doi.org/10.3390/catal10070804

AMA Style

Li R, Li T, Zhou Q. Impact of Titanium Dioxide (TiO2) Modification on Its Application to Pollution Treatment—A Review. Catalysts. 2020; 10(7):804. https://doi.org/10.3390/catal10070804

Chicago/Turabian Style

Li, Ruixiang, Tian Li, and Qixing Zhou. 2020. "Impact of Titanium Dioxide (TiO2) Modification on Its Application to Pollution Treatment—A Review" Catalysts 10, no. 7: 804. https://doi.org/10.3390/catal10070804

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